You are on page 1of 22

Journal of Petroleum Science and Engineering 208 (2022) 109475

Contents lists available at ScienceDirect

Journal of Petroleum Science and Engineering


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

A review on parameters affecting nanoparticles stabilized foam


performance based on recent analyses
Shahrul Aida Ab Rasid a, b, *, Syed M. Mahmood b, Nor Idah Kechut a, Saeed Akbari b
a
PETRONAS Research Sdn. Bhd., 43000, Kajang, Selangor, Malaysia
b
Department of Petroleum Engineering, Universiti Teknologi PETRONAS, 32610, Bandar Seri Iskandar, Perak, Malaysia

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

Keywords: In recent years, much attention has been paid to the addition of nanoparticles (NPs) into the surfactant foam
nanoparticles Stabilized foam structure to form durable foams. Although surfactant foam as an enhanced oil recovery (EOR) technique
Foam stability significantly increases the viscosity of the injected fluid compared to gas injection, it has been observed that the
Retention
foam structure does not have the required stability in cases such as high temperatures and high salinity. Although
Concentration
Salinity
NP addition increases stability, the extent of this increase depends on several parameters, including NP prop­
Flow velocity erties (i.e., size, type, surface wettability) and reservoir properties (i.e., salinity of formation water, presence of
Size oil, reservoir temperature), process parameters (i.e., NP concentration and flow rate), synergistic effects between
Type the surface charge of NPs and the net charge of surfactant, and NP loss in the porous medium. This study aims to
Wettability review and summarize the findings of previous studies to conclude the effects of each of the parameters above on
foam stability, identify differences, and determine gaps for future studies. Throughout this report, a study
background is provided, followed by the concept of stability and how to determine it in different tests for NP-
surfactant foam. Next, the mechanisms of increasing stability by NP addition are briefly presented. Lastly,
previous findings related to the effect of each parameter on NP foam stability are presented and discussed.
Although several detailed reviews of NP-stabilized foam have been published previously, the present study
differs from them in that the effects of additional parameters were investigated by reviewing new findings.
Moreover, an attempt has been made to discuss the effects of parameters from different angles and according to
their role in the mechanisms of NP-stabilized foam, such as particle detachment energy and maximum capillary
pressure. The findings of this review show that i) oil presence lowers foam stability, ii) increasing NP concen­
tration as well as decreasing NP size, temperature, NP retention, and salinity lead to increased stability, and iii)
for parameters including NP surface wettability, NP types, and shear rate/flow velocity there are optimum points
that result in the best foam performance. NP loss in a porous medium is an economically damaging process that
occurs through mechanisms including adsorption, mechanical entrapment and log-jamming, and particles
settling due to gravity. Depending on the type of mechanism, the strategy for minimizing or removing the in­
fluence of these mechanisms differs. However, there is currently no way to determine and measure the contri­
bution of each of these mechanisms to the total retention in dynamic tests, which could be a specific topic for
future research.

1. Introduction originate from various causes, including rock and fluid expansion, dis­
solved gas release, water influx from neighboring aquifers, and gravity.
A combination of three different methods can achieve maximum oil Secondary oil production involves the injection of fluids into a reservoir,
recovery from petroleum reservoirs. The conventional methods of pro­ such as water or gas (into the gas cap), to maintain reservoir pressure
ducing oil from a reservoir involve primary and secondary methods and preserve the pressure difference required to displace the oil toward
(Satter and Iqbal, 2015). The primary production of an oil or gas the production well. A succession of both production stages can extract
reservoir at the expense of the reservoir’s inherent energy might roughly a third of the original oil in place (Gbadamosi et al., 2019).

* Corresponding author. PETRONAS Research Sdn. Bhd., 43000, Kajang, Selangor, Malaysia & Department of Petroleum Engineering, Universiti Teknologi
PETRONAS, 32610, Bandar Seri Iskandar, Perak, Malaysia
E-mail addresses: shahrul.aida.abrasid@gmail.com, aida.rasid@petronas.com (S.A. Ab Rasid).

https://doi.org/10.1016/j.petrol.2021.109475
Received 2 February 2021; Received in revised form 16 August 2021; Accepted 4 September 2021
Available online 8 September 2021
0920-4105/© 2021 Elsevier B.V. All rights reserved.
S.A. Ab Rasid et al. Journal of Petroleum Science and Engineering 208 (2022) 109475

Tertiary recovery is a stage in which Enhanced Oil Recovery (EOR) compared to the viscosity of the single gas phase, causes the displacing
techniques are usually applied to recover the remaining trapped or fluid to be diverted from the high-permeability layer to the low-
by-passed oil by altering the reservoir properties (Akbari et al., 2019a). permeability layers, resulting in more oil being produced from the
Among the various techniques available for EOR, such as chemical bypass layers, leading to a significant increase in oil recovery (Zhu et al.,
injection, gas injection, and thermal oil recovery, the gas injection 2004).
technique is particularly important, with an application rate of about The above advantages come into play if the surfactant foam has the
39% in global EOR projects (AlYousef et al., 2017). Gas injection in­ required stability under reservoir conditions, which is not always the
volves injecting hydrocarbon and non-hydrocarbon gases such as case. Foam stability is the foam’s ability to remain in bubble structure
methane, air, carbon dioxide, natural gas, and nitrogen into reservoirs, against the process of film thinning and coalescence at which the thin
improving oil displacement and volumetric sweep efficiency by film ruptures (Schramm and Wassmuth, 1994). In the reservoir, the
reducing oil viscosity, capillary forces, and interfacial tension. However, injected surfactant foam is in a porous medium in contact with oil, brine,
the high mobility and low density of gas limit the oil recovery from this and possibly high temperature, which may adversely affect the stability
technique. The tendency of a gas to channel through the high perme­ of the surfactant foam. The chemical structure of many surfactants is not
ability zone and separation of gas and liquid due to density difference designed for high-temperature applications and does not have the
and gravity segregation results in low sweep efficiency in the oil reser­ required stability. By a process such as hydrolysis, which is enhanced at
voir (Lake, 1989), see Fig. 1. high temperatures and in the presence of ions in formation water, the
For reservoirs with poor macroscopic sweep efficiency in the Gas surfactants readily precipitate and lose their function in the foam
EOR projects, surfactant foam injection was proposed as an alternative structure, accelerating the destruction of the lamella (Kalam et al.,
method in the 1950s (Boud and Holbrook, 1958). Surfactant foam is 2019). Previous research has demonstrated that oil in a phenomenon
produced by injecting gas into a liquid phase containing surfactant, known as hindered generation inhibits the effective generation of
forming a dispersion of gas bubbles in the liquid separated by a thin film lamella in porous media. On the other hand, depending on the surfactant
called lamellae, see Fig. 2. There are three (3) foam generation mech­ type and the mineralogy of the reservoir rock, the injected surfactant
anisms: snap-off, lamellae division, and leave behind mechanism may be highly absorbed, resulting in a significant reduction of the sur­
(Alcorn et al., 2020; Yekeen et al., 2018b); see Fig. 3. The snap-off factant concentration in the lamella, which significantly reduces the
mechanism is a process where a gas bubble is generated when the gas foam stability.
moves through a pore throat (Gugl, 2020). It is highly dependent on the Various methods have been proposed to increase the stability of
local dynamic capillary pressure at the pore throat (Rossen, 2017). surfactant foam, of which the addition of nanoparticles (NPs) to sur­
Therefore, the capillary pressure at the pore throat must be greater than factant foam is one of the most successful methods for several reasons.
the capillary pressure at the front of the interface for snap-off to occur. First, because NPs are solid, their performance is not dependent on
The foam generation through lamellae division occurs when a foam conditions such as high temperatures and the presence of oil and brine.
bubble approaching a “branch point” in the pore, causing the lamellae to Second, because of their nano size, they are retained very little when
split and formed individual bubbles (Ibrahim and Nasr-El-Din, 2019). moving in a porous medium, and therefore the permeability of the rock
This mechanism is only possible when the existing foam bubbles are remains largely unaffected. Third, their adsorption on the reservoir rock
bigger than the pore’s size and a sufficient pressure gradient exists to is negligible. Fourth, NPs are inexpensive because they may be obtained
displace the lamellae in the pore throat. Lastly, the leave behind from various low-cost sources, such as coal fly-ash. Finally, since they
mechanism creates continuous gas bubbles from the gas front into a can be grafted, their wettability is easily adjustable to obtain a foam with
saturated liquid medium (Sagbana and Abushaikha, 2021). In this acceptable stability.
condition, the continuous gas bubbles accumulated and block the pas­ The extent and durability of foam stability due to the addition of NP
sage, thus resulting to a dead-end pathway. The foam generated through depends on several factors such as NP size, NP surface wettability, NP
this mechanism is reported to generate less stable foam and low flow concentration, a synergistic effect between the surface charge of NPs and
resistance as the bubble generated are parallel to the direction of flow the net charge of surfactant, salinity of formation water, presence of oil,
during the drainage process (Zeng et al., 2016; Hamza et al., 2017). reservoir temperature, flow rate, NP loss in the porous medium, and NP
For two principal reasons, foam consisting of gas trapped in a type. This paper aims to investigate how and to what extent the various
continuous liquid phase has much less mobility (M) than a displacing parameters mentioned affect the foam stability created by the addition
fluid consisting only of gas. First, the relative permeability of the gas of NP by reviewing recent studies in this field. This article goes on to
phase is considerably reduced by confinement in the foam bubbles, and define and explain how stability is measured in various experiments
second, the apparent viscosity of the displacing fluid is considerably performed on foam. In the following, mechanisms will be discussed that
increased by the drag forces exerted on the pore walls by the moving explain how foam stability increases due to NP addition. In the next
bubbles. The high mobility observed with gas injection leads to several section, the effect of various factors on foam stability formed by the
problems, including an early breakthrough. The foam resistance to easy addition of NP will be explained. NP retention studies will be reviewed.
flow, which in some cases has increased the apparent viscosity 1000-fold Although several detailed reviews of NP-stabilized foam have been
published previously (Yekeen et al., 2018a; Zhang et al., 2021), the
present study differs from them in that the effects of additional param­
eters were investigated by reviewing new findings. Moreover, an
attempt has been made to discuss the effects of parameters from
different angles and according to their role in the mechanisms of
NP-stabilized foam, such as particle detachment energy and maximum
capillary pressure. Throughout this article, parameters that require
additional investigation are highlighted, and further analysis is
suggested.

2. Experimental techniques for evaluating the performance of


nanoparticle-surfactant foam

Fig. 1. Gas channeling in high permeability reservoirs and gas segregation due The potential use of NPs as a foam stabilizer has been assessed in
to gravity. recent studies through bulk scale, bubble scale, and pore-scale stability

2
S.A. Ab Rasid et al. Journal of Petroleum Science and Engineering 208 (2022) 109475

Fig. 2. A typical foam structure.

experiments similar to the method used for conventional surfactant significantly, calculating the fluid’s viscosity in the porous medium can
foam (Eftekhari et al., 2015; Farhadi et al., 2016; Guo and Aryana, 2016; be a reliable indicator of foam production and penetration. The foam
Prigiobbe et al., 2015). stability is represented by the pressure drop across the sample or the
In bulk-scale stability experiments, a liquid containing a surfactant corresponding gas apparent viscosity measured. After reaching a
and an NP comes into contact with the gas, and there is sufficient me­ steady-state, the apparent viscosity can be calculated using Darcy’s law
chanical energy for the gas and liquid to generate foam. The container (Equation (1)).
used can be a calibrated cylinder, a high-pressure viewing cell, or a
AK Δp
foamscan instrument. Mechanical energy can be supplied from sources μ= . (1)
Q L
such as shearing stirring by laboratory blender or shaking stirring
(Razali et al., 2018; Wang et al., 2017). Bubbling by gas sparging can be where μ represents the apparent foam viscosity, K stands for perme­
employed, especially if the gas being used is not air. The most important ability, Δp is the pressure drop, L is the length, A is the cross-sectional
parameters measured in this type of test are foamability, foam half-life, area, while Q is the volumetric flow rate. In this equation, the pres­
and drainage half-life. The maximum height that occurs after foaming, sure difference between the inlet and outlet of the core plays a vital role
often when there is no separate liquid phase in the container, is foam­ in calculating viscosity, which must be carefully measured and recorded.
ability. The foam half-life is defined as the time required for the foam to In foam dynamics studies, the mobility reduction factor (MRF) is
shrink to half its original volume at a given temperature and pressure another crucial parameter obtained by dividing the apparent foam vis­
(see Fig. 4a for the determination of the foam half-life.) Drainage cosity by the apparent viscosity of the single-phase gas flow. Foam sta­
half-life is defined as the time required to drain half of the liquid used to bility measured at the core scale is assumed to be imitating the foam
form a foam (see Fig. 4b.) In bulk-scale stability experiments, the foam behavior in the reservoir due to its similar condition.
half-life and the drainage half-life are the most important indicators for
evaluating foam stability. 3. Nanoparticle-mediated foam stabilization mechanisms
Observations and results of bulk scale stability tests are often insuf­
ficient to fully understand the stabilized foam NPs’ performance. The foam structure is not stable over time and therefore changes.
Therefore, it is needed to examine foam texture, bubble morphology, These changes are happened due to coalescence, coarsening, and foam
and thin liquid to understand the mechanisms involved in foam pro­ drainage, see Fig. 6. In the bubbles coalescence, the instability of the
duction, stability, and decay. Such studies are performed in the context inter-bubbles films results in bubble breakage and merging of the two
of bubble-scale stability experiments in which foam microbubbles were smaller bubbles to form larger bubbles due to rupture of liquid films
allowed to stabilize, and the bubble suspension is placed on a micro­ between bubbles (Yekeen et al., 2018a). During foam coarsening, gas
scope slide (AlYousef et al., 2018). Fig. 5 shows an experimental setup diffuses from tiny bubbles to larger bubbles because of the greater
for the bubble scale stability test. The number of lamellae per unit vol­ pressure difference (known as Laplace pressure) between the inner and
ume is used to describe foam texture in these tests. Image analysis exterior of smaller bubbles; as a result, the smaller bubbles fade over
software (e.g., Image-J) is typically used to determine bubbles’ size time, increasing in average bubble size. Foam drainage is the process of
distribution and average diameter. liquid draining through networks of plateau borders in a fresh foam due
Although bulk and bubble-scale stability tests can be employed for to combined gravitational and capillary forces (Sun et al., 2008).
qualitative analyses of NP-stabilized foams, experiments in porous Four main reasons are suggested to explain why these phenomena (i.
media and dynamic mode are unavoidable requirements for quantitative e., coalescence, coarsening, and drainage) occurred (Binks et al., 2008;
evaluation of NP-stabilized foams’ performance. In dynamic studies in a Horozov, 2008; Kaptay, 2006). The first reason is the low energy
porous medium, either pre-generated foam is injected or the foam is required to remove the surfactant from the gas-liquid interface. If this
generated in situ by simultaneous gas and liquid injection. Another way energy is provided, the surfactant (or foaming agent) is readily removed
to create foam in situ is to inject alternating gas and liquid into the from the interface, eventually leading to foam instability. The second
porous medium in a water-alternating-gas (WAG) process. The porous factor is the pressure difference (known as capillary pressure) that exists
media in these types of tests can be any of the glass bead packs (Risal along with the gas-liquid interface, and as this pressure difference in­
et al., 2018), sand packs (Sun et al., 2015), sandstone, or carbonate creases, coalescence occurs more rapidly. The third reason is the low
reservoir cores (Rognmo et al., 2017; Eftekhari et al., 2015). Because the energy required to rupture the thin film and reduce the distance between
development of foam increases the viscosity of the displacing fluid gas bubbles in the surfactant foam structure to zero. The fourth issue is

