You are on page 1of 15

Journal of Petroleum Exploration and Production Technology (2023) 13:2109–2123

https://doi.org/10.1007/s13202-023-01651-0

ORIGINAL PAPER-PRODUCTION ENGINEERING

Optimization of surfactant‑polymer flooding for enhanced oil recovery


M. Elmuzafar Ahmed1 · Abdullah S. Sultan1 · Abdulkarim Al‑Sofi2 · Hasan S. Al‑Hashim1

Received: 28 December 2021 / Accepted: 15 May 2023 / Published online: 15 June 2023
© The Author(s) 2023

Abstract
Chemical enhanced oil recovery applications continue to face a variety of obstacles, particularly in high saline and high-
temperature reservoirs, in addition to high chemical prices. This issue creates difficulty in developing optimal recipes that
can withstand these extreme circumstances and so achieve maximal hydrocarbon recovery at the lowest feasible cost. The
usefulness of surfactant polymer (SP) in mobilizing oil and increasing sweep efficiency in carbonate rocks is assessed in
this article. A thermo-viscosifying polymer and an acrylamido tertiary butyl sulfonate (ATBS)/acrylamide (AM) copolymer
were employed. Surfactants of various grades of amphoteric carboxybetain are used. These potential chemicals were chosen
after a thorough study of previous research, which included long-term thermal stability, fluid rheology, interfacial tension,
adsorption, and microfluidic tests. The contact angles were measured using a captive drop analyzer at high pressure and
high temperature. The core-flooding experiments for slug size and injection sequence optimization were carried out using
12-inch long and 1.5-inch diameter limestone cores. For two weeks, the samples were aged. The trials were carried out at
90 °C. The seawater (SW) salinity utilized in the injection was 57,000 ppm. The findings highlighted the importance of
surfactant-polymer interactions in wettability and fluid rheology. The best chemical combination was carboxybetaine (0.05
wt%) and ATBS/AM (0.25 wt%) which recovered 31.29% of the residual oil saturation (ROS), or 11.63% of the original oil
in place (OIIP). The optimal slug size was 3.5 PV, which recovered 34.21% of the ROS and 17.05% of the OIIP. The optimum
injection sequence was the co-injection of surfactant and polymer SW-S1P1-SW, which extracted 31.29% of the ROS and
11.63% of the OIIP. The recoveries were discovered to be related to the slug’s size. The chemical injection sequence was
critical to the eventual oil recovery. Among the other sequences, SW-SP-SW had the highest recovery (SW-P-S-SW, SW-S-
SW-P-SW, and SW-P-SW-S-SW). This is thought to be owing to the compounds' synergistic impact. We found that there is
no systematic optimization process that combines the effect of chemicals, slug size, and sequence in one study, which gave
us the motivation to cover the research gap.

Keywords Chemical EOR · SP flooding · Carbonate core flooding · Contact angle · Recovery optimization
Abbreviations IFT Interfacial tension
ATBS Acrylamido tertiary butyl sulfonate OIIP Oil initially in place
AM Acrylamide ROS Residual oil saturation
ASP Alkali-surfactant-polymer SW Seawater
CSNs Colloidal silica nanoparticles SP Surfactant-polymer
CMC Critical micelle concentration TVP Thermoviscofying polymer
CEOR Chemical-enhanced-oil-recovery
List of symbols
CA Contact angle
g Gravitational acceleration = 9.8 m/s2
DHR Discovery Hybrid Rheometer
k Permeability (md)
HTHS High-temperature high salinity
Kro Oil relative permeability
Kr w Water relative permeability
NB Bond number
* Abdullah S. Sultan
sultanas@kfupm.edu.sa Nca Capillary number
NM Mobility number
1
Department of Petroleum Engineering, King Fahd University NT Trapping number
of Petroleum and Minerals, Dhahran, Saudi Arabia v Velocity
2
EXPEC ARC, Saudi Aramco, Dhahran, Saudi Arabia