3
S.A. Ab Rasid et al. Journal of Petroleum Science and Engineering 208 (2022) 109475

Fig. 3. Foam generation mechanisms in porous media, (a) leave behind, (b) snap-off, and (c) lamella mobilization and division. Adapted from Almajid and Kovs­
cek (2020).

the low viscosity of the liquid phase in the surfactant foam structure, required for NP with the radius of the particle, R, the interface/surface
which increases the speed of liquid drainage from the liquid film. Adding tension γ αβ and the particle contact angle at the interface, θ (Gbadamosi
NPs to the surfactant foam structure will weaken these four destructive et al., 2019).
phenomena and thus make the structure more stable. The following
paragraphs explain how the addition of NP improves the W r = πR2 γαβ (1 ± cos θ)2 (2)
surfactant-foam stability. The sign "+" denotes particle removal into the gas phase, while the
NP adsorption in the gas-liquid interface is an almost irreversible sign "-" denotes particle removal into the liquid phase. Under similar
process that results in the long-term stability of the foam structure. This conditions, it has been observed that removing NP with a size of 10 nm
is because removing NP from the gas-liquid surface requires high energy and a contact angle close to 90◦ from an NP foam structure takes 1000
called particle detachment energy (Binks and Lumsdon, 2000), which is times more force than removing surfactant from a surfactant foam
almost impossible to supply under reservoir conditions. Equation (2) structure. Hence the addition of NP to the surfactant foam structure
shows the method for calculating the particle detachment energy (Wr )

4
S.A. Ab Rasid et al. Journal of Petroleum Science and Engineering 208 (2022) 109475

Fig. 4. Foam stability determination using a) the average half-life of foam. Adapted from (Osei-Bonsu et al., 2017), b) The drainage half-life of SDS-foam in the
presence of oil. Adapted from (Yekeen et al., 2016).

where V is the drainage velocity, hf is the thin film (lamellae) thickness,


R is the foam film radius while μe is the viscosity of the aqueous phase, πd
is the disjoining pressure. As Equation (4) shows, Pc plays a decisive role
in the liquid drainage velocity equation. The direct effect of Pc reduction
on drainage velocity reduction is noticeable. Although the addition of
NP will increase the hf , it will also increase the viscosity and R at the
same time, which will result in a decrease in the drainage velocity.
The increase in stability due to the increase in the maximum capil­
lary pressure of coalescence (PMax
c ) is another significant effect of adding
NP to surfactant foam. PMax
c is the pressure that causes the gap between
the bubbles to be zero, resulting in the rupture of the liquid film. In
2006, the relation given in Equation (5) was developed for PMax c by
Kaptay (2006), which explains the relationship between this parameter
Fig. 5. Investigation of CO2 foam stability at bubble-scale. Reproduced from and the properties of NP and foam.
(Yekeen et al., 2017a).
2γaw
PMax
c =β (cos θ + z) (5)
R
results in a durable structure.
Moreover, the presence of solid NP in the film’s structure will create where γaw , θ, R, and β are air-water interfacial tension, particles contact
a barrier that prevents gas interface diffusion and will significantly angle, particle radius, and a theoretical packing parameter, respectively.
reduce the rate of coalescence, coarsening, and liquid drainage. Z is a constant and depends on the arrangement of particles in the thin
Depending on the wettability, the arrangement of the NPs in the liquid liquid film, which is zero for a single layer of particles. In this equation,
film varies (Horozov, 2008); see Fig. 7. Comparing the stability created increasing (particle parking at the air-water interface of the foam) and
for the three known arrangements shows that the network structure and decreasing R will increase PMax
c and thus increase stability.
the monolayer structure will provide the most and the least stability,
respectively. This arrangement mechanism delays the possibility of foam 4. Critical parameters affecting nanoparticles stabilized foam
film rupturing as the particles continuously re-arrange themselves until performance
the lamellae reach a thickness that will break. On the other hand, the
presence of NP in the liquid film will increase the fluid viscosity, which Several parameters have been identified to affect the stability of NP-
will reduce the speed of liquid drainage, resulting in more durable foam. stabilized foam. These parameters include NP size, NP surface wetta­
The pressure difference that forms around the gas-liquid interface, bility, NP concentration, a synergistic effect between the surface charge
known as capillary pressure (Pc ), plays a decisive role in determining of NPs and the net charge of surfactant, salinity of formation water,
foam stability. High-pressure difference results in low foam stability. presence of oil, reservoir temperature, flow rate, NP loss in the porous
Adding NP reduces Pc . Equation (3) shows how foam properties affect Pc medium, and NP type. The effect of each parameter toward NP-
(Alzobaidi et al., 2017). R is inversely related to Pc , and as R increases, Pc stabilized foam stability is discussed in the following subsections.
will decrease. Equation (4) depicts how Pc affecting the drainage
velocity. 4.1. Effect of particles size
/
Pc = γ R(1 − ∅)0.5 (3)
The stability of NP-stabilized foam is highly dependent on NP size at
the same surface characteristics (Kim et al., 2016). Summary of findings
2h3f
V= (Pc − πd ) (4) related to the NP size effect on foam stability is presented in Table 1. Hu
3μe R2 et al. (2018) generated the bulk foam using a shaker (150 rpm) for 12 h

5
S.A. Ab Rasid et al. Journal of Petroleum Science and Engineering 208 (2022) 109475

Fig. 6. Scheme of the three main foam destabilization mechanisms. (a) Drainage of the liquid due to gravity, (b) coalescence of two bubbles, and (c) coarsening of the
bubbles due to the Laplace pressure. Adapted from (Fameau and Salonen, 2014).

500, 1000, and 5000 ppm improved foam half-life and thus increased
stability. However, these results were not true for NPs calcium sulfate
and iron oxide under all the above conditions and decreased the time of
the foam half-life and hence the foam stability. As a result, depending on
the type of surfactant foam, there are NPs in micro size that can be used
to enhance stability.
In fact, small-size NPs can readily be adsorbed to the gas-liquid
interface due to greater diffusivity and a higher number of NPs avail­
able in situ (Kim et al., 2016). On the other hand, it has also been
observed that large size NPs can also exist in the lamellae and at the
plateau border depending on the lamellae thickness, thus improved the
foam stability by delaying the liquid drainage process (Fameau and
Fig. 7. Nanoparticle arrangement at the thin liquid film: (a) a monolayer of
Salonen, 2014). When Equations (2) and (5) are compared, it is obvious
bridging particles, (b) a bilayer of closed-packed particles, and (c) a network of
that reducing particle size, R, has the opposite effect on particle
particle aggregates inside the film. Adapted from (Massarweh and Abush­
aikha, 2021).
detachment energy and maximum capillary pressure as raising particle
detachment energy and decreasing maximum capillary pressure. Even if
the rest of the key parameters such as NP wettability, surface coating etc.
to ensure adsorption equilibrium and they observed that foam stability
is kept constant, it seems that an optimal NP size exists which is favor­
was inversely proportional to particle size, as presented by Fig. 8. In
able for foam stability. Therefore, further studies are recommended to
addition, Kim et al. (2016) observed that apparent foam viscosity in­
experimentally prove whether such an optimal particle size exists for a
creases with decreasing NP size from their sand pack and core flooding
given NP surfactant foam system.
experiments.
In a study by Rafati et al. (2016), foam half-life measurements were
conducted using barium sulfate NP (1.38 μm), calcium sulfate NP (7.61 4.2. Effect of nanoparticle surface wettability
μm), calcium carbonate NP (12.16 μm), strontium sulfate NP (64.14
μm), and iron oxide NP (131.04 μm). The results of this study show that The surface wettability of NPs has been described as one of the pa­
the addition of NPs including barium sulfate, calcium carbonate, and rameters involved during the NPs’ foam stabilization mechanism (Binks
strontium sulfate into the structure of surfactant foam at concentrations and Lumsdon, 2000). Table 2 provides a summary of the influence of NP
of 0.05, 0.1, 0.5, and 0.1 wt% and at surfactant concentrations of 200, surface wettability on foam stability. The NP surface wettability is
determined through the contact angle between the particles with the

6
S.A. Ab Rasid et al. Journal of Petroleum Science and Engineering 208 (2022) 109475

Table 1
Table summary of the influence of nanoparticles size on foam stability.
Authors Nanoparticle Types Nanoparticle Surfactant Main Findings Related to Particle Size
Size (nm)

Kim et al. SiO2 5 1. The smaller NPs can be more easily adsorbed at the decane/brine or CO2/
(2016) 12 brine interface due to the greater diffusivity and the higher number
25 concentration.
80 2. For NPs with the same surface wettability, the internal structure, and
consequently stability and apparent viscosity of NP-stabilized emulsions/foams
is strongly dependent on NP size.
Rafati et al. CaCO3, CaSO4, BaSO4, SrSO4, 12,000, AOS (anionic) 1. SrSO4 showed less stable foam compared to BaSO4 and CaCO3 as the particles
(2016) Fe2O3 7600, are much larger and make monolayer bridges which induce disjoining pressure
1300, in the lamellae and destabilizes foam by dewetting and film rupture.
64,000,
131,000,
Guo and Fumed SiO2 T30, 179–223, AOS (anionic), 1. Larger size NPs generated foams with slightly lower formability and stability
Aryana Nano clay 356-419, SDS (anionic), LAPB compared to SiO2 NPs, which is in large part due to its particle size
(2016) (amphoteric)
Hu et al. SiO2 NPs (SNPs) 20, 100, 500 CTAB (cationic), SDS 1. Among the three SNPs (20 nm, 100 nm, and 500 nm), 20 nm SNP exhibited
(2018) (anionic) the most effective improvement for foam stability. Smaller SNP was able to
create a more compact solid-like layer, which might effectively prevent both
coarsening and coalescence.
Razali et al. SiO2 A300 (hydrophilic), SiO2 7, 12, <100, 25 MFOMAX 1. The highest effect of catalytic behavior on the foam stability of MFOMAX can
(2018) R816 (hydrophobic), ZnO, TiO2 (zwitterionic) be seen when A300 was used as the foam stabilizer. This may be due to the large
surface area/smaller particle size of A300 as compared to other NPs.
2. Larger surface area/smaller particle size provides more spaces for the
surfactant to attach to the NP’s surface and promotes catalytic behavior.

not firmly adhere to the gas-liquid interface (Horozov, 2008; Kruglyakov


et al., 2011). Singh and Mohanty (2015) visually observed the retarding
process of liquid draining from foam lamellae consisting of hydrophilic
NPs in anionic surfactant using a confocal microscopy image. Their re­
sults show that the lamellae thickness was thicker in the presence of NPs,
indicating slow liquid drainage. The foam film became planar after the
critical film thickness is achieved, and further draining of liquid in the
lamellae caused the liquid to be drawn toward the particles. The NP
packing and rearrangement in the lamellae stabilize the film. The NPs
may transition from a bilayer to a monolayer arrangement, as shown in
Fig. 10. Most importantly, hydrophilic NPs in a similar charge surfactant
do not deteriorate foamability. Instead, it promotes the surfactant
adsorption onto the gas-liquid interface through electrostatic repulsion,
Fig. 8. Half-lives versus cationic cetyltrimethylammonium bromide (CTAB) thus effectively reducing the solution’s interfacial tension (Vatanparast
concentration for CTAB solution and CTAB/silica nanoparticles (SNPs) disper­ et al., 2018).
sions. Data has been extracted from (Hu et al., 2018). Partially hydrophobic NPs were found to enhance foam stability by
adhering to the gas-liquid interface. These NPs can achieve the
gas-liquid surface. Hydrophilic particle has a contact angle of less than maximum attachment energy with a contact angle close to 90◦ , thus
90◦ ; therefore, a significant portion of it is in the liquid phase. A particle improve foam stability (Binks and Lumsdon, 2000; Denkov et al., 1992;
with the opposite characteristics is known as a hydrophobic NP, as Horozov, 2008; Kaptay, 2006). Arriaga et al. (2012) reported that bulk
shown in Fig. 9 (Binks et al., 2008). foam stability increased as NPs become partially hydrophobic in the
Hydrophilic NPs improve foam stability, although the particles do presence of negatively charged silica NPs with a cationic surfactant. As

Table 2
Table summary of the influence of nanoparticles’ surface wettability on foam stability.
Authors Nanoparticle Types Nanoparticles Surfactant Main Findings
Wettability

Kruglyakov et al. Aerosil SiO2 Hydrophobic 1. Increasing the concentration of hexylamine led to the increasing of
(2011) hydrophobicity of the particle aggregates and the increasing of the contact
angle. The increase of the stability agrees well with the increase of the contact
angle (θ = 26–56◦ ).
Arriaga et al. SiO2 NP Hydrophilic n-amylamine 1. Without the addition of amine, the particles are too hydrophilic to stabilize
(2012) (cosurfactant) foams. With increasing ca the particles are rendered increasingly hydrophobic
cationic) and therefore act as foam stabilizers.
Singh and SiO2 NPs Nyacol DP Hydrophilic AOS (anionic) 1. The present case of hydrophilic NPs, after initial drainage, a critical film
Mohanty 9711, SiO2, PEG coating thickness is achieved, and the film becomes planar. Further drainage causes the
(2015) capillary pressure to draw liquid toward the particle, thus stabilizing the film by
a bridging mechanism
Zargartalebi et al. SiO2 AEROSIL816 Partially SDS (anionic) 1. The improvement in surfactant flooding efficiency was more considerable for
(2015) Fumed SiO2 AEROSIL300 hydrophobic, slightly hydrophobic NP-augmented surfactant solutions with respect to
Hydrophilic hydrophilic ones.
Choi et al. (2020) SiO2 Hydrophilic CTAB (cationic) 1. SiO2 NP modified with ICP (increasing contact angle from 62◦ to 82◦ )
improved foam stability compared with un-modified SiO2.

7
S.A. Ab Rasid et al. Journal of Petroleum Science and Engineering 208 (2022) 109475

Fig. 9. Hydrophilic and hydrophobic nanoparticles at their respective contact angle. Adapted from Binks et al. (2008).

Fig. 10. Hydrophilic nanoparticles stabilization mechanism against film rupture as liquid continuously drained from the foam film or lamellae. Adapted from
(Vatanparast et al., 2018).

the NPs become partially hydrophobic, the contact angle increases along both mechanisms, PMax c and particle detachment energy must be
with the foam stability. However, an increase of contact angle beyond considered simultaneously. A detailed explanation can be found else­
90◦ could form large aggregates. Thus reducing the foam stability as NPs where (Kaptay, 2006), but here is a summary of such analysis. Although
cannot “be attached” at the gas-liquid interface due to gravitational both detachment energy and maximum capillary pressure should pro­
force (Kruglyakov et al., 2011). Nevertheless, the enhancement of foam vide as high (positive) values as possible to ensure foam stability, it can
stability was more considerable for partially hydrophobic NPs (Zargar­ be seen that the conclusions made after detachment energy and
talebi et al., 2015). Therefore, identifying the contact angle of NPs used maximum capillary pressure are partly excluding each other. The
to maximize foam stability is crucial as it corresponds to different NP following conclusions can be drawn from detachment energy: (i) the
stabilization mechanisms of foam. particle will be stable only at the gas-liquid interface if the contact angle
In relation to the contact angle, θ, and how it affects foam stability, is significantly greater than 0◦ but significantly less than 180◦ , and (ii)

Fig. 11. The probability (PRGLi ) that the particles are stabilized at the gas-liquid interface (black curve). The probability (PRTLf ) that the liquid films are stabilized by
a single layer of particles (red curve), and the complex probability (PRfoam : PRGLi × PRTLf ) that the foams are stabilized at a liquid volume ratio of 0.5 (blue curve).
Adapted from (Kaptay, 2006). (For interpretation of the references to color in this figure legend, the reader is referred to the Web version of this article.)

8
S.A. Ab Rasid et al. Journal of Petroleum Science and Engineering 208 (2022) 109475

the particle will be most stable at the gas/liquid interface when the films will be stable under the same conditions. This complex probability
contact angle equals 90◦ , and (iii) according to numerous writers, once will be proportional to the probability that a given type of foam under
particles are stabilized at the gas/liquid interface, they will inevitably given circumstances is stable. Consequently, the stability of foams will
stabilize foams. As a result, it is critical to emphasize that this conclusion be proportional to PRGLi × PRTLf . This complex probability is shown in
does not follow from detaching energy. Maximum capillary pressure, on Fig. 11 (blue curve) as a function of contact angle. The ideal contact
the other hand, can be used to draw the following conclusions: (i) a angle for the formation of foams is about 70◦ , as shown in the figure.
higher positive value of PMax
c ensures that a thin liquid film between the This has been reported to be in complete agreement with the experi­
bubbles of a foam can withstand a higher pressing force; and (ii) from mental results.
maximum capillary pressure, one can conclude that the film’s stability is
greatest at 0◦ and decreases to zero at 90◦ .
4.3. Effects of hydrophilic nanoparticles concentration
Let us examine the PRGLi the probability that the solid particles are
stabilized by the detachment energy under the same other conditions at
The influence of NP concentration on foam stability has been
the gas-liquid interface. Fig. 11 shows the probability as a function of
demonstrated experimentally by previous researchers. Table 3 summa­
contact angle (black colored curve). The particles will be most likely to
rizes the previous findings related to the influence of NPs concentration
be at the gas-liquid interface at θ = 90◦ , as shown in the figure. From the
on foam stability. Apart from the enhancement of bulk foam stability,
maximum capillary pressure, PRTLf the probability (cos θ) that the thin
the increased average pressure drop, apparent viscosity, and the addi­
liquid film between the gas bubbles is stabilized by a single particle tional oil recovery in flooding experiments were also reported (Zargar­
layer. Fig. 11 shows this probability as a function of contact angle (curve talebi et al., 2015; Karakashev et al., 2011; Maurya and Mandal, 2018; Li
with red color). The thin liquid film between the gas bubbles is most et al., 2019; Singh and Mohanty, 2018; Rognmo et al., 2018; Sun et al.,
stable at θ = 0◦ . Now let us consider PRfoam as the probability of the 2020; Hu et al., 2018; Yang et al., 2017b; Yekeen et al., 2017a).
complex event that both the particles at the interface and the thin liquid Karakashev et al. (2011) conducted an experimental study of foam

Table 3
Table summary of the influence of nanoparticles concentration on foam stability.
Reference Nanoparticle Surfactant Main Findings

Karakashev et al. Sepiolite NPs, High purity SiO2 SDS (anionic) 1. The silica micro-spheres slightly increased the foamability at 1 wt%, 2 wt
(2011) %, and 4 wt% (with increasing NP concentration).
Kumar and Al2O3, ZrO2, CaCO3, BN, SiO2 SDS (anionic), CTAB (cationic), Polysorbate 1. The presence of NP in the surfactant solution has the strong ability to
Mandal (2017) 80 (nonionic) increase the accumulation of surface active species near the interface,
hence stable foam was obtained at optimum concentration of NPs.
Yang et al. Alumina sol (hydrophilic) SDS (anionic) 1. Foam stability was enhanced at an appropriate SDS/AlOOH
(2017b) concentration ratio.
2. Foam generated from SDS/AlOOH dispersions with higher NP
concentrations exhibited lower foamability because of increases in SDS
adsorption and viscosity.
Yekeen et al. SiO2 (hydrophilic), Modified SDS (anionic) 1. In the presence of hydrophilic SiO2 and Al2O3 NPs, foam stability (half-
(2017a) Al2O3 hydrophilic life) increased with increasing NP concentration from 0.05 wt% to 1 wt%.
Beyond 1 wt%, foam stability started to decrease with the increasing NP
concentration. This behavior can be attributed to the excessive
agglomeration of the NPs at the foam interface and bulk structure.
2. For the modified SiO2–SDS CO2 and air-foam, foam stability (half-life)
increases with the increasing NP concentration.
Maurya and SiO2 NPs SDS (anionic), CTAB (cationic) 1. The emulsion viscosity increased with the increase in silica NP
Mandal (2018) concentration.
Alyousef et al. NexSil 20 non-modified SiO2) Alcohol Ethoxylate (nonionic) 1. Increasing NP concentration does not necessarily generate a more stable
(2018) foam/emulsion. The addition of excessive solid particles or surfactants
might decrease foam stability and reduce the maximum capillary pressure
of coalescence as a result of forming large and dense aggregates.
Singh and LN1 - PEG surface coat SiO2, Anionic surfactant with a sulfonate head 1. These results show that emulsion stability can be tuned by varying the
Mohanty (2018) LN2 - GLYMO surface coat group concentration of surfactant and NPs from ultra-stable (half-life order of
SiO2 years) to moderately stable (half-life order of minutes).
2. A clear synergy between surfactant and Si-NP2 in stabilizing foam under
high-salinity conditions was observed with foam resistance factor
increasing from 1.84 (surfactant case) to 5.66 (surfactant+0.5 wt% Si-
NP2). Average pressure drops increase as the NPs concentration increases.
Rognmo et al. SiO2 A - 23.3 nm,B - 20.3 nm Linear alcohol ethoxylate 1. An increase in NP A concentration increased the oil recovery further.
(2018)
Hu et al. (2018) SiO2 NPs (SNPs) hydrophilic CTAB (cationic), SDS (anionic) 1. The initial foam volume decreases to some extent in the presence of SNP,
because the adsorption of CTAB onto SNPs reduced the number of free
CTAB molecules, and the surface tension increased.
2. CTAB molecules adsorbed on the negatively charged SNPs surface via
electrostatic interaction. This adsorption not only reduced their surface
charges but also rendered them to be more hydrophobic than the bare
surface.
Li et al. (2019) Hydrophilic SiO2 NPs (SC) HHSB 1. The coalescence rate of SC-CO2 foam gradually became very slow with
the increasing NP concentration.
2. With an increase in NP concentration, the apparent viscosity of SC-CO2
foam increased greatly.
Majed Almubarak Surface modified SiO2 Tallowtrimethylammoniumchlodride 1. The surfactant and NPs concentrations are a crucial parameter for foam
et al (2020) (cationic) stability where, an optimal concentration of NPs for strong foam
generation.
Xuezhen and Ethyl cellulose (ECNP) CH3 70PO-100EOH (nonionic) 1. The synergy between the surfactant and ECNP dramatically improved the
Mohanty (2020) foamability and stability when their concentrations were optimized.

9
S.A. Ab Rasid et al. Journal of Petroleum Science and Engineering 208 (2022) 109475

mixture, reduction of zeta potential, particle aggregation, and foam


stability.
In a positive-charge NP and anionic surfactant mixture, an electrical
attraction interaction exists. The electrical attraction between the
oppositely charged surfactant and NP causes the surfactant monomers to
be adsorbed on the NPs’ surface (Veyskarami and Ghazanfari, 2018).
Thus, an in-situ modification of NPs’ wettability from hydrophilic to
partially hydrophobic and vice versa may occur (Arriaga et al., 2012;
Llamas et al., 2019) as the opposite-charge surfactant forms a monolayer
at the NP surface (Farhadi et al., 2016; Panahpoori et al., 2019). Further
increase of NP hydrophobicity may increase contact angle above 90◦ ,
resulting in the coagulation of solid NPs (Kruglyakov et al., 2011).
Therefore, it is necessary to determine the optimum NPs and surfactant
concentration to avoid NPs becoming extremely hydrophobic, thereby
avoiding flocculation (Binks and Rodrigues, 2007).
In a similar-charge NP and surfactant system where electrical
repulsion exists, the nanofluid dispersion’s zeta potential can either
remain the same or further increase (Guo and Aryana, 2016; Vatan­
parast et al., 2018; Maurya and Mandal, 2018). A more significant
Fig. 12. Foamability reduction observed with decreasing cationic cetyl­ positive or negative value of zeta potential usually corresponds to
trimethylammonium bromide (CTAB) surfactant concentration at fixed silica smaller average particle size and less agglomeration, as shown in Fig. 13
nanoparticles (SNPs) concentration of 0.1 wt%. Adapted from Hu et al. (2018). (Panahpoori et al., 2019). Therefore, NPs remained in the lamellae, thus
stabilizing the foam through the particle arrangement mechanism by
stability using sepiolite NPs mixed in sodium dodecyl sulfate (SDS) slowing liquid drainage (Hu et al., 2018). According to Vatanparast et al.
surfactant using a foam tester. Their results show that sepiolite NPs (2018), the zeta potential of hydrophilic silica NPs in the diluted
significantly increase the foam half-life with increasing NP concentra­ nanofluid dispersion nearly remained constant by adding SDS, implying
tion at fixed surfactant concentrations. Similar bulk foam stability re­ that the surfactant molecules do not modify the surface charges of the
sults in decane were observed by Maurya and Mandal (2018). Increasing NPs. This result was somewhat predictable because of the same elec­
NP concentration has been shown to improve dynamic foam stability trical charge on the NPs’ surface and the surfactant’s head. However, in
and apparent foam viscosity (Li et al., 2019) and improve fluid diversion the presence of NPs, the surfactant performs better in terms of lowering
during the Berea core flooding experiment (Singh and Mohanty, 2018) IFT. The electrostatic repulsion between SDS molecules and NPs is
and at reservoir conditions (Rognmo et al., 2018). responsible for this behavior.
However, a previous study by Hu et al. (2018) has also reported a Although both charges interaction between NPs and surfactants
reduction in foamability when NPs and surfactant used does not have a enhance foam stability based on a different mechanism, it has been re­
similar charge, as presented in Fig. 12. As NP concentration increases, ported that a longer foam half-life is achievable between oppositely
more surfactant molecules are adsorbed on NP surfaces; consequently, charged NPs and surfactants (Hu et al., 2018; Yekeen et al., 2019a).
zeta potential reduction was observed experimentally (Yang et al., Thus, confirming that the electrostatic attraction is more prominent in
2017b). Zeta potential is an exponential difference between the outer enhancing foam stability. However, identifying the optimum concen­
boundary, an immobile stern layer of the NP, and bulk solution. tration of oppositely charged NPs and the surfactant is crucial to avoid
Reduction in zeta potential or zero zeta potential is achieved when there particle aggregation and precipitation in solution as the particles’ total
is a balance between the concentration of ions attached to the surface net charge at the NP surface is being neutralized (Esmaeilzadeh et al.,
and the surface charge of the NPs (Ahmadi et al., 2011). Since fewer 2014; Veyskarami and Ghazanfari, 2018).
surfactant molecules are available in the solution, the foamability of
NP-stabilized foam decreases with an increase in NP concentration 4.5. Effect of salinity
(Yekeen et al., 2017a; Yang et al., 2017b). Similar findings were
observed by Alyousef et al. (2018) using hydrophilic silica in three To date, some researchers had observed a detrimental effect of
non-ionic surfactants. salinity toward NP-stabilized foam performance, see Table 5. Noor et al.
In summary, it has been proven that increasing NP concentration (2019) investigated foam stability at varying Alpha Olefin Sulfonate
may enhance foam stability. However, it may also decrease its foam­ (AOS) surfactant, silica NP, and brine concentrations by observing foam
ability depending on the NPs and surfactant type used. Therefore, it is half-life and bubble size analysis at standard conditions. Their result
crucial to identify an optimum NP concentration to achieve optimum shows that adding 1.0 wt% of sodium chloride (NaCl) reduced the NPs
foamability and foam stability in combination with the surfactant used stabilized foam stability as larger foam bubbles were created, causing
before any NP-surfactant foam real application. faster foam collapse. A similar outcome was observed by Xiao et al.
(2016) through dynamic foam stability measurement. Their results show
that the maximum foam apparent viscosity was achieved at a lower foam
4.4. Synergy effect between nanoparticles surface charge and surfactant quality, indicating that the salt ions in the solution reduce NP-stabilized
net charge foam’s stability.
On the contrary, Li et al. (2019) reported from their experiments
The interaction between NP surface charge and surfactant net involving salinities between 2 wt% to 10 wt%; foam half-life signifi­
charges critically affects NP-assisted foam stability as demonstrated by cantly improved with increasing salinity, although the foamability was
the main findings in Table 4. It may be an electrical attraction between slightly decreased. Similar results were also observed by Xu et al. (2018)
oppositely charged NPs and surfactant net charge or an electrical and Qian et al. (2019) up to 5 wt% salinity.
repulsion between similar charges of the two. NPs and surfactant Different results may be due to different ionic interactions between
mixture of similar and opposite charge has been observed to enhance salt, NPs, and surfactant. Several researchers observed that increasing
foam stability compared to conventional surfactant foam (Hu et al., the salinity resulted in aggregation due to a reduction in surface tension
2018). The synergy effect can be observed through NP dispersion in a and zeta potential, as shown in Fig. 14, thus decreasing foam stability

10
S.A. Ab Rasid et al. Journal of Petroleum Science and Engineering 208 (2022) 109475

Table 4
Table summary of synergy effect between nanoparticles surface charge and surfactant net charge on foam stability.
Reference Nanoparticle Surfactant Main Findings