13
Vol.:(0123456789)
2110 Journal of Petroleum Exploration and Production Technology (2023) 13:2109–2123

ΔPcap Capillary forces its attention shifted toward carbonate reservoirs due to the
ΔPgrav Gravitational forces increasing number of maturing carbonate fields (Al-Hashim
ΔPvis Viscous forces et al. 2005; Azad and Sultan 2014; Han et al. 2013; Levitt
λo Oil mobility et al. 2013; Wang et al. 2015).
λw Water mobility SP flooding is more feasible than individual flooding of
μo Oil viscosity surfactant or polymer and ASP flooding by combining the
μw Oil viscosity benefits of IFT reduction and mobility control and by avoid-
σ Interfacial tension ing operational and precipitation problems associated with
ρ Density (g/cc) ASP flooding (Luo et al. 2013). Ding et al. did a compara-
tive study between SP flooding and AP flooding in term of
IFT, emulsification, and recovery and found that ultra-low
Introduction IFT and oil-in-water dominates SP while water-in-oil (W/O)
emulsification dominates the AP. AP is better than SP in
Chemical EOR technologies have been used for decades and terms of heavy oil recovery, however, SP can be improved
their development is increasing rapidly due to the massive by introducing foam to assist the ultra-low IFT (Ding et al.
need for hydrocarbons in the world and because most of 2020).
the reservoirs have reached the tertiary recovery phase. The The amount of chemicals used in CEOR is of great
main reason for using surfactants is to lower the interfacial importance which leads to the need for concentration and
tension (IFT) between water and oil. While the use of poly- slug size optimization; the optimum slug we are looking for
mer is generally linked with surfactant for mobility control is the slug that gives the highest possible recovery with the
to ensure the best sweep efficiency. Additionally, viscoelastic lowest cost.
surfactants can be used as they can control mobility as well The rock wettability affects the EOR process by influenc-
as reduce the IFT (Ahmed et al. 2022; Azad et al. 2014; ing the oil recovery, even though oil-wet rock will enhance
Elmuzafar Ahmed et al. 2023). the film drainage (Al-Hashim et al. 2005; Omran et al.
Polymer flooding (P) for enhanced oil recovery (EOR) is 2020). The highest overall oil recovery will be achieved
a commonly utilized technique to enhance the effectiveness under water-wet conditions (Zhu et al. 2013). Therefore, it
of oil displacement by reducing the mobility ratio between is very important to restore the original wettability before
the displacing and displaced fluids. However, this technique starting a core-flooding experiment to avoid an optimistic
does not have any impact on the residual oil, which requires recovery.
high capillary forces to be mobilized. In order to overcome High salinity causes chemical precipitation due to
this limitation, surfactants are added to the process, which the presence of divalent cations such as ­Ca++ and ­Mg++
can reduce the interfacial tension (IFT) between oil and (Mohammadi et al. 2008; Tabary et al. 2013) as well as
water, increase the ratio of viscous forces to capillary forces decreasing the polymer viscosity and elasticity (Dang et al.
(capillary number), and hence improve the technique, known 2013). On the other hand, high temperature affects the
as SP flooding. chemical stability and leads the polymer to degrade and act
In carbonate reservoirs, particularly in the Middle East, inefficiently (Azad and Sultan 2014; Han et al. 2013). To
CEOR process still faces challenges such as high salin- solve this problem a comprehensive study of the chemical
ity, high temperature, and reservoir heterogeneity which was done which include thermal stability, fluid rheology,
adversely affect recovery and efficiency. For instance, high micro-emulsion stability, and IFT under these harsh condi-
temperatures cause hydrolysis that affects the polymer's sta- tions (Kamal et al. 2014; Kamal, Sultan, and Hussein 2015;
bility, whereas the surfactants may precipitate in high salin- Kamal, Sultan, Al-Mubaiyedh, and Hussein 2015; Kamal,
ity mediums, making it difficult to screen and choose the Sultan, Al-Mubaiyedh, Hussein et al. 2015). Then the best
best chemicals to use. Additionally, carbonate reservoirs are candidates should be tested by core flooding to choose the
known to have low permeability, with fractures possessing one that gives the highest recovery. Besides, the occurrence
high permeability, leading to channeling and reduced sweep of monovalent cations such as N ­ a+ will decrease the pH
efficiency. For higher permeability formation it is preferred of the solution due to the ion exchange with the ­H+on the
to use higher polymer molecular weight as it is found to rock surface. Therefore the pH drop below the isoelectric
increase the sweep efficiency while for lower permeability point (Somasundaran and Hanna 1979) will result in higher
ones, there is an optimum value that should not be exceeded adsorption of chemicals on carbonate rock especially anionic
(Fang et al. 2022). surfactants (Glover et al. 1979; Journal et al. 2008; Lu and
The early studies of CEOR flooding focused mainly on Pope 2017; Somasundaran and Hanna 1979).
sandstone reservoirs (Arihara et al. 1999; Chiou and Kel- Studying the flooding sequence is important to iden-
lerhals 1981; Osterloh and Jante 1992). However, recently tify the best sequence that minimizes the fingering and

13
Journal of Petroleum Exploration and Production Technology (2023) 13:2109–2123 2111

maximizes the recovery. The literature reported many Capillary number ­(Nca) is the ratio of viscous force to
sequences such as polymer pre-flush to minimize the sur- the capillary force acting on the displaced fluid, we can
factant adsorption (Wang et al. 2015) and SP co-injection increase it by decreasing the IFT using a surfactant or by
followed by polymer injection (Zhenquan et al. 2013) and increasing the polymer viscosity.
other sequences.
𝚫𝐏𝐯𝐢𝐬 𝝁v
Dead oil is very common in lab experiments due to Nca = = (2)
the difficulty of handling live oil and maintaining it in its 𝚫Pcap 𝝈
reservoir composition, however, gases such as methane
where:
and ­CO2 can be added to dead oil in known ratios under
Nca : Capillary number;
pressure to achieve live oil for core flooding purposes (Sui
ΔPvis : viscous forces;
et al. 2019).
ΔPcap : Capillary forces;v : velocity;σ: interfacial tension.
When chemicals pass through reservoir rock, adsorption on
Bond number (­ N B) is the ratio between hydrostatic
the surface of the rock can affect the feasibility of the process.
forces and capillary forces. So, in SP, we can only
Therefore, it is an important element to be considered in the
increase the Bond number by decreasing the IFT through
design to select chemicals that are economical and technically
surfactant because the increment by increasing density
achieve the goal. Some additives have proven to be effective
is negligible.
in reducing surfactant adsorption such as Colloidal silica
nanoparticles (CSNs). It is confirmed with SDS surfactant
( )
k
ΔPgrav Δ𝜌g
and reached up to 61% reduction in adsorption by using 25% NB = =
𝜙
(3)
CSNS with 2500 ppm SDS (Kesarwani et al. 2021). ΔPcap 𝜎
In addition, hydrodynamic dispersion can occur when
fluids mix while flowing through a permeable formation, where:
resulting in a dilution of the slugs and decreased effective- NB: Bond number;
ness. Another important factor to consider is viscous fin- ΔPgrav : gravitational forces;ρ : density (g/cc);g : gravita-
gering, which occurs when low-viscosity fluid displaces tional acceleration = 9.8 m/s2;k : Permeability (md).
higher-viscosity fluid. In the SP-EOR process, the differ- At constant Bond number, hydrocarbon recovery
ence in molecular weight of the polymers may cause slow increases with the increase of Capillary number up to a
movement in the pore system and result in chromatographic specific limit and then decreases dramatically because of
separation. It is crucial to investigate and evaluate adsorption the higher viscous forces and the flow stability (Tabrizy
and chromatographic separation through core flood experi- 2014).
ments in the lab before scaling up the process to the field to Trapping number is a combination of Bond number and
avoid potential failure. Capillary number in a way that can sufficiently address the
Enhanced oil recovery (EOR) is implemented to recover combined effect of capillary, viscous, and buoyancy forces
residual oil held up by capillary and viscous forces in porous in three dimensions.
media. This relationship can be expressed using dimension- √
less numbers such as the Mobility number, capillary number, NT = N2ca + 2Nca NB sin 𝜶 + N2B (4)
Bond number, and Trapping number. The Mobility num-
ber represents the ratio of the mobility of the displacing where:
phase to that of the displaced phase. Since oil has a lower For horizontal flow 𝛼 = 0°, for vertical flow = 90°.
mobility number, the viscosity of the displacing phase can NT:Trapping number.
be increased using polymer, or the viscosity of the oil can As far as we know and based on our survey of the lit-
be reduced using heat. erature, there is no established optimization process that
integrates the impact of chemicals, slug size, and sequence
in a single study. Although some of these effects have been
[ ]
K rw
𝝀
investigated separately or addressed within larger projects
𝝁w
NM = w = [ ] (1)
𝝀o Kro
with limited emphasis. This has motivated us to address
𝝁o
this research gap and explore this area further.
where:
NM : Mobility Number; λo : Oil mobility; λw : Water
mobility;
Kro : Oil relative permeability;μo : Oil viscosity;
Kr w : Water relative permeability;μw : Water viscosity.

13
2112 Journal of Petroleum Exploration and Production Technology (2023) 13:2109–2123

Experimental section Table 2  Seawater composition Salt Concen-


tration
Materials (mg/L)

NaCl 41,041
Synthetic formation brine with total dissolved solids (TDS) CaCl2 .2H2 O 2384
of 229,870 ppm was used to saturate the cores initially then MgCl2 .6H2 O 17,645
displaced with oil to establish the initial water saturation. Na2 SO4 6343
The composition of the formation brine is given in Table 1. NaHCO3 165
Synthetic seawater similar to Arabian Gulf water with TDS 57,612
a TDS of 57,612 ppm was used for water and chemical
flooding as well as for the bulk fluid of the contact angle
experiment, its composition is in Table 2.
The brines were prepared carefully using deionized both of them received in a solution with an active con-
water. The weights of the salts were measured using an tent of 40%. Both surfactants have similar IFT values
accurate weight balance of up to 4 decimal digits. They and their critical micelle concentration (CMC) is around
have been stirred for at least three hours to ensure com- (0.025 wt%). The optimum surfactant concentration was
plete dissolution, then filtered with 0.5-micron filter chosen to be 0.05 wt% to compensate for any probable
paper. loss of surfactant due to retention or adsorption. We kept
The crude oil used in this research was medium oil it as low as possible because there is no point in using
with 24.6 API and 0.89987 g/cc density measured at additional surfactant concentration and increasing the
room temperature. Its viscosity at 90 °C is 3 cp and the cost of chemicals without an expected reduction in IFT
density at that temperature is 0.77748 g/cc. It has been as shown in Fig. 2.
analyzed using Saturates, Aromatic, Resin, and Asphal- Those chemicals will be used with the following
tene (SARA) test as shown in (Table 3) which measures formulations in the contact angle, rheology, and core-
the mass percentage of each component in the crude oil flooding Table 4.
mixture.
We used two types of polymers and two types of sur-
factants in this work. Acrylamido tertiary butyl sulfonate Rheology
(ATBS)/acrylamide (AM) copolymer produced by SNF
FLOERGER. It has 8 million Dalton molecular weight The seawater-based solutions, as previously described,
with the structure in Fig. 1a and it will be referred to as are prepared by first placing the solvent in a wide beaker
(P1). Thermoviscosifying polymer obtained from Hengju and using a magnetic stirrer to create a vortex. The poly-
Polymer Co. laboratories, its molecular weight is 7.08 mer powder is then gradually added to the top of the
million Dalton, it contained 8% thermos sensitive mono- vortex to facilitate dissolution. If a surfactant is present,
mer and has 3% degree of hydrolysis with the structure the surfactant solution is added to create the vortex.
in Fig. 1b and it will be referred to as (P2). The equipment employed for this process is the Dis-
While the surfactants were Amphoteric surfactant covery Hybrid Rheometer (DHR3) provided by TA
SS-880 Carboxybetaine provided by Oil Chem Technol- Instruments Trios, with the use of a concentric cylinder
ogy, USA which will be referred to later in the paper geometry. The resulting data were fitted to established
as (S1). The other surfactant (S2) was Amphoteric sur- models such as the Carreau-Yasuda.
factant SS-885 Carboxybetaine from the same source,

Table 1  Formation brine Table 3  Saturates, Aromatic, Resin, and Asphaltene (SARA) test
Salt Concen-
composition analysis for Crude oil
tration
(mg/L) Components Composi-
tion (wt%)
NaCl 165,546
CaCl2 .2H2 O 69,723 Saturates (Alkanes) 26.64
MgCl2 .6H2 O 18,724 Asphaltene 6.90
Na2 SO4 2395 Resin 5.54
NaHCO3 529 Poly aromatic hydrocarbons 60.92
TDS 229,870 Total 100

13
Journal of Petroleum Exploration and Production Technology (2023) 13:2109–2123 2113

Fig. 1  Polymers chemical structures: a copolymer and b Thermoviscofying polymer

Fig. 2  The interfacial tension


for Carboxybetaine amphoteric
surfactant with concentration
showing the critical micelle
concentration (CMC)

Table 4  Chemical formulations for the selected candidates and their Riyadh and Indiana limestone core. The discs were cut to
concentration used in the study a diameter of 1 inch and a thickness of 3 mm, then care-
SP combination Surfactant Polymer fully smoothed using sandpapers with increasing grades
of fineness. We used silicon carbide abrasive paper elec-
S1P1 SS-880 (0.05 wt%) ATBS/AM. (0.25 wt%) tro-coated that was made in Korea for sandpapering. The
S1P2 SS-880 (0.05 wt%) TVP. (0.25 wt%) discs were then dried in an oven at 100 °C for 24 h to
S2P1 SS-885 (0.05 wt%) ATBS/AM. (0.25 wt%)
remove all water. To restore the discs to their native state
S2P2 SS-885 (0.05 wt%) TVP. (0.25 wt%) reservoir wettability, we immersed them in crude oil and
heated them in an oven at 90 °C for one week for outcrop
discs and two weeks for Indiana discs. The drop fluid
Contact angle used was the same crude oil used for saturation, which
was filtered with a 0.7-micron filter and filled into the
The process of measuring contact angle is a way to deter- drop tank in the machine using a manual screw pump.
mine the wettability of a rock by observing the angle of The bulk fluid was a combination of surfactant and poly-
a drop on its surface when surrounded by another fluid. mer in four different ratios, prepared using the seawater
Typically, the drop fluid used is oil, and the surround- described earlier with the polymer powder or surfactant
ing fluid is either air or reservoir fluid, such as seawater solution at the required concentration. The IFT 700 pro-
or EOR chemicals. For this study, we used surfactant or vided by VINCI Technologies was used for contact angle
polymer as the surrounding fluid. To prepare the discs, measurement at a temperature of 90 °C and pressure of
we used two types of rocks: carbonate outcrop rock from 1500 psi. VINCI Technologies also provided computer

13
2114 Journal of Petroleum Exploration and Production Technology (2023) 13:2109–2123

software for image analysis and angle calculation. Before displacement pumps for injection fluids, and a fluid separa-
attaching the disc holder, air bubbles were released from tor or fraction collector (Fig. 4). Computer software is used
the bulk line to prevent them from affecting vision. The to monitor the pressure transducers, pumps, and separator
cell was filled first with the bulk fluid and connected to camera in real time. The core is loaded inside the sleeve and
the bulk and drop fluid lines. After reaching the required pushed to ensure no spaces between the plugs, and the sleeve
temperature, the disc was placed on the disc holder and is placed inside the core holder and tightened carefully to
attached to the cell to start building pressure. The drop avoid any leaks. The oven is turned on to 90 °C, and the
was released carefully, and the drop angle was measured reservoir pressure is gradually built up to 1500 psi overnight
for at least half an hour of stabilization time. Contact to guarantee a stabilized temperature inside the core. The
angle is measured through the oil phase, with water-wet injection rate is 1 cc/min for all fluids, starting with water-
conventionally being defined as greater than 90° and oil- flooding, followed by chemical flooding, and finishing with
wet as less than 90°. Figure 3 shows this convention for water-flooding again.
contact angle.