Binks and Rodrigues SiO2 CTAB (cationic) 1.A synergism exists between particles and surfactants, thus enhancing emulsion stability, in
(2007) contrast to complete instability of emulsion with particles or surfactant alone.
Arriaga et al. (2012) SiO2 n-amylamine (cosurfactant 1.The results (bulk behavior of mixed silica NPs-n-amylamine dispersion) confirm that n-
cationic) amylamine adsorb at the particle surface through electrostatic binding, producing in situ
particle hydrophilization, which is responsible for the observed foam stabilization.
Farhadi et al. (2016) SiO2 Ethyl hexadecyl dimenthyl 1.NPs and opposite charge surfactants have less interfacial tension. Thus, generate smaller
ammonium bromide (cationic) foam with uniform bubbles. With higher NPs concentration, more NPs can participate at the
interface and prevents coalescence. Since more aggregates are available in aqueous film, it
reduces drainage velocity, which increased bulk viscosity.
Guo and Aryana Fumed SiO2 T30, AOS (anionic), 1. It is proposed that particle size and distribution affect NP-assisted foam stability and the
(2016) Nano clay (by Sigma SDS (anionic), particle surface charge neutralization leads to particle aggregation.
Aldrich) LAPB (amphoteric) 2. Addition of surfactant leads to its physical adsorption on NPs’ surface, reducing surface
energy of NPs and particle-particle interactions. It increases the colloidal stability of the NPs
in suspension, and the layer of surfactant helps prevent the particles from aggregating.
Hu et al. (2018) SiO2 (SNPs) CTAB (cationic), SDS (anionic) 1. SNPs-CTAB adsorbed at the air–water interface and formed a thick, compact, NP-enhanced
bubble film, which significantly slowed drainage, gas diffusion, and coalescence and then
increased the foam stability.
2. Even if the SNPs in the existence of SDS were hydrophilic, the foam stability was still
enhanced since these particles presented in the aqueous phase of the foam film and slowed
down the drainage.
Veyskarami and SiO2 SDS (anionic), CTAB (cationic) 1. Opposite-charged surfactant could make an in-situ modification on silica NP surface
Ghazanfari (2018) through adsorption of surfactant monomers on NP surface. Linking opposite-charged NP and
surfactant can cause aggregation and precipitation of NPs in the bulk solution.
2. No interaction between like-charged surfactant and NP was observed.
Vatanparast et al. SiO2 (9 nm & 30 SDS (anionic), 1. In the presence of NPs, the surfactant behaves more efficient in reducing the IFT. This
(2018) nm) DBSA (anionic) behavior is attributed to the electrostatic repulsion between SDS molecules and NPs,
promoting surfactant adsorption.
Maurya and Mandal SiO2 SDS (anionic), 1. It has been found that cationic surfactant forms highly stable emulsion due to the electric
(2018) CTAB (cationic) attraction between the CTAB and SiO2.
2. Whereas, in the case of anionic surfactant, emulsion stability depends on the relative
concentration of SDS and SiO2 NP due to electrostatic repulsion between SDS and SiO2. In
NP–surfactant solution, the interfacial tension (IFT) may change if there is electrostatic
repulsion between similar-charged NP and surfactant.
Llamas et al. (2019) TiO2, SiO2 CTAB (cationic) 1. When both types of NP are made amphiphilic, synergetic effects have been observed due to
the formation of mixed NP layers at the liquid-air interface. These synergetic effects on the
surface properties of these mixed dispersions have been then exploited to obtain foams,
characterized by excellent stability features.
Panahpoori et al. TiO2 CTAB (cationic) 1. Absorbance spectrum analysis for nanofluids and NP-CTAB solutions proved the role of
(2019) CTAB molecules in NPs stabilization, possibly by the development of repulsive forces between
surface-modified NPs.
Yekeen et al. (2019a) Al2O3, CuO, SDS (anionic), 1. The result confirmed that the influence of electrostatic attraction between the anionic
TiO2, ZnO, SDBS (anionic), surfactant and opposite charge NPs could be dominant over the influence of electrostatic
ZrO2, SiO2 CTAB (cationic) repulsion (similar charge surfactant and NPs) in promoting foam stability.
Fu and Liu (2020) SiO2 SDS (anionic), 1. Anionic surfactant yielded the most stable foam (highest apparent viscosity) compared to
CTAB (cationic), the cationic and nonionic surfactant.
Ecosurf EH-9 (nonionic)
Zhao et al. (2021) SiO2 A (NPA), SiO2 AOT (anionic) 1. Hydrophilic NPA can enhance the surface activity of AOT surfactants due to electrostatic
B (NPB) repulsion force and the increase of ionic strength.
2. Partially hydrophobic NPB has synergistic effects with AOT for stabilizing CO2 foam when
they are mixed in proper proportions.

(Vatanparast et al., 2018; Eide et al., 2018). It is difficult to assume that stabilized foam by NPs at high salinity conditions. Therefore, further
the aggregation of NPs in the presence of salts necessarily leads to the research may be conducted to understand the actual mechanism be­
formation of unstable foams. Although a high accumulation of NP par­ tween NPs and salt ions in the presence of surfactant molecules and their
ticles in the foam structure and at locations other than the plateau in­ effect on NP fluid mixture stability and the stabilized foam performance.
terfaces and the gas-liquid interface leads to foam instability because NP
hinders the presence of particles in the interface, a moderate accumu­
lation of NP particles in the gas-liquid interface leads to higher foam 4.6. Effect of oil
stability, even in the presence of salts.
The monovalent salt ions have been reported to have a lesser impact Oil is destructive toward conventional surfactant foam and can cause
on foam stability as compared to divalent salt ions, as exhibited in them to break down instantly (Eftekhari et al., 2015). However, the
Fig. 15(Yekeen et al., 2017b). This is due to the effectiveness of divalent addition of NP to the surfactant foam was able to reduce the detrimental
cations screening NP charges being higher than monovalent cations impact of oil on foam stability (Eftekhari et al., 2015; Razali et al., 2018;
(Jafari Daghlian Sofla et al., 2018; Yang et al., 2017b; Yekeen et al., Rognmo et al., 2018; Qian et al., 2019; Yang et al., 2017b; Veyskarami
2017b; Eftekhari et al., 2015; Eide et al., 2018). When divalent salt ions and Ghazanfari, 2018; Noor et al., 2019; Singh and Mohanty, 2017b).
completely shield the electrical double layer of NPs, they can destabilize Table 6 provides a summary of previous studies related to the influence
the NPs due to aggregation of the particles. When this happens, fewer of oil on foam stability.
NPs are available in the lamellae when the foam is generated, resulting Yang et al. (2017b) conducted an experimental investigation on
in poor foam stability. Although the NPs’ actual interaction with cations foam stability in the presence of kerosene oil at room temperature by
salt was unclear, laboratory analysis had successfully created a highly observing the foam half-life of Sodium Dodecyl Sulfate (SDS) anionic
surfactant stabilized by alumina NPs. Their results show that the foam

11
S.A. Ab Rasid et al. Journal of Petroleum Science and Engineering 208 (2022) 109475

Fig. 13. Zeta potential and corresponding average particle size of 0.03 wt% TiO2 nanofluid at variable pH reproduced from Panahpoori et al. (2019).

half-life in the presence of NPs was longer, indicating better foam sta­ Further investigation may be conducted to understand the effect of oil
bility. Similar results were also reported by other researchers using a and the NP stabilization mechanism involved during static stability and
wide range of oils viscosities from light to heavy oils having viscosity up dynamic stability test.
to 508 cp (Eftekhari et al., 2015; Qian et al., 2019; Veyskarami and
Ghazanfari, 2018; Noor et al., 2019) and even at high-temperature 4.7. Effect of temperature
conditions of 110 ◦ C (Razali et al., 2018). The stability improvement
of NP-stabilized foam in the presence of oil was also observed in flow The foam stability in the presence of NPs has been reported to
experiments by showing higher apparent viscosity and additional oil decrease with increasing temperature, see Table 7. Chea et al. (2018)
recovery (Risal et al., 2018; Yang et al., 2017a; Xuezhen and Mohanty, investigated the foamability and stability of silica and alumina NPs
2020). Interestingly, Singh and Mohanty (2017b) reported a complete stabilized foam at temperatures ranging from 25 ◦ C to 90 ◦ C in the
blocking by NPs stabilized foam during flow experiments indicating that presence of anionic surfactant. They observed that foam half-life de­
the oil being swept does not break the stabilized foam generated. creases as temperature increases at constant NP and surfactant con­
Furthermore, in a study by Xu et al. (2020) the oil washing efficiency of centration and salinity. Similar results were observed when silica and
the foam was greatly increased after the addition of NPs. This is due to alumina NPs were used, which is also in agreement with findings by
the fact that the NPs can be irreversibly adsorbed onto the liquid film of Emrani et al. (2017) for tests with silica and anionic surfactant and Singh
the bubble interface, and a dense layer is formed at the interface via and Mohanty (2017a, 2017b) using surface-modified silica in anionic
flocculation, increasing the viscoelasticity of the surfactant-NP foam’s surfactant at 80 ◦ C. Likewise, several investigations on NPs stabilized
liquid film. foam core flooding experiments reported that the mobility resistance
Furthermore, a study by Veyskarami and Ghazanfari (2018) clearly factor significantly decreased at 50 ◦ C (Alyousef et al., 2017) and 80 ◦ C
reveals that the presence of oil significantly reduces foam stability of (Wang et al., 2017), indicating weaker foam at the higher temperature.
silica NP–CTAB and silica NP–SDS see Fig. 16. However, increasing the Mo et al. (2014b) found a decrease in oil recovery by NP-stabilized
concentration of silica NP in the foam structure enhances stability in the foam as temperature increased in an investigation of the temperature
presence of oil substantially. As a result, it can be stated that if the NP is effect on residual oil recovery by CO2 foam in the presence of silica NPs
suitably selected for addition to an surfactant foam structure, the foam in sandstone cores. Their findings show that silica-stabilized foam’s oil
stability at particular NP concentrations will be greatly enhanced. recovery decreases by almost 5% as the temperature increased from
However, similar to conventional foam, lighter oil, either synthetic or 25 ◦ C to 60 ◦ C. Similar outcomes were reported by Fu et al. (2018) using
crude oil, is more detrimental to foam stability than heavier oil (Singh Berea sandstone core at 2 wt% NaCl.
and Mohanty, 2015; Maurya and Mandal, 2018; Yekeen et al., 2016). In Temperature influences the diffusion rate, surfactant adsorption at
addition, NPs stabilized foam was reported to collapse faster as more oil the gas/water interface and rock surface, and liquid viscosity, all of
is available (Bashir et al., 2019). which contribute to foam stability. As temperature increases, surfactant
Besides, Veyskarami and Ghazanfari (2018) reported contradicting molecules’ mobility increases; therefore, fewer surfactant molecules are
foam stability performance between static stability test and foam flow adsorbed at the gas-liquid interface (Yang et al., 2017b; Zhu et al.,
test when identifying a suitable combination of NPs and surfactants in 2019). This finding was confirmed by a laboratory experiment where
the presence of oil. His findings show that like-charge NPs and surfac­ Emrani and Nasr-El-Din (2015) observed that the surface tension of AOS
tants exhibit better half-life in the presence of kerosene oil with a vis­ surfactant increases at higher temperature conditions above 100 ◦ C.
cosity of 1.6 cP. Remarkably, the opposite behavior is observed during Since CO2 gas has a lower solubility in liquid at a higher temperature, it
foam flow in the Hele-Shaw cell, as the opposite-charge NPs and sur­ further elevates the increasing behavior of surface tension. In addition,
factants display higher differential pressure in the presence of the same NPs’ mobility also increases as temperature increases; thus, particles
oil. It indicates how dynamic stability analysis is necessary to distinguish were inhibited from being adsorbed onto the gas-liquid interface, thus
ideal parameters or various NPs and surfactants to be used in conjunc­ decreasing foam stability (Yu et al., 2012b). It has also been discovered
tion with oil type for field application. There is still a lack of under­ that a decrease in liquid viscosity caused by an increase in temperature
standing of this phenomenon due to limited references available. reduces the drainage half-life (Emrani et al., 2017).

12
S.A. Ab Rasid et al. Journal of Petroleum Science and Engineering 208 (2022) 109475

Table 5
Table summary of the influence of salinity on foam stability.
Reference Nanoparticle Surfactant Salt Main Findings

Xiao et al. (2016) SiO2 ANS (anionic) NaCl 1. The salinity significantly negatively affected the apparent viscosity of foams
generated with NP + ANS systems.
2. In the presence of salt, maximum apparent viscosity is achieved at lower
foam quality.
Yekeen et al. Hydrophilic SiO2, SDS (anionic) NaCl, 1. There is a transition NaCl, CaCl2, and AlCl3 salt concentration for foam
(2017b) Hydrophilic Al2O3 CaCl2, stabilized by NPs and surfactant, beyond which the foam stability increased
AlCl3, with increasing salt concentrations. The presence of salt improved the foam
stability due to the increased NPs hydrophobicity and the reduced
electrostatic repulsion between the particles, which resulted in particles
aggregation at the gas-liquid interface of the foam.
2. The rate of bubble coalescence was significantly lower in monovalent salt
ions than the divalent and trivalent salt ions.
Eide et al. (2018) Modified SiO2 (Levasil NaCl 1. Divalent salt ions are more detrimental to NP stability than monovalent
CC301) ions, most likely due to a reduction in zeta-potential that causes aggregation
Jafari Daghlian Hydrophilic SiO2 NaCl, Na2SO4, 1. Divalent salt ions screen the charge of silica NPs more effectively than
Sofla et al. (2018) MgCl2, MgSO4 monovalent ions where divalent ions in the electrical double layer of NPs can
destabilize the NPs in the solution.
Xu et al. (2018) SiO2( AES (anionic-nonionic NaCl, 1. High salinity (≥50,000 mg/L) is conducive to the stability of S-AK/AES
S-AK) surfactant), SDS (anionic) CaCl2 natural gas foams at a high temperature of 50 ◦ C.
2. The addition of NaCl or CaCl2 promotes the interaction between S-AK and
AES, improving the structural strength of the foaming agent.
Vatanparast et al. Hydrophilic SiO2 SDS (anionic), NaCl 1. The impact of NPs on the interfacial tension of the surfactant solution
(2018) 9 nm & 30 nm DBSA (anionic) weakens in the presence of NaCl, and it completely vanishes at a high
concentration. Therefore, at the high salt concentration, the electrostatic
repulsion almost disappears due to the entire surface screening of the
negatively charged NPs’ by Na+ ions.
Qian et al. (2019) Alumina-coated SiO2, AES (anionic-nonionic NaCl 1. Salinity increased the performance of partial monolayer coverage NP-
SNOWTEX-AK (ST-AK) surfactant) surfactant foam but decreased the performance of surfactant-only foam. The
synergistic interaction between NPs and surfactants is responsible for this
difference.
Noor et al. (2019) SiO2 (SNP) AOS (anionic) NaCl 1. Adding brine to the solution with dispersed SNP can reduce the foam
stability due to a decrease in the particles’ electrostatic repulsion.
Li et al. (2019) Hydrophilic SiO2 HHSB (amphoteric) NaCl, 1. The stability of super critical-CO2 foam improved with the increase in
CaCl2 salinity due to the screened electrostatic repulsion between negatively
charged silica NPs.
2. The super critical-CO2 foam volume decreased slightly with an increase in
salinity, and the increased foam’s half-life.
Wang et al. (2020b) SiO2 (Ludox CL) Sulfobetaine (LHSB) NaCl, 1. High concentration of NaCl and CaCl2 reduce the adsorption of LHSB on NP.
CaCl2 2. Significant adsorption reduction is observed in CaCl2, thus less NP can be
adsorbed at the gas-liquid interface.