Core‑flooding Results and discussion

Core-flooding is a laboratory method used to imitate reser- Rheology


voir recovery. To do this, a cylindrical rock sample is pre-
pared to match the conditions of the downhole formation in Studying the rheological properties of chemicals intended
terms of temperature, pore pressure, and overburden pres- for injection into a reservoir is crucial, especially when deal-
sure. The core is cut into small 2-inch plugs with a 1.5-inch ing with heterogeneous formations. It helps to comprehend
diameter and polished using a grinding machine to create the divergence and impact of shear on viscosity, whether it
a composite core. The plugs are then cleaned with toluene exhibits shear-thinning or shear-thickening behavior. Rheo-
solvent and dried overnight in a 90 °C oven. Air perme- logical properties are also required for surface calculations
ability is measured and corrected to Klinkenberg liquid per- related to injection pressure and flow rate. In this section,
meability. The core is then saturated with formation water we will delve into the rheological properties of chemical
brine, vacuumed for three hours to remove any air from the combinations where polymers are mixed with surfactants, as
pores, and subjected to 3000 psi pressure overnight using a detailed in Table 4. Figure 5 illustrates the impact of chang-
positive displacement pump. Porosity and pore volume are ing chemicals on shear scan behavior for all SP combina-
calculated from the weight difference between the saturated tions listed in Table 4. For S1P1, the shear scan displays
and dry cores. Oil saturation is achieved by displacing the shear-thinning behavior in the (0.001–5) 1/s shear rate range,
formation water with crude oil using a centrifuge machine. with no apparent effect of shear above 5 1/s. The data was fit
The cores are wrapped carefully with nylon and aluminum to the Carreau-Yasuda model, and the zero-shear viscosity
sheets and aged for two weeks in a 90 °C oven to restore the was found to be 35.6 cp at the given temperature and salinity.
native state reservoir wettability. On the other hand, the shear scan for S2P1 exhibits the same
The core-flooding experiment is conducted using a shear-thinning behavior in the initial part and then reaches a
VINCI core-flooding machine, which consists of a core plateau. The data was fit to the Carreau-Yasuda model, and
holder with a rubber sleeve, accumulators linked with the zero-shear viscosity estimation was 52.8 cp. The same

Fig. 3  Contact angle convention


used throughout this experimen-
tal set

13
Journal of Petroleum Exploration and Production Technology (2023) 13:2109–2123 2115

Fig. 4  Schematic diagram of the core-flooding system

Fig. 5  The effect of chemicals


on rheology at 85 °C with
seawater

trend was observed for S1P2 and S2P2, except the zero-shear that the presence of surfactant does not affect rheology in the
viscosity extracted by fitting the data to the Carreau-Yasuda (3–100) 1/s shear rate range because, for the same polymer,
model gives 124 cp and 23.5 cp, respectively. It is evident the two curves with different surfactants overlapped, but

13
2116 Journal of Petroleum Exploration and Production Technology (2023) 13:2109–2123

there is a difference in the first range, which is either minor, of 38° as plotted in Fig. 6c. But the latter one containing
as in P1, or significant, as in P2. The disparity in P2 suggests S2 exhibited a contact angle of 70° indicating a water-wet
that P2 performs better with S1 than with S2. rock Fig. 6d. On the other hand, the two weeks-aged Indi-
ana limestone was used in this set of experiments to study
Contact angle the effect of the four different chemicals on the wettability
taking the seawater contact angle as a benchmark. The sea-
Contact angle measurements were conducted on two distinct water contact angle stabilized at 4° indicating the original
types of rocks with varying aging times. The first rock sam- oil-wet rock as shown in Fig. 7e. Comparing this result to
ple was an outcrop rock that consisted of pure calcite and the previous rock with seawater we can say that 1 week is
had been aged for a week. The second sample was an Indi- not enough to restore the native state wettability on the disc
ana limestone rock that had been aged for two weeks. Both samples. The contact angle for S1P1 spread to the rock sur-
rocks were anticipated to have oil-wet properties, making face after 50 s showing strongly oil-wet rock as shown in
them ideal for exploring the impact of aging on wettability. Fig. 7a. While the contact angle for S2P1 stabilized at 12°
The one-week-aged outcrop rock was utilized to examine the for more than 500 s measured through the oil phase indicat-
impact of four distinct chemical formulations on wettability ing a strongly oil-wet rock as in Fig. 7b. Investigating the
in comparison to seawater wettability without any chemical contact angle for P2 solutions with S1 and S2 Fig. 7c and d
treatments.The contact angle for the seawater stabilized at respectively resulted in extremely oil-wetness for S1P2. The
129° which means the rock is water wet as in Fig. 6e. The S2P2 solution maintained a contact angle of 22° for over
contact angle for S1P1 stabilized at 52° as shown in Fig. 6e 1000 s during oil wetting, indicating strong oil-wet proper-
showing weakly oil-wet rock. While S2P1 is 34° indicating ties. Seawater has a benchmark contact angle of 4°. There-
an oil-wet rock more than S1P1 as shown in Fig. 6b. On fore, mixtures containing SS-885 can shift the wettability of
the other hand, both of the forthcoming solutions consist the rock from oil-wetness to water-wetness, ranging from 4
of P2, the former one containing the S1 has a contact angle degrees to 12° when combined with ATBS/AM and to 22°

Fig. 6  Contact angles for crude oil droplet in contact with calcite outcrop discs with the bulk fluid of a S1P1, b S2P1, c S1P2, d S2P2, e seawa-
ter (SW)

13
Journal of Petroleum Exploration and Production Technology (2023) 13:2109–2123 2117

Fig. 7  Contact angles for crude oil droplet in contact with Indiana limestone disc with the bulk fluid of a S1P1, b S2P1, c S1P2, d S2P2, e sea-
water (SW)

when combined with TVP. In contrast, other surfactant mix- a remarkably high recovery for water flooding. However,
tures increased oil-wetness to the maximum limit, causing Buckley et al. (1996). reported a similarly high water-flood-
the oil droplet to spread across the rock surface. ing recovery rate of 80% after two weeks of aging. Later, a
2.7 PV slug of S1P2 was injected, which extracted 33.41%
Core‑flooding results of the ROS, corresponding to 6.574% of the OIIP. In the
S2P2 experiment, 7 PVs of seawater injection resulted in an
The effectiveness of the selected chemicals in releasing oil recovery of 58.17% of the OIIP. A 2.7 PV slug of S2P2
residual oil trapped by capillary forces or bypassed by water- was then injected and successfully extracted 9.07% of the
flooding through fingering was assessed through core-flood- ROS, representing 3.794% of the OIIP. The pressure drop
ing experiments. Additionally, the optimal combination of stabilized at this stage, indicating negligible oil production.
the four chemical formulations, along with the appropriate With SP injection, the pressure drop increased due to the
slug size and sequence, was determined. high polymer viscosity, but it declined again after polymer
breakthrough.
The effect of chemicals The combination S1P1 showed the most promising incre-
mental recovery out of OIIP as can be seen in Fig. 8 and can
In the S1P1 core-flooding experiment, water-flooding be considered as the optimum combination above the other
resulted in the recovery of 62.8% of the oil initially in place three. Figure 9 displays the resistance factor changes during
(OIIP). Chemical flooding was able to recover 31.29% of the the flooding process. Following the water flooding period,
residual oil saturation (ROS), which corresponds to 11.63% the resistance factor initially rises at the start of the chemi-
of the OIIP. Once 5 pore volumes (PV) of seawater had been cal flooding process due to the formation of an oil bank
injected, the pressure drop stabilized as no more oil was pro- ahead of the water. However, it subsequently declines once
duced. However, when SP injection began, the pressure drop the water breaks through. As the process shifts to post-water
increased dramatically due to the high polymer viscosity. For injection, the resistance factor declines once more due to the
the S2P1 experiment, 7 PVs of seawater injection resulted reduced viscosity. Notably, S1P1 exhibits higher resistance
in an oil recovery of 57.7% of the OIIP. Subsequently, a 2.7 compared to the other combinations on average, which may
PV slug of S2P1 was injected, which extracted 14.87% of be attributed to the formation of micro gel at the inlet face.
the ROS, representing 6.285% of the OIIP. After 4 PVs of This micro gel can partially obstruct the pores, causing a
seawater injection, no more oil was produced, and water reduction in permeability.
injection continued up to 7 PV for consistency with other Table 5 contains a detailed summary of this set of experi-
core floods. ments. Figure 9 displays the resistance factor changes during
In the S1P2 experiment, 7 PVs of seawater injection the flooding process. Following the water flooding period,
resulted in an oil recovery of 80.3% of the OIIP, which is the resistance factor initially rises at the start of the chemical

13
2118 Journal of Petroleum Exploration and Production Technology (2023) 13:2109–2123

Fig. 8  The effect of chemicals


incremental recovery out of
original oil in place (OIIP)

Fig. 9  The effect of the chemi-


cal resistance factor (The ratio
of the chemical pressure drop
to the last pre-water flooding
pressure drop)

Table 5  Summary of the effect Core-flood S1P1 S2P1 S1P2 S2P2


of chemicals core-flooding
experiments Polymer concentration (wt%) 0.25 0.25 0.25 0.25
Surfactant concentration (wt%) 0.05 0.05 0.05 0.05
Length (inches) 11.46 11.49 11.42 11.5
Porosity (%) 16.7 18 17 15.8
Oil permeability (mD) 98.14 98 179 39
Pore volume (cc) 55.018 64.4 56.19 52.29
Swi (fraction) 0.31 0.3 0.296 0.294
Initial oil volume (cc) 37.83 45.03 39.55 36.9
Remaining oil saturation 0.372 0.45 0.197 0.418
Injected PV of chemicals 2.7 2.7 2.7 2.7
Recovery % of OIIP after water-flooding 62.80% 55.00% 80.30% 58.17%
Recovery % of OIIP by chemical flooding 11.63% 6.29% 6.57% 3.79%
Recovery % of ROS by chemical flooding 31.29% 13.96% 33.41% 9.07%

13
Journal of Petroleum Exploration and Production Technology (2023) 13:2109–2123 2119

flooding process due to the formation of an oil bank ahead stabilized at 3 PV. Thus, the optimum slug size is 3 PV
of the water. However, it subsequently declines once the of S1P1, as explained in Fig. 10. Figure 11 illustrates the
water breaks through. As the process shifts to post-water resistance factor, which indicates permeability reduction due
injection, the resistance factor declines once more due to the to polymer retention at the end of chemical flooding. The
reduced viscosity. Notably, S1P1 exhibits higher resistance resistance factor for the 2.7 PV slug size is higher post-flood-
compared to the other combinations on average, which may ing than pre-water flooding, indicating retention. Conversely,
be attributed to the formation of micro gel at the inlet face. the 1.7 PV and 3.5 PV slug sizes have lower resistance after
This micro gel can partially obstruct the pores, causing a post-water flooding than pre-water flooding, indicating per-
reduction in permeability. meability enhancement.