Although the increasing temperature has a detrimental effect on is commonly measured during foam flow experiments using a capillary
foam stability in the presence of NPs, experimental studies have proven tube (Xiao et al., 2016) or flowing in a porous media (Rognmo et al.,
that NPs still have significantly enhanced foam stability as compared to 2017; Yekeen et al., 2017b). Table 8 summarizes the effect of flow ve­
conventional surfactant foam (Bashir et al., 2019; Yang et al., 2017b; locity on foam stability. Apparent foam viscosity is considered to
Zhu et al., 2019; Li et al., 2019). Bashir et al. (2019) experimental results represent the foam stability in this condition. The flow velocity in­
show that the NPs can extend foam half-life and improve foam stability fluences the NPs-stabilized foam stability and describes the rheological
compared to surfactant foam at an optimum NP concentration at 85 ◦ C. behavior of the foam flow. A condition in which apparent foam viscosity
Similar results were observed by Yang et al. (2017b) in the presence of decreases with increasing velocity describes the shear-thinning flow
NPs as the temperature increased from 22.5 ◦ C to 80 ◦ C. At an optimum behavior (Rognmo et al., 2017). On the contrary, the condition at which
NP concentration, an optimal amount of NPs can be adsorbed at the apparent foam viscosities increase with increasing flow velocity is
gas-liquid interface to enhance the disjoining pressure, thus slowing known as shear-thickening flow behavior.
down foam collapse at high temperatures (Li et al., 2019). Surfactants, Several studies have experimentally reported a decrease in apparent
on the other hand, can decompose at high temperatures, rendering them foam viscosity with increasing flow velocity during foam flow experi­
ineffective for foam stabilization. Protecting surfactants from degrada­ ments. Through co-injection flow experiments, Yekeen et al. (2017b)
tion and improving foam stability are critical factors that can be ach­ reported that the apparent foam viscosity reduces with increasing in­
ieved by adding NPs to the solution (Emrani et al., 2017). jection flow rates. The results obtained can attribute to shear-thinning
In summary, NP-stabilized foam stability decreases with increasing behavior similar to conventional surfactant foam.
temperature, similar to surfactant foam behavior. However, it has been Xiao et al. (2016) studied the effect of shear rates on foam rheology
reported that NPs improve the foam stability compared to surfactant at different foam qualities and salinities in the presence of silica NPs
foam at high-temperature conditions, at the optimum concentration. using the flow loop apparatus at 1140 psi and temperature of 40 ◦ C.
However, limited research is available in studying the foam stability They reported that the foam exhibited shear thinning behavior at
effect at a temperature above 90 ◦ C. As a result, additional research and all-foam qualities. However, different rheology behavior was observed
studies are required to improve the understanding of foam stability for at high salinity conditions (5 wt%), whereby foam exhibits pure shear
those high-temperature field applications. thickening behavior at all-foam qualities in the absence of oil. In addi­
tion to that, they also identified a critical shear rate at which
4.8. Effect of flow velocity NP-surfactant foam was stable in high salinity conditions.
Based on the discussed findings, it can be deduced that the shear rate
The effect of flow velocity or injection velocity toward foam stability is not the only parameter affecting phenomenon characteristics (shear-

13
S.A. Ab Rasid et al. Journal of Petroleum Science and Engineering 208 (2022) 109475

thinning or shear-thickening) and precisely the magnitude of the wt%, 0.2 wt% and 1.0 wt%), and particle sizes (5 nm and 25 nm).
apparent viscosity. The foam rheology behavior may differ at different However, their observation shows that all NPs tested, regardless of
NPs or surfactant concentrations (Maurya and Mandal, 2018) and particles concentration and particles sizes, do not significantly adsorb
different salinity (Xiao et al., 2016). It is commonly understood that onto the mineral surfaces.
foam without NPs exhibits shear thinning behavior, but there is still a Based on previous researches, NP retention via mechanical trapping
lack of understanding of the extent of shear thickening behavior in is insignificant as the NPs’ size is several magnitudes smaller than the
NP-stabilized foam, which requires in-depth investigation. pore throat size (Sun et al., 2021). The trapping of NPs is commonly
examined through flooding experiments and analyzed based on differ­
4.9. Effect of nanoparticles retention ential pressure, produced NPs concentration, and permeability of the
media pre- and post-flooding. Rognmo et al. (2017) observed that the
NPs retention in porous media indirectly reduces foamability and silica NPs breakthrough at approximately one pore injection volume
foam strength (Rognmo et al., 2017). A summary of the findings related during the NPs flooding experiments. The author obtained an average
to the influence of NPs retention on foam stability is provided in Table 9. NPs retention of 401 μg/g from four core flood experiments permeability
NPs are retained in porous media via three different mechanisms; ranging between 1506md and 2447md. In addition to that, no change in
attachment to rock surface due to electrostatic interaction results in the differential pressure measurements during NPs flooding indicating
particle adsorption, mechanical entrapment and log-jamming result in that NP retention is not due to mechanical trapping. However, Sun et al.
pore channel plugging, and particles settling due to gravitational effect (2021) observed that the NPs retention decreases from 12.1% to 4.1%
caused by large density differences between particles and aqueous with increasing sand pack permeability ranging between 231md and
phase, see Fig. 17. The retention mechanism is a function of NPs 4320md. The author also reported that a 10.2% decrease of permeability
adsorption to either the rock surface or the liquid interface (Metin et al., was observed a post-brine flooding in the 231md sand pack.
2012). It is preferable to have NPs adsorbed at the liquid interface rather On the other hand, Yu et al. (2012a) observed no change in differ­
than the rock surface in a system. Excessive NP retention can negatively ential pressure and permeability during silica NPs flooding using sand­
affect the reservoir’s porosity, permeability, wettability, and flow ca­ stone core with 33md and 57.6md permeability. With no permeability
pacity (Rognmo et al., 2017; Sun et al., 2021; Yu et al., 2012a). The impairment, a slight change in differential pressure during limestone
degree of retention should be minimized to maintain in-depth foam­ (132md) core flooding was observed. Therefore, the NPs were adsorbed
ability and keep costs low. Therefore, knowledge of nanofluid retention due to their electrostatic interaction with rock surfaces instead of me­
characteristics in the porous media is essential when implementing chanical trapping. Interestingly, although the silica NPs breakthrough
NP-surfactant foam flow strategies. immediately during the dolomite core flooding, differential pressure
Adsorption of surface-active agents, like surfactants and NPs, occurs during brine flooding after NP flooding continuously increases. This
when surface-active agents dissolved in an aqueous phase are injected response indicates that particle plugging occurred and permeability is
into a porous media and attract the pore wall instead of migrating to the changed. The result obtained may be due to enhanced interaction be­
gas-liquid interface. Adsorption adversely affects the foaming process tween the nano-silica and pore surface in the extremely low perme­
and may result in ineffective and uneconomical use of foam stabilizers as ability (5.29md) dolomite core. This low permeability may result in
the amount of surface-active agents available for stabilization decreases higher interactions between moving particles and pore surfaces that
with an increasing degree of adsorption. The adsorption of NPs onto rock promote pore-surface processes like a deposition and pore-throat pro­
surfaces may vary between different rock types and minerals due to cesses like plugging, screening, and bridging.
electrostatic interaction bet ween the NPs and the rock surface. Yu et al. The NPs retention via particle adsorption and mechanical trapping is
(2012a) study the silica NPs adsorption on sandstone, limestone, and minimal based on the reviewed literature. Different rock types and
dolomite. The equilibrium adsorption results show that NPs adsorption mineral compositions play a vital role in the adsorption of NPs onto rock
is highest on limestone, which mainly consists of calcite (5.501 mg/g) surfaces due to electrostatic interaction. Interestingly, different surface
compared to sandstone (1.272 mg/g) and dolomite (0 mg/g). The author coat nanosilica, NPs concentration, and particle size do not significantly
highlights that the low adsorption of nano-silica with sandstone may be affect the adsorption of NPs onto the rock surface. Since the reviewed
due to similar composition, while the high adsorption observed with researchers utilize silica NPs in their study, the degree of NP adsorption
limestone is due to high electrostatic force between the two surfaces. may differ using other types of NPs.
Additionally, Metin et al. (2012) conducted adsorption experiments Currently, only the total amount of retention can be calculated from
using calcite and quartz sand. The author extended the analysis to the analysis of dynamic test results without determining the contribu­
observe the effect of different surface coat particles (unmodified SiO2, tion of each retention mechanism to the total amount of NPs lost in the
sulfonate-modified, and PEG-modified SiO2), NPs concentration (0.04 porous medium. It is essential to differentiate the contribution of each of

Fig. 14. Schematic of electrostatic repulsion behavior in the absence (left) and presence of salt ions (right). Adapted from Vatanparast et al. (2018).

14
S.A. Ab Rasid et al. Journal of Petroleum Science and Engineering 208 (2022) 109475

adsorption onto the surface sample (sandstone, carbonate, and


bentonite) decreases with increasing silica NPs concentration at fixed
surfactant concentration. The author pointed that the observed behavior
was due to stronger hydrogen bonding between surfactant and NP
(Ahmadi and Shadizadeh, 2013; Saxena et al., 2019). Since competitive
adsorption exists between NPs and rock surface, higher surfactant
molecules adsorption onto NPs surface will lead to lesser surfactant
adsorption onto the rock surface. Therefore, more surfactant molecules
will be available in the bulk liquid, thus reduce the loss of surfactant
onto rock surface and improving foam performance.

4.10. Effect of nanoparticle types

The use of NPs to enhance the stability of foam is not limited to SiO2.
Several researchers have studied the effect of metal oxides NPs (Yekeen
et al., 2019a), fly-ash NPs (Wang et al., 2020a; Lee et al., 2017; Phong
et al., 2020), and carbon nanotubes (Raja Ramanathan and Nasr-El-Din,
2020). A summary of findings related to the influence of NPs type on
foam stability is provided in Table 11. Overall, all the NPs used exhibit
enhancement of foam stability compared to surfactant foam. The
enhancement of foam stability is a function of NPs’ characteristics such
as density, size, shape, wettability, and surface charges that contribute
to the stabilization mechanism (Rafati et al., 2016).
Yekeen et al. (2019a) conducted extensive research on the effect of
surfactant (anionic, cationic, and nonionic) and NP types (Al2O3, ZrO2,
SiO2, TiO2, ZnO, CuO, and carbon nanotubes). It was observed that the
electrostatic attraction between the NPs surface charge and surfactant
Fig. 15. Effect of (a) monovalent salt ions (Na⁺) and (b) divalent salt ions (Ca2⁺) head-groups play a vital role in stabilizing the foam. The interaction
concentration on the foamability of the nano-ash-AOS mixture. Reproduced allows better NP adsorption at the gas-liquid interface and tighter
from Eftekhari et al. (2015). packing at the lamellae, thus slowing down the liquid drainage and
enhancing foam stability. In a foam half-life measurement using fly ash
these mechanisms in retention because if they are the significant amount NPs combined with MFOMAX surfactant, Phong et al. (2020) emphasize
and needed to be treated, the method of reducing or eliminating the that any NPs can be used to stabilize surfactant foam. It is because of the
impact of these mechanisms varies depending on their type. Mechanical characteristic of irreversible NPs adsorption at the gas-liquid interface
entrapment, for example, can be minimized or even eliminated by by acting as a barrier between the fluid to reduce the liquid drainage and
reducing the NP size, whereas adsorption can be reduced by altering the inhibit bubble coalescence. A similar observation was made by Wang
NP or coating the NP’s surface with chemicals that repel the NP from the et al. (2020a) using fly ash combined with AOS and SDS surfactant.
rock surface. The findings of polymer flooding dynamics research Recent studies have evaluated the ability of carbon nanotubes in
(Akbari et al., 2019b; Al-Hajri et al., 2018), for which such a separation enhancing foam stability. Raja Ramanathan and Nasr-El-Din (2020)
of mechanisms has been done, can be used to design such a method, compares the foam performance between silica and concentration
which could be the focus of future investigations. It should be noted that multiwalled carbon tube (MWCNT) combined with AOS surfactant
while static tests can provide a general and qualitative overview of the through foam half-life measurements. The results show that MWCT
amount of adsorption on the rock surface, the quantitative use of static exhibit 4.6 times foam stability compared to silica NPs (2.9 times).
test results to evaluate adsorption in the dynamic state is associated with Concurrently, the MWCNT exhibits a higher mobility reduction factor
a high risk of error for three main reasons. First, exposed surfaces may (5.5) than silica (3.5) during the foam flooding experiments. The author
not have similarities to the initially consolidated surface surfaces during emphasize that the different performance observed were attributed by
the disaggregation of consolidated rock. Second, the static test does not the MWCNT’s characteristic (rod-like structure, size, zeta potential, and
measure the loss of mechanical entrapment in porous media. Third, the partial hydrophobicity) compared to silica NPs.
disaggregated rock’s wettability may differ from the wettability of the The NP’s characteristics mainly contribute to the improvement of
reservoir rock (Manichand and Seright, 2014; Dang et al., 2014). foam stability in the presence of NPs. The potential use of other NPs,
Due to the strong electrostatic adsorption between opposite charge especially the abundant by-product particles such as fly ash, should be
NP and surfactant molecules, competitive adsorption of surfactant further explored and studied for potential low-cost products and
molecules with NPs and rock surface exists (Yekeen et al., 2017b). sustainability.
Insignificant loss of surfactant molecules onto the rock is reported
through several experimental investigations, see Table 10. Yekeen et al. 5. Conclusion
(2019b) observed that the amount of free surfactant molecules decreases
in the presence of NPs. The silica NPs were more effective in reducing After reviewing the effects of the main parameters involved in the
cationic surfactant adsorption onto surface samples (clay-rich shale and surfactant foam stabilization by NP, the general conclusion of this paper
quartz-rich shale) compared to anionic and non-ionic surfactants. In is that increasing the stability of the surfactant by adding NP is possible
addition to that, Saxena et al. (2019) observed that the surfactant in almost every situation. However, the degree of stability enhancement

15
S.A. Ab Rasid et al. Journal of Petroleum Science and Engineering 208 (2022) 109475

Table 6
Table summary of the influence of oil on foam stability.
Reference Nanoparticle Surfactant Oil Type Main Findings

Eftekhari et al. Fly ash coal AOS (anionic) Shell Ondina 1.Bulk NP-foam shows higher stability in the presence of hexadecane and Ondina
(2015) 919, 919 compared to surfactant foam.
Hexadecane 2. NP can be used to stabilize foam in the presence of crude oil at high
temperatures and pressure.
3.The mechanism of the nano ash foam stability in the presence of oil is not clear. It
may be due to the presence of NPs at the gas− liquid interface preventing the oil
droplets from entering and breaking the lamellae.
Singh and Mohanty SiO2 α-olefin sulfonate Crude oil: 1.The foam volume decreased drastically in crude oil A, indicating that the foam
(2015) Nyacol DP 9711, (anionic) A (9cp), (with and without NP) unstable.
SiO2 B (382cp), 2. The foam half-life increased with crude oil B and C.Higher foam stability was
PEG coating, C (36.2cp) observed in surfactant− NP blends compared to surfactant foam.
3.With and without NP, lighter crude oils are more destabilizing to foams than
heavier oils.
Yang et al. (2017b) AlOOH CTAB (cationic) Kerosene 1.The SDS/AlOOH foam exhibited better stability than the SDS foam at high
temperatures and in the presence of oil.
Yekeen et al. (2016) unmodified SiO2, SDS (anionic) Hexadecane 1.The bubble size distribution of SiO2-SDS-foam in the presence of paraffin oil and
modified SiO2 NP (3.3cp), hexadecane oil shows that foam bubble size increases with decreasing of oil
Decane (0.9cp), viscosity and density.
Paraffin oil 2.Oils with low viscosity and density are more detrimental to foams and the rate of
(20.8cp) coalescence and coarsening reduces in the presence of NPs.
3.The addition of SiO2 NPs into the surfactant solution could improve foam
stability in the presence of oil.
Razali et al. (2018) SiO2 A300 MFOMAX Light crude oil 1. Hydrophilic SiO2 has significantly improved the foam half-life by 2 times longer
Hydrophilic, zwitterionic) (45◦ API) than the surfactant alone in the presence of light crude oil.
SiO2 R816 2.The SiO2 (hydrophilic) NPs have significantly reduced the detrimental effects of
Hydrophobic, light crude oil and strengthen the foam. The foam generated is more wet and stable
ZnO, TiO2 in the presence of light crude oil.
Rognmo et al. (2018) SiO2 (23.3 nm & 20.3 Linear alcohol Crude oil (7.6cp) 1. The pressure gradients of SiO2 NPs stabilize foam in the presence of oil and
nm) ethoxylate remain stable during oil displacement.
Maurya and Mandal SiO2 SDS (anionic) Decane 1.The emulsion containing paraffin oil shows better stability as compared with
(2018) (0.859cp), decane oil after 10 days.
Paraffin (56cp) 2.Paraffin oil provides a higher viscosity to the stabilized emulsion. The increase in
viscosity is mainly due to an increase in the viscosity of the dispersed phase.
Veyskarami and SiO2 (SNP) SDS (anionic), Kerosene (1.6cp) 1. Static stability analysis revealed that both in situ modified and unmodified SNP
Ghazanfari (2018) CTAB (cationic) slightly improved foam stability in the absence and presence of oil.
2. In foam static stability analysis, the like-charged NP and surfactant system gave
a better performance than the opposite-charged one in the absence and presence of
oil.
3. Foam flow in Hele–Shaw cell shows that the foam of SiO2 NP and like-charged
surfactant showed lower mobility in the absence of oil, and opposite-charged
surfactant–NP foam lowers mobility in the presence of oil.
4. Foam flow analysis suggest opposite-charged NP and surfactant systems are
more attractive than like-charged ones for EOR application.
Singh and Mohanty Polyethylene glycol- Amphoam (anionic) Crude oil (32cp) 1. Significant crossflow of oil from low-permeability to high-permeability was
(2018) coated SiO2 NP1, observed for the case of surfactant foam flood. Conversely, the Si-NP2 resulted in
GLYMO-coated SiO2 no crossflow with complete blocking of the high-permeability region due to the
NP2 formation of an in-situ emulsion. Such selective plugging of high-permeability
channels via NPs with optimum surface coating could potentially recover oil from
heterogeneous reservoirs.
Noor et al. (2019) SiO2 SDS (anionic) White mineral oil 1. Adding NPs at specific concentrations to synthetic saltwater and oil shows
promise when it comes to increasing the half-time of foam. It might help slow the
drainage of thin aqueous film and, therefore, produce a more stable foam.
Qian et al. (2019) Alumina-coated silica, AES (anionic- Mineral oil 1. NP-surfactant foam is highly stable at partial monolayer coverage and uniquely
SNOWTEX-AK (ST-AK) nonionic surfactant) (18cp), stable when contacting an oil phase.
Heavy oil 2. NPs present in the foam structure prevented oil from entering the foam
(508cp) structure.
Bashir et al. (2019) Fumed SiO2, AOS (anionic) North Sea oil 1. CO2-foam can show better stability in the presence of oil when the optimum
Rice husk ash (RHA) (54.71cp) concentrations of SiO2/RHA NPs (0.2 wt%) and xanthan gum (0.3 wt%) were
added.