The effect of slug size The effect of the injection sequence

To determine the most cost-effective size, three different Four different sequences were tested to determine which one
sizes were tested: 1.7 PV, 2.7 PV, and 3.5 PV. The SP slug is the most effective in terms of oil recovery. The recovery
used for injection was selected based on the chemical exper- percent of the residual oil was used as the deciding factor.
iments, with S1P1 being identified as the optimal choice The sequences tested were SW-SP-SW, SW-P-S-SW, SW-
due to the incremental recovery achieved, as presented in S-SW-P-SW, and SW-P-SW-S-SW. The SP slug used in the
Fig. 8. The full summary of this set is elaborated in Table 6. experiments was S1P1. The graph depicted in Fig. 13 dis-
Starting with a 1.7 PV slug size core-flooding experiment, plays the Resistance Factor profile for all tested sequences.
the SP slug was injected after pre-water-flooding and post- It is important to note that the fluids containing polymer,
water-flooding. For S1P1 flooding, no oil was produced after whether combined with surfactant or used alone, exhibit
injecting 5 PV of seawater, resulting in an oil recovery of higher resistance. Conversely, when surfactant was intro-
56.76% of the OIIP. A 1.7 PV injection of the S1P1 slug duced after water flooding, the resistance factor was less
extracted 26.24% of the ROS, which is equivalent to 11.34% than 1, indicating a lower pressure drop than water flood-
of the OIIP. The 2.7 PV experiment involved injecting the ing. This can be attributed to the reduction in oil saturation,
SP slug between two water-flooding sessions. S1P1 flood- which subsequently increases the relative permeability of the
ing resulted in an oil recovery of 62.8% of the OIIP by pre- water phase, leading to a decrease in pressure drop.
water flooding. The chemical flooding extracted 31.29% of Table 7 summarizes the results of all the tests. In the
the ROS, which represents 11.63% of the OIIP. For the final SW-SP-SW sequence, injecting one pore volume (PV) of
experiment of the set, a 3.5 PV slug size was used, with 5 seawater resulted in oil recovery of 63.28% of the original
PV of seawater injection recovering 50.16% of the OIIP. oil in place (OIIP). Then, injecting 2.7 PV of the S1P1 slug
The chemical slug extracted 34.21% of the ROS, which rep- successfully extracted 31.29% of the residual oil saturation
resents 17.05% of the OIIP. It's worth noting that different (ROS), which represents 11.63% of the OIIP. The SW-P-S-
plugs produced different recoveries, even when using the SW sequence resulted in oil recovery of 55.7% of the OIIP
same slug size. The optimal slug size was determined by by injecting 5 PVs of seawater. After that, 1.5 PV of the
observing the general trend, whereby incremental recovery P1 slug was injected, followed by 1.5 PV of the S1 slug,

Table 6  Summary of the effect Core-flood 1.7 PV S1P1 2.7 PV S1P1 3.5 PV S1P1
of slug size core-flooding
experiments Polymer concentration (wt%) 0.25 0.25 0.25
Surfactant concentration (wt%) 0.05 0.05 0.05
Length 11.46 11.46 11.46
Porosity (%) 13 16.7 13.7
Oil permeability (md) 49.5 98.14 0.7
Pore volume (cc) 43.51 55.018 44.72
Swi 0.33 0.31 0.32
Initial oil volume (cc) 29 37.83 30.5
Remaining oil saturation 0.432 0.372 0.498
Injected PV of chemicals 1.7 2.7 3.5
Recovery % of OIIP after water-flooding 56.76% 62.80% 50.16%
Recovery % of OIIP by chemical flooding 11.34% 11.63% 17.05%
Recovery % of ROS by chemical flooding 26.24% 31.29% 34.21%

13
2120 Journal of Petroleum Exploration and Production Technology (2023) 13:2109–2123

Fig. 10  Optimizing the slug


size by injecting different slug
sizes ranges from 1.7 to 3.5 PV
presented in recovery out of
residual oil saturation (ROS)

Fig. 11  The effect of slug size


on the resistance factor

then SW. This extraction process extracted 17.26% of the et al. 2015). Based on the results depicted in Fig. 12, the
ROS, which represents 7.63% of the OIIP. The SW-S-SW- sequence SW-S1P1-SW showed the highest ROS recovery.
P-SW sequence showed water-flooding recovery of 52.78% This is because it shows a synergistic effect that cannot be
of the OIIP by injecting 5 PVs of seawater. Injecting 1.5 achieved by either surfactants or polymers alone. This result
PV of the S1 slug and then 1.5 PV of the P1 slug, followed is similar to the findings of Felix et al. (2015), who showed
by SW, resulted in an additional 12.08% of the ROS being that the SW-SP sequence is more effective than the SW-
extracted, which represents 5.702% of the OIIP. Lastly, the S-P sequence and polymer-augmented surfactant soaking.
SW-P-SW-S-SW sequence showed oil recovery of 52.11% of It is also worth noting that the recovery achieved by the
the OIIP by injecting 5 PVs of seawater, followed by 1.5 PVs polymer is much higher than that achieved by the surfactant,
of the P1 slug, then SW, and 1.5 PVs of the S1 slug. In total, which supports the findings of Bataweel et al. (2012) that
25.52% of the ROS was extracted, which represents 12.24% surfactants alone do not increase recovery unless there
of the OIIP. It is important to note that polymer flooding is a controlling agent for mobility. The graph depicted in
resulted in high recovery, with around 9% of the OIIP. Previ- Fig. 13 displays the Resistance Factor profile for all tested
ous research has shown that polymer flooding can even reach sequences. It is important to note that the fluids containing
15% of the OIIP (Kamal, Sultan, Al-Mubaiyedh, Hussein polymer, whether combined with surfactant or used alone,

13
Journal of Petroleum Exploration and Production Technology (2023) 13:2109–2123 2121

Table 7  Summary of the effect of injection sequence core-flooding experiments


Core-flood SW-P1-S1-SW SW-S1P1-SW SW-S1-SW-P1-SW SW-P1-SW-S1-SW

Polymer conc (wt%) 0.25 0.25 0.25 0.25


Surfactant conc (wt%) 0.05 0.05 0.05 0.05
Length 11.44 11.46 11.38 11.45
Porosity (%) 16.4 16.7 16.8 13
Oil permeability (md) 82.57 98.14 52 18.44
Pore volume (cc) 54.014 55.018 55.01 43.48
Swi 0.31 0.31 0.31 0.3
Initial oil volume (cc) 38.125 37.83 37.83 30.3
Remaining oil saturation 0.451 0.372 0.481 0.479
Injected PV of chemicals 1.5 P + 1.5 S 2.7 1.5 S + 1.5 P 1.5 P + 1.5 S
Recovery % of OIIP after water-flooding 54.92% 62.80% 51.92% 52.11%
Recovery % of OIIP by chemical flooding 7.63% 11.63% 5.70% 12.24%
Recovery % of ROS by chemical flooding 16.93% 31.29% 11.86% 25.57%

Fig. 12  The effect of sequence


recovery out of the residual oil
saturation (ROS)

exhibit higher resistance. Conversely, when surfactant was AM) and thermoviscofying polymer (TVP). The optimi-
introduced after water flooding, the resistance factor was less zation involved investigating the surfactant polymer (SP)
than 1, indicating a lower pressure drop than water flood- slug size, injection sequence, and their compatibility. The
ing. This can be attributed to the reduction in oil saturation, following observations were made:
which subsequently increases the relative permeability of the
water phase, leading to a decrease in pressure drop. • The addition of surfactant did not affect the mixture's
rheology at shear rates between 3–100 1/s.
• TVP exhibited better compatibility with SS-880 than
Conclusions SS-885.
• Thermal aging of calcite rock discs for two weeks was
A comprehensive optimization study was conducted sufficient to restore wettability, while one week was not
to identify the optimal formulation for enhanced oil adequate.
recovery under harsh conditions that yields the high- • The optimal chemical combination was carboxybetaine
est recovery. The selected candidates were two ampho- (0.05 wt%) and ATBS/AM (0.25 wt%), which resulted in
teric surfactants, SS-880 and SS-885, and two polymers, 31.29% recovery of the Residual Oil Saturation (ROS),
Acrylamido tertiary butyl sulfonate/acrylamide (ATBS/ equivalent to 11.63% of the OIIP.

13
2122 Journal of Petroleum Exploration and Production Technology (2023) 13:2109–2123

Fig. 13  The effect of sequence


resistance factor

• The optimal slug size was determined to be 3.5 PV, References


resulting in a ROS recovery of 34.21% and an OIIP
recovery of 17.05%. Ahmed ME, Sultan AS, Mahmoud M, Singh K, Kamal MS, Patil S,
• The optimal injection sequence involved co-injecting Kanj M (2022) Evaluation of the dynamic interfacial tension
between viscoelastic surfactant solutions and oil using porous
surfactant and polymer SW-S1P1-SW, yielding a ROS micromodels. Langmuir 38(20):6387–6394. https://​doi.​org/​10.​
recovery of 31.29% and an OIIP recovery of 11.63%. 1021/​acs.​langm​uir.​2c004​69
Al-Hashim HS, Obiora V, Al-Yousef HY, Fernandez F, Nofal W (2005)
Alkaline surfactant polymer formulation for carbonate reservoirs.
Acknowledgements The authors would like to acknowledge Saudi Pet Sci Technol 23(5–6):723–746. https://​doi.​org/​10.​1081/​LFT-​
ARAMCO for funding this Project (CPM2297) through the research 20003​3098
institute, Center for Integrated Petroleum Research (CIPR), and the Arihara N, Yoneyama T, Akita Y, XiangGuo L (1999) Oil recovery
technical support by EXPEC ARC. Special thanks to the College of mechanisms of alkali-surfactant-polymer flooding. In: Proceed-
Petroleum & Geosciences and the department of Petroleum Engineer- ings of SPE Asia Pacific oil and gas conference and exhibition.
ing for the permission to use the equipment and resources in the labs. Doi: https://​doi.​org/​10.​2118/​54330-​MS
Azad MS, Sultan AS, Nuaim SA, Mahmoud MA, Hussein IA (2014)
Author contributions The manuscript was written through the contri- Could VES be a part of a hybrid option to recover heavy oil in
butions of all authors. All authors have approved the final version of complex heavy oil reservoirs. Soc Pet Eng SPE Heavy Oil Conf
the manuscript. Canada 2014:3. https://​doi.​org/​10.​2118/​170191-​ms
Azad MS, Sultan AS (2014) Extending the applicability of chemical
Funding This is study was funded by the Center for Integrated Petro- EOR in high salinity, high temperature & fractured carbonate res-
leum Research (CIPR) for project (CPM2297). ervoir through viscoelastic surfactants. In: SPE annual technical
symposium and exhibition held in Al-Khobar, Saudi Arabia. Doi:
Declarations https://​doi.​org/​10.​2118/​172188-​ms
Bataweel, M. A., Shivaprasad, A. K. Y., & Nasr-El-din, H. A. T. A.
Conflict of interest There are no conflicts of interest to declare. U. (2012). Low-tension polymer flooding using amphoteric sur-
factant in high salinity/high hardness and high temperature condi-
Open Access This article is licensed under a Creative Commons Attri- tions in sandstone cores. In: Society of petroleum engineers - SPE
bution 4.0 International License, which permits use, sharing, adapta- EOR conference at oil and gas West Asia, OGWA, pp 1–23. Doi:
tion, distribution and reproduction in any medium or format, as long https://​doi.​org/​10.​2118/​155676-​ms
as you give appropriate credit to the original author(s) and the source, Buckley JS, Bousseau C, Liu Y (1996) Wetting alteration by brine and
provide a link to the Creative Commons licence, and indicate if changes crude oil: from contact angles to cores. SPE J 1(03):341–350.
were made. The images or other third party material in this article are https://​doi.​org/​10.​2118/​30765-​PA
included in the article's Creative Commons licence, unless indicated Chiou CSSPEM, Kellerhals GESPEM (1981) Polymer/surfactant trans-
otherwise in a credit line to the material. If material is not included in port in micellar flooding. Soc Petroleum Eng. https://​doi.​org/​10.​
the article's Creative Commons licence and your intended use is not 2118/​9354-​PA
permitted by statutory regulation or exceeds the permitted use, you will Dang CTQ, Chen ZJ, Nguyen NTB, Bae W, Phung TH (2013) Devel-
need to obtain permission directly from the copyright holder. To view a opment of isotherm polymer/surfactant adsorption models in
copy of this licence, visit http://​creat​iveco​mmons.​org/​licen​ses/​by/4.​0/. chemical flooding. In: SPE Asia Pacific oil and gas conference
and exhibition. Doi: https://​doi.​org/​10.​2118/​147872-​MS
Ding MC, Wang Y, Yuan F, Zhao H, Li Z (2020) A comparative
study of the mechanism and performance of surfactant- and