depends on several parameters and covers a wide range from low to 2. Both hydrophilic and partially hydrophobic NPs can improve
high. The parameters studied in this work include NP properties (i.e., foam stability. However, the improvement is more considerable
size, type, surface wettability) and reservoir properties (i.e., salinity of with partially hydrophobic NPs.
formation water, presence of oil, reservoir temperature), process pa­ 3. Increasing NP concentration does not necessarily improve foam
rameters (i.e., NP concentration and flow rate), synergistic effects be­ stability. This behavior is highly dependent on the types of NPs
tween the surface charge of NPs and the net charge of surfactant, and NP and surfactant used due to their complex synergic interaction.
loss in the porous medium. The key findings are summarized below: Preliminary NPs and surfactant type and concentration screening
are crucial to achieving optimum enhancement of foam stability.
1. The NPs stabilized foam stability is inversely proportional to NP 4. Similar charges lead to repulsion between the surfactant and the
size. The NP size, which has been proven to improve foam half- NP, causing the surfactant to move toward the gas-liquid inter­
life, ranges between 7 nm and 10 μm. face and decrease the interfacial tension. When surfactant and NP

16
S.A. Ab Rasid et al. Journal of Petroleum Science and Engineering 208 (2022) 109475

Fig. 16. Effect of nanoparticles concentration on foam stability of silica nanoparticle–CTAB and silica nanoparticle–SDS a) in the absence of oil, and b) in the
presence of oil at different surfactant concentrations. Adapted from Veyskarami and Ghazanfari (2018).

Table 7
Table summary of the influence of temperature on foam stability.
Reference Surfactant Temp (◦ C) Main Findings

Yu et al. (2012b) SiO2 25, 45, 60 The height of SiO2–CO2 foam decrease with increasing temperature. As the temperature
reaches 60 ◦ C, no CO2 foam was observed in the sapphire tube.
Mo et al. (2014a) SiO2 25, 45, 60 1.The residual oil recovery by CO2/SiO2 flooding decreased from 62.6% to 52.1% when
the temperature increased from 25 ◦ C to 60 ◦ C.
Emrani and SiO2, Fe2O3 AOS (anionic) Ambient to 1.At ambient temperature, NPs promote CO2-foam formation in AOS solutions. Adding a
Nasr-El-Din 150 small amount of NPs (0.1 wt%) to AOS solutions could improve the CO2-foam stability.
(2015) 2. NPs can help to stabilize the system faster at 150 ◦ C.
Yang et al. (2017b) AlOOH SDS (anionic) 22.5, 80 1.The SDS/AlOOH foam exhibited much better stability than the SDS foam at high
temperatures and in the presence of oil.
2.Increasing temperature accelerate the drainage rate.
Emrani et al. (2017) SiO2 AOS (anionic) 25, 65.5 1.The results of this work show that foam half-life decreases as the temperature increases
from 25 to 65.5 ◦ C.
2. NP-based foams are more stable over time compared to surfactants foam. Adding SiO2
NPs to the AOS solution improves the foam stability and increases the MRF eight-fold.
3. NP solutions can be used to improve CO2 foam stability.
Singh and Mohanty PEG surface coat Anionic surfactant 25, 80 1.The aqueous stability and salt tolerance of both NPs decreased with an increase in the
(2017a) SiO2, (sulfonate head group) temperature.
GLYMO surface 2.The decay of foam was increased significantly at the higher temperature of 80 ◦ C as
coat SiO2 compared to 25 ◦ C.
Alyousef et al. Surface-modified CNF (anionic) 25, 50 1.The foamability tests showed that the rise in temperature reduces the pressure drop in
(2017) SiO2 the core flood experiment, which can be attributed to the reduction in foam stability and
strength.
Wang et al. (2017) Fumed SiO2, DDAB, 40, 80 1.The amphiphilic particles foam system can stabilize the foam better under two
T30 SiO2, H34 CAB temperatures.
SiO2 2.The resistance factors of surfactants and the mixture of NPs and surfactants decreased
significantly at 80 ◦ C as compared to those at 40 ◦ C.
Chea et al. (2018) SiO2, Al2O3 AOS (anionic) 25, 40, 60, 1.Generally, the presence of NPs does not significantly enhance the foamability in both
90 neutral and alkaline base fluids with increasing temperature.
2.Foam half-life degraded with increasing temperature, the presence of NPs did enhance
its half-life at that temperature.
3. Al2O3 possessed a better stabilizing effect than SiO2 at all temperatures regardless of
the system pH.
4.Adding NPs in foam does not change the bubble’s shape at any time and in any
condition.
5. The presence of NPs in the foam could retard its liquid drainage rate and thus slowing
down the bubble coalescence process that will eventually increase foam half-life and
stability at any temperature.
Fu et al. (2018) SiO2 25, 60 1.The residual oil recovery by CO2/SiO2 flooding decreased as the temperature increased.
Zhu et al. (2019) AlOOH CTAB (anionic) 30, 50, 70 1.Stability enhancement was observed by NPs addition under high-temperature
EAPB (amphoteric) conditions, though temperature would accelerate the liquid drainage and gas diffusion.
Bashir et al. (2019) Fumed SiO2, AOS (anionic) 25, 85 1. Using optimum concentrations of additives (e.g. NPs) can enhance the stability of the
Rice husk ash foam structure at high-temperature conditions.
(RHA)
Li et al. (2019) Hydrophilic SiO2 HHSB 70 1.The improved foam stability at high temperature was attributed to the enhanced
disjoining pressure with the addition of NPs, thereby preventing film thinning.

17
S.A. Ab Rasid et al. Journal of Petroleum Science and Engineering 208 (2022) 109475

Table 8
Table summary of the influence of flow velocity on foam stability.
Reference Nanoparticle Surfactant Shear Rate/Flow rate/ Main Findings
Measurement

Xiao et al. SiO2 ANS 5950− 17,850 1/s, 1.The non-Newtonian foam rheology is significantly affected by shear rate. The
(2016) (anionic) Flow Loop apparatus foams generated with NP, NP + NaCl, and NP + ANS systems exhibited shear-
thinning behavior at all qualities, with maximum foam viscosity at the lowest shear
rate.
2.At lower salinity, systems exhibited a transition from shear-thinning at the low
quality (50%) to shear-thickening behavior at the high quality (60, 70, 80%).
3.At higher salinity, systems exhibited a shear-thickening behavior at all qualities.
4.The critical shear rate for foam generation starts at 8925 1/s⫺1 (9 ml/min) in high
salinity environments.
Yekeen et al. Hydrophilic SiO2, SDS 0.1–0.5 ml/s, 1.Generally, the foam (with or without NPs) apparent viscosity decreased with
(2017b) Hydrophilic Al2O3 (anionic) Etched glass micromodel increasing flow rate. This can be attributed to the shear-thinning effect.
2.Shear thinning behavior is expected in this low-quality regime. Hence, the
bubbles rheology decreases with the increasing flow rate.
3.At the lowest flow rate, the apparent viscosity of the SDS-foam increases from
20.34 cp to 44.91 cp during the flow process of Al2O3-SDS foam and 56.73 cp during
the flow process of SiO2-SDS foam.
Rognmo et al. Silane modified colloidal AOS 120–240 ml/h, Co- 1. Surfactant-stabilized foam demonstrates a shear-thinning behavior, with a
(2017) SiO2 (anionic) injection in sandstone core significant reduction in apparent viscosity with an increase in superficial velocity,
whereas NP-stabilized foam exhibits no changes in apparent viscosity with changes
in superficial velocities.
Rahim Risal Bare SiO2, Surface SDS 0.5–2.0 cc/min, Capillary 1. The apparent viscosity is decreasing as the shear rate decreasing indicates a
et al. (2018) modified SiO2-A and (anionic) viscometer, pseudoplastic behavior of power law fluid.
SiO2–B Glassbead pack 2. At low shear rate, the apparent viscosity for surface-modified SiO2-stabilized
foams are higher than the conventional foam.

Table 9
Table summary of the influence of nanoparticles retention on foam stability.
Reference Nanoparticle Rock or mineral types Main Findings

Helene F. Fullerol, Single-wall carbon nanotubes Silicate beads 1. High mobility index calculated for Fullerol and SWNT agrees as their low value
Lecoanet et al. (SWNT), SiO2, Alumoxane, n-C60, Anatase, for attachment efficiency and rapid breakthrough.
(2004) Feroxane 2. A saturation or blocking of deposition sites observed in the porous medium
except for metal oxane. This suggests that mobility of these NPs may increase over
time as deposition sites become saturated over progressively larger distances within
the porous medium.
Yu et al. (2012a) SiO2 Sandstone #1 (33md), Static test:
Sandstone #2 1. Static adsorption test: limestone (5.501 mg/g) > sandstone (1.272 mg) >
(57.6md), dolomite (0 mg/g)
Limestone (132md), 2. Low adsorption of SiO2 in sandstone may be due to similar composition and
Dolomite (5.29md) probably due to the clay content
3. Higher adsorption in limestone may be due to high electrostatic attraction force
between NPs and limestone surface.
4. All static adsorption test reach equilibrium in less than 12 h.
Dynamic test:
1. The resulted recovered NPs is in alignment with static test.
2. Low perm (33mD) sandstone recovered 23% injected np while high perm
(57.6mD) recovered 86% NPs, with almost zero np during post flush at no change in
pressure & permeability.
3. NP breakthrough occurs after 2 PV np injection in limestone core, 32.6%
recovered NP and no change in pressure and perm.
4. NP breakthrough immediately in dolomite. When 2 PV particle dispersion were
injected, silica concentration in the effluent is reached to the injection
concentration. 95.8% np recovery after 1.5 PV brine. However, the pressure
continuously increases even after brine injection, indicating plugging and change in
perm.
Metin et al. (2012) Unmodified SiO2, Calcite mineral, 1. Insignificant adsorption of unmodified, sulfonate, and PEG-modified SiO2 NPs
Sulfonate-modified SiO2, Quartz sand observed on quartz and calcite surfaces.
Quat-modified SiO2, 2.Increase in particle size from 5 to 25 nm or the addition of NaCl less than CSC
PEG-modified SiO2 (critical salt concentration) does not promote the adsorption of NPs onto mineral
surface.
Rognmo et al. SiO2 Bentheimer sandstone 1. NP breakthrough occurred after + - 1 PV; less 1 PV indicating poor volumetric
(2017) sweep efficiency and more 1 PV indicating better volumetric sweep efficiency
2. Satisfaction of retention were observed between 2.6 and 3.8 PV.
3. Average retention for all cores (4) was 401 μg/g.
4. NP elusion was calculated to 79 μg/g. Hence 19.6% of retained NP were
mobilized during the post-flood.
5. Differential pressure during flooding remained constant; no plugging or increase
in viscosity. Hence the mechanical trapping was unlikely.
Sun et al. (2021) SiO2 Silica sand pack system 1. The NPs can flow through the sand pack easily because of the big difference
(231–4320 md) between particles size and pore throat size.
2. The permeabilities for all sand packs decrease after nanofluid injection; however,
the permeability loss rate is about 1.3% when the sand pack is 4320mD.
3. The relatively high permeability formation is suitable choice when injecting
nanofluid, which will cause less damage and recover more particles from the
effluent.
18
S.A. Ab Rasid et al. Journal of Petroleum Science and Engineering 208 (2022) 109475

7. The shear-thinning behavior of foam in the presence of NPs is not


always seen; in certain circumstances, shear-thickening phe­
nomenon has been reported. The behavior of the NP-surfactant
foam appears to be determined by a combination of factors
such as NP, surfactant, and salt concentrations, as well as shear
rate/flow velocity. However, there is still a lack of understanding
on the extend of shear-thickening behavior observed. Therefore,
in-depth investigation is recommended to improve the under­
standing on the NPs stabilized foam behavior.
8. NP loss in porous media is unavoidable. However, by selecting
the correct type and size of NPs, the retention can be minimized.
9. NP reduced surfactant adsorption on the rock surface because the
surfactant favored adsorption on the NP surface over adsorption
on the rock surface. Such a preference has resulted in the
continued presence of surfactant in the foam structure and
Fig. 17. Mechanisms of nanoparticles’ retention in porous media. consequently the longer stability of that structure.
10. Regardless of the type of NPs used, several studies show that the
have opposite charges, adsorption of the surfactant on the NP’s presence of NPs increased the stability of surfactant foam. How­
surface improves the surface properties of the NP from hydro­ ever, the degree of stability enhancement varied depending on
philic to partially hydrophobic and leads to higher foam stability. the properties of the NPs (i.e., density, size, shape, wettability,
However, an excessive increase in hydrophobicity may lead to and surface charge).
particle aggregation resulting in NP deposition and reduced foam
stability. Therefore, the concentrations of NPs and surfactants In terms of future studies, five points should be considered:
that have opposite charges must be optimally determined.
5. In most laboratory studies, increasing salinity decreased the sta­ i) Even if the rest of the key parameters such as NP wettability,
bility of NP-surfactant foam. However, in a few cases, depending surface coating etc. is kept constant, it seems that an optimal NP
on the NP and surfactant used, an increase in stability was size exists which is favorable for foam stability. Therefore, further
observed due to the shielding effect of the salt ions. studies are recommended to experimentally prove whether such
6. The presence of oil and the temperature rise lower the stability of an optimal particle size exists for a given NP surfactant foam
the foam even in the presence of the NP, however this reduction system.
in stability is much less than if the NP had not been added to the ii) NP loss in a porous medium is an economically damaging process
structure of the surfactant foam. that occurs through mechanisms including adsorption, mechan­
ical entrapment and log-jamming, and particles settling due to

Table 10
Summary of findings related to competitive adsorption of surfactant molecules in the presence of nanoparticles and rock surface.
Reference Surfactant Rock surface Main Findings