13
Journal of Petroleum Exploration and Production Technology (2023) 13:2109–2123 2123

alkali-polymer flooding in heavy-oil recovery. Chem Eng Sci high-temperature, high salinity carbonates. In: SPE Middle East
219:115603. https://​doi.​org/​10.​1016/j.​ces.​2020.​115603 oil and gas show and conference, pp 1–15. Doi: https://​doi.​org/​
Elmuzafar Ahmed M, Sultan AS, Saikia T, Mahmoud M, Patil S, Kanj 10.​2118/​164241-​MS
M (2023) Investigating effects of chelating agents on viscoelastic Lu J, Pope GA (2017) Optimization of gravity-stable surfactant flood-
surfactant flooding at the pore scale using micromodels. Energy ing. SPE J. https://​doi.​org/​10.​2118/​174033-​PA
Fuels 37(2):1070–1080. https://​doi.​org/​10.​1021/​acs.​energ​yfuels.​ Luo P, Wu Y, Huang S (2013) Optimized surfactant–polymer flooding
2c037​77 for western canadian heavy oils. In: SPE heavy oil conference
Fang Y, Yang E, Guo S, Cui C, Zhou C (2022) Study on micro remain- Canada, 2013. Doi: https://​doi.​org/​10.​2118/​165396-​ms
ing oil distribution of polymer flooding in Class-II B oil layer of Mohammadi H, Delshad M, Pope GA (2008) Mechanistic modeling
Daqing Oilfield. Energy. https://​doi.​org/​10.​1016/j.​energy.​2022.​ of alkaline / surfactant / polymer floods. Spe 110212(2006):1–13.
124479 https://​doi.​org/​10.​2118/​110212-​PA
Felix U, Ayodele TO, Olalekan O (2015) Surfactant-polymer flooding Omran M, Akarri S, Torsaeter O (2020) The effect of wettability and
schemes (a comparative analysis). In: Society of petroleum engi- flow rate on oil displacement using polymer-coated silica nano-
neers - SPE Nigeria annual international conference and exhibi- particles: a microfluidic study. Processes. https://​doi.​org/​10.​3390/​
tion, NAICE. Doi: https://​doi.​org/​10.​2118/​178367-​ms PR808​0991
Glover C, Puerto M, Maerker J, Sandvik E (1979) Surfactant phase Osterloh WT, Jante Jr MJ (1992) Surfactant-polymer flooding with ani-
behavior and retention in porous media. SPE J 19(3):183–193. onic PO/EO surfactant microemulsions containing polyethylene
https://​doi.​org/​10.​2118/​7053-​PA glycol additives. In: Eighth symposium on enhanced oil recovery.
Han M, Alsofi A, Fuseni A, Zhou X, Hassan S, Aramco S (2013) IPTC Doi: https://​doi.​org/​10.​2118/​24151-​ms
17084 Development of chemical EOR formulations for a high Somasundaran P, Hanna HS (1979) Adsorption of sulfonates on reser-
temperature and high salinity carbonate reservoir. Iptc. https://​ voir rocks. Soc Petroleum Eng J. https://d​ oi.o​ rg/1​ 0.2​ 118/7​ 059-P ​ A
doi.​org/​10.​2523/​17084-​MS Sui X, Chen Z, Kurnia I, Han X, Yu J, Zhang G (2019) Alkaline-sur-
Journal SPE, Miller CA, Hirasaki GJ, Miller CA, Puerto M (2008) factant-polymer flooding of active oil under reservoir conditions.
Recent advances in surfactant EOR recent advances in sur- Fuel 262:116647. https://​doi.​org/​10.​1016/j.​fuel.​2019.​116647
factant EOR. In: International petroleum technology conference, Tabary R, Bazin B, Douarche F, Moreau P (2013) Surfactant flood-
Kuala Lumpur Malaysia, vol 3(5). Doi: https://​doi.​org/​10.​2118/​ ing in challenging conditions: towards hard brines and high
115386-​MS temperatures. In: SPE Middle East oil and gas show and confer-
Kamal MS, Sultan AS, Al-Mubaiyedh UA, Hussien IA, Pabon M ence, MEOS, proceedings, pp 1–16. Doi: https://​doi.​org/1​ 0.​2118/​
(2014) Evaluation of rheological and thermal properties of a 164359-​ms
new fluorocarbon surfactant-polymer system for EOR appli- Tabrizy VA (2014) Investigation of C ­ O2 enhanced oil recovery using
cations in high-temperature and high-salinity oil reservoirs. dimensionless groups in wettability modified chalk and sandstone
J Surfactants Deterg 17(5):985–993. https://​doi.​org/​10.​1007/​ rocks. J Petroleum Eng 2014:1–16. https://​doi.​org/​10.​1155/​2014/​
s11743-​014-​1600-7 430309
Kamal MS, Sultan AS, Al-Mubaiyedh UA, Hussein IA (2015a) Review Wang J, Han M, Fuseni AB, Cao D (2015) Surfactant adsorption in
on polymer flooding: rheology, adsorption, stability, and field surfactant-polymer flooding for carbonate reservoirs (2).pdf. In:
applications of various polymer systems. Polym Rev 55:491–530. SPE Middle East oil and gas show and conference. Doi: https://​
https://​doi.​org/​10.​1080/​15583​724.​2014.​982821 doi.​org/​10.​2118/​172700-​ms
Kamal MS, Sultan AS, Al-Mubaiyedh UA, Hussein IA, Feng Y Zhenquan L, Zhang A, Cui X, Zhang L, Guo L, Shan L (2013) A
(2015b) Rheological properties of thermoviscosifying polymers successful pilot of dilute surfactant-polymer flooding in shengli
in high-temperature and high-salinity environments. Can J Chem oilfield. In: SPE improved oil recovery symposium, pp 1–6. Doi:
Eng 93(7):1194–1200. https://​doi.​org/​10.​1002/​cjce.​22204 https://​doi.​org/​10.​2118/​154034-​MS
Kamal MS, Sultan AS, Hussein IA (2015c) Screening of amphoteric Zhu Y, Zhang Y, Hou Q, Yuan H, Jian G (2013) IPTC 16433 effect of
and anionic surfactants for cEOR applications using a novel main factors on oil recovery of surfactant-polymer flooding. In:
approach. Colloids Surf A 476:17–23. https://​doi.​org/​10.​1016/j.​ Society of petroleum engineers - International petroleum technol-
colsu​rfa.​2015.​03.​023 ogy conference, IPTC. Doi: https://d​ oi.o​ rg/1​ 0.2​ 523/i​ ptc-1​ 6433-m​ s
Kesarwani H, Sharma S, Mandal A (2021) Application of novel col-
loidal silica nanoparticles in the reduction of adsorption of sur- Publisher’s Note Springer Nature remains neutral with regard to
factant and improvement of oil recovery using surfactant polymer jurisdictional claims in published maps and institutional affiliations.
flooding. ACS Omega. https://d​ oi.o​ rg/1​ 0.1​ 021/a​ csome​ ga.1​ c0029​ 6
Levitt D, Klimenko A, Jouenne S, Chamerois M, Bourrel M
(2013) Overcoming design challenges of chemical EOR in

13

You might also like