Ahmadi and Hydrophilic Saponin (zizyphus Shale sandstone 1. Mechanism of adsorption of saponin onto hydrophilic SiO2 is hydrogen bonding
Shadizadeh SiO2, spina-chriti) rock between hydroxyl groups of SiO2 and saponin.
(2013) Hydrophobic 2. Hydrophobic SiO2 is more effective than hydrophilic SiO2 to inhibit adsorption losses
SiO2 onto shale sandstone.
Yekeen et al. SiO2, Al2O3 SDS (anionic) Kaolinite . Surface tension of surfactant increases in the presence of NPs; attributed by the surfactant
(2017b) adsorption on the NPs’ surfaces which reduced the concentration of free surfactant in the
bulk solution.
2. Competitive adsorption of the surfactant molecules on the NP and kaolinite surface
exists; as more surfactant adsorbed on NP then less will be adsorbed on kaolinite.
3. NP and adsorbed surfactant molecule formed a negative charge cluster that resides in
bulk solution rather than being adsorbed onto kaolinite.
Saxena et al. (2019) SiO2, Al2O3 Soapnut fruit Sandstone 1. As the surfactant concentration increases, more amount of surfactant was adsorbed onto
(anionic) (negative charge), the sample surface
negative charge), 2. When NPs are added to the system, they also act as an adsorbent leading to competitive
Bentonite adsorption of surfactant molecules
(negative charge) 3. It was observed that the adsorption of surfactant molecules on the NP surface was higher
due to stronger hydrogen bonding. Thus, less amount surfactant molecule available to be
adsorbed onto the rock surface.
4. The bonded NP and surfactant molecule stays in the bulk solution reduces of surfactant
adsorption onto the rock surface
5. The adsorption reduction capability of SiO2 is greater than the reduction capacity of
Al2O3 can be explained by the almost rounded structure compared to Al2O3.
Yekeen et al. SiO2 CTAB (cationic), organic-rich shale, 1. A threshold surfactant concentration exists, below which there is a complete adsorption
(2019b) SDS (anionic), Quartz rich shale of surfactant monomers on the adsorbent surface (shale) - in the absence of NP.
Triton-X-100 (non- 2. The amount of surfactant adsorbed on the adsorbent (shale) surface decrease with
ionic) increasing temp - in the absence and presence of NP.
3. The surfactant adsorption on the shale increase as the salinity concentration increases.
4. The maximum adsorption capacity of surf onto KPK shale decrease in the presence of
SiO2 NP.
5. NP also serves as an adsorbent; consequently, there is a competition for the adsorption of
surf molecules between np and the shale surface.
6. In the presence of NP, the concentration of free surfactant molecules in the bulk solution
reduces due to the high adsorption on the surface of NPs.

19
S.A. Ab Rasid et al. Journal of Petroleum Science and Engineering 208 (2022) 109475

Table 11
Table summary of the influence of nanoparticles type on foam stability.
Authors Nanoparticle Surfactant Main Findings

Yekeen et al. (2019a) Al2O3, CuO, SDS (anionic), 1.Considerable reduction in n-decane-water interfacial tension and improvement in
TiO2, ZnO, SDBS (anionic), foam stability was observed at 80 ◦ C, due to electrostatic attraction between the charges
ZrO2, SiO2 CTAB (cationic) on NPs surfaces and surfactant head-groups.
Phong et al. (2020) Fly ash, MFOMAX (zwitterionic) 1.Surfactant with NPs is able to improve foam stability regardless of the type and
Fabricated fly ash concentration of NPs. The irreversible adsorption of the NPs at the gas-liquid interface
improves the foam stability by reducing the direct contact between the fluids, thus
reducing the effect of liquid drainage, bubble coalescence, and bubble coarsening.
2. As the concentration of NPs increases, the foam stability also increases due to the
increase of NPs in the gas-liquid interface, which helps form an aggregate particle layer
to reduce liquid drainage.
3. The larger NPs size, the higher the tendency to aggregate and causes liquid drainage.
The increase in gravitational force also causes the film thinning and bubble coalescence.
Rossen and Wang (1999), Fly ash AOS (anionic), 1. The adsorption of fly ash particles on the surface of the liquid film can improve the
Wang et al. (2020a) SDS (anionic), alkylphenol viscoelasticity of the bubbles, thus prevent the drainage of the liquid layer and gas
ethoxylates (nonionic) diffusion between bubbles effectively, delay the aggregation and rupture speed of foam
and improve the foam stability.
2. The surface tension decreases with the increase in fly ash content. The adsorption of
fly ash particles on the surface of the liquid film reduces surface tension, thus improve
the foam stability.
3. Fly ash particles can form a dense shell-like structure on the liquid film’s surface, thus
preventing the liquid drainage, the gas diffusion between bubbles, and delayed thinning
the foam film.
Raja Ramanathan and SiO2, multiwalled carbon AOS (anionic) 1. NP aggregation yielded poor foam stability at high concentrations.
Nasr-El-Din, 2020 tube (MWCNT) 2. The MWCNT and SiO2 NPs enhanced the foam half-life of the AOS by 4.6 and 2.9
times, respectively.
3. The SiO2 NPs and MWCNT started to aggregate upon the addition of salt due to
increased Van der Waals attraction forces. MWCNT was found to be more stable than the
SiO2 NPs at the same salinity.
4. Core flood experiments showed an increase in the pressure drop across the core during
the foam injection stage.
5. AOS/MWCNT foam performed better over AOS/SiO2 NPs was attributed to its rod-like
structure, size, zeta potential, and partial hydrophobicity.

gravity. Depending on the type of mechanism, the strategy for References


minimizing or removing the influence of these mechanisms dif­
fers. However, there is currently no way to determine and mea­ 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.
sure the contribution of each of these mechanisms to the total J. Petrol. Sci. Eng. 112, 239–247.
retention in dynamic tests, which could be an interesting topic for Ahmadi, M., Habibi, A., Pourafshry, P., Ayatollahi, S., 2011. Zeta Potential Investigation
future research. and Mathematical Modeling of Nanoparticles Deposited on the Rock Surface to
Reduce Fine Migration. In: SPE Middle East Oil and Gas Show and Conference, 25-28
iii) Further research may be conducted to understand the actual September 2011 2011 Manama. Bahrain,. SPE.
mechanism between NPs and salt ions in the presence of surfac­ Akbari, S., Mahmood, S.M., Ghaedi, H., Al-hajri, S.J.P., 2019a. A new empirical model
tant molecules and their effect on NP fluid mixture stability and for viscosity of sulfonated polyacrylamide polymers, 11, 1046.
Akbari, S., Mahmood, S.M., Nasr, N.H., Al-Hajri, S., Sabet, M., 2019b. A critical review of
the stabilized foam performance. concept and methods related to accessible pore volume during polymer-enhanced oil
iv) Field applications of NP-stabilized foam and challenges encoun­ recovery. J. Petrol. Sci. Eng. 182, 106263.
tered in field implementation are essential issues that must be Al-Hajri, S., Mahmood, S.M., Abdulelah, H., Akbari, S., 2018. An overview on polymer
retention in porous media. Energies 11, 2751.
addressed in future studies.
Alcorn, Z.P., Foyen, T., Gauteplass, J., Benali, B., Soyke, A., Ferno, M., 2020. Pore- and
v) Modifications of NP’s wettability for foam stabilization is one of core-scale insights of nanoparticle-stabilized foam for CO2-enhanced oil recovery.
the major controlling parameters of foam stabilization by NPs Nanomaterials (Basel) 10.
and should be discussed in future review papers. The authors may Almajid, M.M., Kovscek, A.R., 2020. Pore network investigation of trapped gas and foam
generation mechanisms. Transport Porous Media 131, 289–313.
explain the mechanisms of NPs modifications through chemi­ Alyousef, Z., Almobarky, M., Schechter, D., 2017. Enhancing the stability of foam by the
sorption and physisorption, comparing the applications, advan­ use of nanoparticles. Energy Fuel. 31, 10620–10627.
tages and disadvantages of surface wettability modification using Alyousef, Z.A., Almobarky, M.A., Schechter, D.S., 2018. The effect of nanoparticle
aggregation on surfactant foam stability. J. Colloid Interface Sci. 511, 365–373.
polymers and surfactants. Alzobaidi, S., Lotfollahi, M., Kim, I., Johnston, K.P., Dicarlo, D.A., 2017. Carbon dioxide-
in-brine foams at high temperatures and extreme salinities stabilized with silica
Declaration of competing interest nanoparticles. Energy Fuel. 31, 10680–10690.
Arriaga, L.R., Drenckhan, W., Salonen, A., Rodrigues, J.A., Íñiguez-Palomares, R., Rio, E.,
Langevn, D., 2012. On the long-term stability of foams stabilised by mixtures of
The authors declare that they have no known competing financial nano-particles and oppositely charged short chain surfactants. Soft Matter 8.
interests or personal relationships that could have appeared to influence Bashir, A., Sharifi Haddad, A., Rafati, R., 2019. Nanoparticle/polymer-enhanced alpha
olefin sulfonate solution for foam generation in the presence of oil phase at high
the work reported in this paper. temperature conditions. Colloid. Surface. Physicochem. Eng. Aspect. 582.
Binks, B.P., Lumsdon, S.O., 2000. Influence of particle wettability on the type and
Acknowledgement stability. Langmuir 16, 8622–8631.
Binks, B.P., Rodrigues, J.A., 2007. Synergistic interaction in emulsions stabilized by a
mixture of silica nanoparticles and cationic surfactant. Langmuir 23, 3626–3636.
The authors would like to thank the PETRONAS Research Sdn. Bhd., Binks, B.P., Kirkland, M., Rodrigues, J.A., 2008. Origin of stabilisation of aqueous foams
Malaysia; Universiti Teknologi PETRONAS, Malaysia for supporting this in nanoparticle–surfactant mixtures. Soft Matter 4.
research through research management grants. Boud, D.C., holbrook, O.C., 1958. Gas Drive Oil Recovery Process. Google Patents.
Chea, T.B., Akhir, N.A.M., Idris, A.K., 2018. Investigation on the effect of types of
nanoparticles and temperature on nanoparticles-foam stability. In: International

20
S.A. Ab Rasid et al. Journal of Petroleum Science and Engineering 208 (2022) 109475

Conference on Industrial Engineering and Operations Management. IEOM Society Massarweh, O., Abushaikha, A.S., 2021. A Review of Recent Developments in CO2
International, Bandung, Indonesia. Mobility Control in Enhanced Oil Recovery. Petroleum.
Choi, M., Choi, W.K., Jung, C.H., Kim, S.B., 2020. The surface modification and Maurya, N.K., Mandal, A., 2018. Investigation of synergistic effect of nanoparticle and
characterization of SiO2 nanoparticles for higher foam stability. Sci. Rep. 10, 19399. surfactant in macro emulsion based EOR application in oil reservoirs. Chem. Eng.
Dang, T.Q.C., Chen, Z., Nguyen, T.B.N., Bae, W., 2014. Investigation of isotherm polymer Res. Des. 132, 370–384.
adsorption in porous media. Petrol. Sci. Technol. 32, 1626–1640. Metin, C.O., Baran JR., J.R., Nguyen, Q.P., 2012. Adsorption of surface functionalized
Denkov, N.D., Ivanov, I.B., Kralchevsky, P.A., Wasan, D.T., 1992. A possible mechanism silica nanoparticles onto mineral surfaces and decane/water interface. J. Nano Res.
of stabilization of emulsions by solid particles. J. Colloid Interface Sci. 150, 589–593. 14, 1246.
Eftekhari, A.A., Krastev, R., Farajzadeh, R., 2015. Foam stabilized by fly ash Mo, D., Jia, B., Yu, J., Liu, N., Lee, R., 2014a. Study nanoparticle-stabilized CO2 foam for
nanoparticles for enhancing oil recovery. Ind. Eng. Chem. Res. 54, 12482–12491. oil recovery at different pressure, temperature, and rock samples. In: SPE Improved
Eide, Ø., Føyen, T., Skjelsvik, E., Rognmo, A., Fernø, M., 2018. Nanoparticle stabilized Oil Recovery Symposium. Society of Petroleum Engineers, Tulsa, Oklahoma, USA.
foam in harsh conditions for CO2 EOR. In: Abu Dhabi International Petroleum Mo, D., Jia, B., Yu, J., Liu, N., Lee, R.L., 2014b. Study nanoparticle-stabilized CO2 foam
Exhibition & Conference. Society of Petroleum Engineers, Abu Dhabi, UAE. for oil recovery at different pressure, temperature, and rock samples. In: SPE
Emrani, A.S., Nasr-El-Din, H.A., 2015. Stabilizing CO2-Foam Using Nanoparticles. SPE Improved Oil Recovery Symposium. SPE, Tulsa, Oklahoma, USA.
European Formation Damage Conference and Exhibition, Budapest, Hungary: SPE. Noor, M.Z.B.M., Teng, W.Y., Irawan, S., 2019. Stabilization of silicone dioxide
Emrani, A.S., Ibrahim, A.F., Nasr-El-Din, H.A., 2017. Evaluation of mobility control with nanoparticle foam in tertiary petroleum production. Indonesian J. Chem. 19.
nanoparticle-stabilized CO2 foam. In: SPE Latin America and Caribbean Petroleum Osei-Bonsu, K., Grassia, P., Shokri, N., 2017. Relationship between bulk foam stability,
Engineering Conference. surfactant formulation and oil displacement efficiency in porous media. Fuel 203,
Esmaeilzadeh, P., Hosseinpour, N., Bahramian, A., Fakhroueian, Z., Arya, S., 2014. Effect 403–410.
of ZrO2 nanoparticles on the interfacial behavior of surfactant solutions at air–water Panahpoori, D., Rezvani, H., Parsaei, R., Riazi, M., 2019. A pore-scale study on
and n-heptane–water interfaces. Fluid Phase Equil. 361, 289–295. improving CTAB foam stability in heavy crude oil− water system using TiO2
Fameau, A.-L., Salonen, A., 2014. Effect of particles and aggregated structures on the nanoparticles. J. Petrol. Sci. Eng. 183.
foam stability and aging. Compt. Rendus Phys. 15, 748–760. Phong, G.M., Pilus, R.M., Mustaffa, A., Thangavel, L., Mohamed, N.M., 2020.
Farhadi, H., Riahi, S., Ayatollahi, S., Ahmadi, H., 2016. Experimental study of Relationship between Fly Ash Nanoparticle-Stabilized-Foam and Oil Production in
nanoparticle-surfactant-stabilized CO2 foam: stability and mobility control. Chem. Core Displacement and Simulation Studies. Fuel, p. 266.
Eng. Res. Des. 111, 449–460. Prigiobbe, V., Worthen, A.J., Johnston, K.P., Huh, C., Bryant, S.L., 2015. Transport of
Fu, C., Liu, N., 2020. Study of the synergistic effect of the nanoparticle-surfactant- nanoparticle-stabilized CO $$_2$$ 2 -foam in porous media. Transport Porous Media
polymer system on CO2 foam apparent viscosity and stability at high pressure and 111, 265–285.
temperature. Energy Fuel. 34, 13707–13716. Qian, C., Telmadarreie, A., Dong, M., Bryant, S.L., 2019. Synergistic Effect between
Fu, C., Yu, J., Liu, N., 2018. Nanoparticle-stabilized CO2 foam for waterflooded residual Surfactant and Nanoparticles on the Stability of Foam in EOR Processes. SPE Western
oil recovery. Fuel 234, 809–813. Regional Meeting, San Jose, California, USA: SPE.
Gbadamosi, A.O., Junin, R., Manan, M.A., Agi, A., Yusuff, A.S., 2019. An overview of Rafati, R., Haddad, A.S., Hamidi, H., 2016. Experimental study on stability and
chemical enhanced oil recovery: recent advances and prospects. Int. Nano Lett. 9, rheological properties of aqueous foam in the presence of reservoir natural solid
171–202. particles. Colloid. Surface. Physicochem. Eng. Aspect. 509, 19–31.
Gugl, R., 2020. Modelling and Simulation of Foam-Assisted Water-Alternating-Gas Rahim Risal, A., Manan, M.A., Yekeen, N., Mohamed Samin, A., Azli, N.B., 2018.
Injection in Naturally Fractured Carbonate Reservoirs. University of Leoben. Rheological properties of surface-modified nanoparticles-stabilized CO2 foam.
Guo, F., Aryana, S., 2016. An experimental investigation of nanoparticle-stabilized CO 2 J. Dispersion Sci. Technol. 39, 1767–1779.
foam used in enhanced oil recovery. Fuel 186, 430–442. Raja Ramanathan, O.A., Nasr-El-Din, Hisham A., 2020. A New Effective Multiwalled
Hamza, M.F., Sinnathambi, C.M., Aljunid Merican, Z.M., Soleimani, H., Karl D, S., 2017. Carbon Nanotube-Foam System for Mobility Control. Abu Dhabi International
An overview of the present stability and performance of EOR-foam. Sains Malays. 46, Petroleum Exhibition & Conference, Abu Dhabi, UAE.
1641–1650. Razali, N., Chai, I.C.H., Zainal, S., A Manap, A.A., 2018. Nano-sized particles as foam
Horozov, T., 2008. Foams and foam films stabilised by solid particles. Curr. Opin. Colloid stabiliser designed for application at thigh temperature and light crude oil condition
Interface Sci. 13, 134–140. for EOR - a fluid-fluid case study. Offshore Technology Conference Asia. Kuala Lumpur.
Hu, N., Li, Y., Wu, Z., Lu, K., Huang, D., Liu, W., 2018. Foams stabilization by silica Risal, A.R., Manan, M.A., Yekeen, N., Azli, N.B., Samin, A.M., Tan, X.K., 2018.
nanoparticle with cationic and anionic surfactants in column flotation: effects of Experimental investigation of enhancement of carbon dioxide foam stability, pore
particle size. J. Taiwan Inst. Chem. Eng. 88, 62–69. plugging, and oil recovery in the presence of silica nanoparticles. Petrol. Sci. 16,
Ibrahim, A.F., Nasr-El-Din, H.A., 2019. CO 2 Foam for Enhanced Oil Recovery 344–356.
Applications. Foams-Emerging Technologies. IntechOpen. Rognmo, A.U., Horjen, H., Fernø, M.A., 2017. Nanotechnology for improved CO 2
Jafari Daghlian Sofla, S., James, L.A., Zhang, Y., 2018. Insight into the stability of utilization in CCS: laboratory study of CO 2 -foam flow and silica nanoparticle
hydrophilic silica nanoparticles in seawater for Enhanced oil recovery implications. retention in porous media. Int. J. Greenh. Gas Control 64, 113–118.
Fuel 216, 559–571. Rognmo, A.U., Heldal, S., Fernø, M.A., 2018. Silica nanoparticles to stabilize CO2-foam
Kalam, S., Kamal, M.S., Patil, S., Hussain, S.J.P., 2019. Role of counterions and nature of for improved CO2 utilization: enhanced CO2 storage and oil recovery from mature
spacer on foaming properties of novel polyoxyethylene cationic gemini surfactants, oil reservoirs. Fuel 216, 621–626.
7, 502. Rossen, W.R., 2017. Foams in Enhanced Oil Recovery. Foams. Routledge.
Kaptay, G., 2006. On the equation of the maximum capillary pressure induced by solid Rossen, W.R., Wang, M.W., 1999. Modeling foams for acid diversion. SPE J. 4 (2), 9.
particles to stabilize emulsions and foams and on the emulsion stability diagrams. Sagbana, P.I., Abushaikha, A.S., 2021. A comprehensive review of the chemical-based
Colloid. Surface. Physicochem. Eng. Aspect. 282–283, 387–401. conformance control methods in oil reservoirs. J. Petrol. Explor. Prod. Technol. 11,
Karakashev, S.I., Ozdemir, O., Hampton, M.A., Nguyen, A.V., 2011. Formation and 2233–2257.
stability of foams stabilized by fine particles with similar size, contact angle and Satter, A., Iqbal, G.M., 2015. Reservoir Engineering: the Fundamentals, Simulation, and
different shapes. Colloid. Surface. Physicochem. Eng. Aspect. 382, 132–138. Management of Conventional and Unconventional Recoveries. Gulf Professional
Kim, I., Worthen, A.J., Johnston, K.P., Dicarlo, D.A., Huh, C., 2016. Size-dependent Publishing.
properties of silica nanoparticles for Pickering stabilization of emulsions and foams. Saxena, N., Kumar, A., Mandal, A., 2019. Adsorption analysis of natural anionic
J. Nanoparticle Res. 18. surfactant for enhanced oil recovery: the role of mineralogy, salinity, alkalinity and
Kruglyakov, P.M., Elaneva, S.I., Vilkova, N.G., 2011. About mechanism of foam nanoparticles. J. Petrol. Sci. Eng. 173, 1264–1283.
stabilization by solid particles. Adv. Colloid Interface Sci. 165, 108–116. Schramm, L.L., Wassmuth, F., 1994. Foams: basic principles. In: SCHRAMM, L.L. (Ed.),
Kumar, S., Mandal, A., 2017. Investigation on stabilization of CO 2 foam by ionic and Foams: Fundamentals and Applications in the Petroleum Industry. United States of
nonionic surfactants in presence of different additives for application in enhanced oil America.
recovery. Appl. Surf. Sci. 420, 9–20. Singh, R., Mohanty, K.K., 2015. Synergy between nanoparticles and surfactants in
Lake, L.W., 1989. Enhanced Oil Recovery. Prentice Hall. stabilizing foams for oil recovery. Energy Fuel. 29, 467–479.
Lecoanet, Helene F., Bottero, Jean-Yves, Wiesner, M.R., 2004. Laboratory assessment of Singh, R., Mohanty, K.K., 2017a. Nanoparticle-stabilized foams for high-temperature,
the mobility of nanomaterials in porous media. Environ. Sci. Technol. 38, high-salinity oil reservoirs. In: SPE Annual Technical Conference and Exhibition (San
5164–5169. Antonio, Texas, USA).
Lee, J.K., Ko, J., Kim, Y.S., 2017. Rheology of fly ash mixed tailings slurries and Singh, R., Mohanty, K.K., 2017b. SPE-187165-MS nanoparticle-stabilized foams for high-
applicability of prediction models. Minerals 7. temperature, high-salinity oil Reservoirs.pdf>. In: SPE Annual Technical Conference
Li, W., Wei, F., Xiong, C., Ouyang, J., Shao, L., Dai, M., Liu, P., Du, D., 2019. A novel and Exhibition. Society of Petroleum Engineers, San Antonio, Texas, USA.
supercritical CO2 foam system stabilized with a mixture of zwitterionic surfactant Singh, R., Mohanty, K.K., 2018. Study of nanoparticle-stabilized foams in harsh reservoir
and silica nanoparticles for enhanced oil recovery. Front Chem. 7, 718. conditions. Transport Porous Media 131, 135–155.
Llamas, S., Ponce Torres, A., Liggieri, L., Santini, E., Ravera, F., 2019. Surface properties Sun, Q., Tan, L., Wang, G. J. I. J. O. M. P. B., 2008. Liquid foam drainage: an overview,
of binary TiO2 - SiO2 nanoparticle dispersions relevant for foams stabilization. 22, 2333–2354.
Colloid. Surface. Physicochem. Eng. Aspect. 575, 299–309. Sun, Q., Li, Z., Wang, J., Li, S., Li, B., Jiang, L., Wang, H., Lü, Q., Zhang, C., Liu, W., 2015.
Majed Almubarak, Z.A., Almajid, Muhammad, Almubarak, Tariq, Ng, Jun Hong, 2020. Aqueous foam stabilized by partially hydrophobic nanoparticles in the presence of
Enhancing Foam Stability through a Combination of Surfactant and Nanoparticles. surfactant. Colloid. Surface. Physicochem. Eng. Aspect. 471, 54–64.
Abu Dhabi International Petroleum Exhibition & Conference. Abu Dhabi, UAE). Sun, S., Wang, Y., Yuan, C., Wang, H., Wang, W., Luo, J., Li, C., Hu, S., 2020. Tunable
Manichand, R.N.N., Seright, R.S.S., 2014. Field vs. Laboratory polymer-retention values stability of oil-containing foam systems with different concentrations of SDS and
for a polymer flood in the tambaredjo field. SPE Reservoir Eval. Eng. 17, 314–325. hydrophobic silica nanoparticles. J. Ind. Eng. Chem. 82, 333–340.

21
S.A. Ab Rasid et al. Journal of Petroleum Science and Engineering 208 (2022) 109475

Sun, Q., Liu, W., Li, S., Zhang, N., Li, Z., 2021. Interfacial rheology of foam stabilized by surfactant-foam stability in presence of silicon dioxide and aluminum oxide
nanoparticles and their retention in porous media. Energy Fuel. 35, 6541–6552. nanoparticles. J. Petrol. Sci. Eng. 159, 115–134.
Vatanparast, H., Shahabi, F., Bahramian, A., Javadi, A., Miller, R., 2018. The role of Yekeen, N., Manan, M.A., Idris, A.K., Padmanabhan, E., Junin, R., Samin, A.M.,
electrostatic repulsion on increasing surface activity of anionic surfactants in the Gbadamosi, A.O., Oguamah, I., 2018a. A comprehensive review of experimental
presence of hydrophilic silica nanoparticles. Sci. Rep. 8, 7251. studies of nanoparticles-stabilized foam for enhanced oil recovery. J. Petrol. Sci. Eng.
Veyskarami, M., Ghazanfari, M.H., 2018. Synergistic effect of like and opposite charged 164, 43–74.
nanoparticle and surfactant on foam stability and mobility in the absence and Yekeen, N., Manan, M.A., Idris, A.K., Samin, A.M., Risal, A.R., 2018b. Mechanistic study
presence of hydrocarbon: a comparative study. J. Petrol. Sci. Eng. 166, 433–444. of nanoparticles–surfactant foam flow in etched glass micro-models. J. Dispersion
Wang, K., Wang, G., Lu, C., Pei, C., Wang, Y., 2017. Preparation and investigation of Sci. Technol. 39, 623–633.
foaming amphiphilic fluorinated nanoparticles for enhanced oil recovery. Materials Yekeen, N., Padmanabhan, E., Idris, A.K., 2019a. Synergistic effects of nanoparticles and
(Basel) 10. surfactants on n-decane-water interfacial tension and bulk foam stability at high
Wang, T., Fan, H., Yang, W., Meng, Z., 2020a. Stabilization mechanism of fly ash three- temperature. J. Petrol. Sci. Eng. 179, 814–830.
phase foam and its sealing capacity on fractured reservoirs. Fuel 264. Yekeen, N., Padmanabhan, E., Idris, A.K., Ibad, S.M., 2019b. Surfactant adsorption
Wang, Y., Wang, J., Fan, H., Du, F., Zhou, W., Yang, J., 2020b. Effect of inorganic salt on behaviors onto shale from Malaysian formations: influence of silicon dioxide
foam properties of nanoparticle and surfactant systems %. J Tenside Surfactants nanoparticles, surfactant type, temperature, salinity and shale lithology. J. Petrol.
Detergents 57, 382–388. Sci. Eng. 179, 841–854.
Xiao, C., Balasubramanian, S.N., Clapp, L.W., 2016. Rheology of supercritical CO2 foam Yu, J., An, C., Mo, D., Liu, N., Lee, R., 2012a. Study of adsorption and transportation
stabilized by nanoparticles. In: SPE Improved Oil Recovery Conference (Tulsa, behavior of nanoparticles in three different porous media. In: Eighteenth SPE
Oklahoma, USA). Improved Oil Recovery Symposium (Tulsa, Oklahoma, USA).
Xu, L., Rad, M.D., Telmadarreie, A., Qian, C., Liu, C., Bryant, S.L., Dong, M., 2018. Yu, J., Liu, N., Li, L., Lee, R.L., 2012b. Generation of Nanoparticle-Stabilized
Synergy of surface-treated nanoparticle and anionic-nonionic surfactant on Supercritical CO2 Foams. Carbon Management Technology Conference, Orlando,
stabilization of natural gas foams. Colloid. Surface. Physicochem. Eng. Aspect. 550, Florida, USA: SPE.
176–185. Zargartalebi, M., Kharrat, R., Barati, N., 2015. Enhancement of surfactant flooding
Xu, Z., Li, B., Zhao, H., He, L., Liu, Z., Chen, D., Yang, H., Li, Z., 2020. Investigation of the performance by the use of silica nanoparticles. Fuel 143, 21–27.
effect of nanoparticle-stabilized foam on EOR: nitrogen foam and methane foam. Zeng, Y., Muthuswamy, A., Ma, K., Wang, L., Farajzadeh, R., Puerto, M., Vincent-
ACS Omega 5, 19092–19103. Bonnieu, S., Eftekhari, A.A., Wang, Y., Da, C., Joyce, J.C., Biswal, S.L., Hirasaki, G.J.,
Xuezhen, W., Mohanty, K., 2020. Nanoparticle-Stabilized Strong Foam for EOR in 2016. Insights on foam transport from a texture-implicit local-equilibrium model
Fractured Reservoirs. SPE Annual Technical Conference & Exhibition, Denver, with an improved parameter estimation algorithm. Ind. Eng. Chem. Res. 55,
Colorado, USA. 7819–7829.
Yang, W., Wang, T., Fan, Z., 2017a. Highly stable foam stabilized by alumina Zhang, Y., Liu, Q., Ye, H., Yang, L., Luo, D., Peng, B., 2021. Nanoparticles as foam
nanoparticles for EOR: effects of sodium cumenesulfonate and electrolyte stabilizer: mechanism, control parameters and application in foam flooding for
concentrations. Energy Fuel. 31, 9016–9025. enhanced oil recovery. J. Petrol. Sci. Eng. 202.
Yang, W., Wang, T., Fan, Z., Miao, Q., Deng, Z., Zhu, Y., 2017b. Foams stabilized by in Zhao, J., Torabi, F., Yang, J., 2021. The Synergistic Role of Silica Nanoparticle and
situ-modified nanoparticles and anionic surfactants for enhanced oil recovery. Anionic Surfactant on the Static and Dynamic CO2 Foam Stability for Enhanced
Energy Fuel. 31, 4721–4730. Heavy Oil Recovery: an Experimental Study. Fuel, p. 287.
Yekeen, N., Idris, A.K., Manan, M.A., Samin, A.M., 2016. Experimental study of the Zhu, T., Ogbe, D., Khataniar, S.J.I., Research, E.C., 2004. Improving the foam
influence of silica nanoparticles on the bulk stability of SDS-foam in the presence of performance for mobility control and improved sweep efficiency in gas flooding, 43,
oil. J. Dispersion Sci. Technol. 38, 416–424. 4413–4421.
Yekeen, N., Idris, A.K., Manan, M.A., Samin, A.M., Risal, A.R., Kun, T.X., 2017a. Bulk and Zhu, J., Yang, Z., Li, X., Hou, L., Xie, S., 2019. Experimental study on the microscopic
bubble-scale experimental studies of influence of nanoparticles on foam stability. characteristics of foams stabilized by viscoelastic surfactant and nanoparticles.
Chin. J. Chem. Eng. 25, 347–357. Colloid. Surface. Physicochem. Eng. Aspect. 572, 88–96.
Yekeen, N., Manan, M.A., Idris, A.K., Samin, A.M., Risal, A.R., 2017b. Experimental
investigation of minimization in surfactant adsorption and improvement in

22

You might also like