You are on page 1of 16

Journal of Molecular Liquids 275 (2019) 265–280

Contents lists available at ScienceDirect

Journal of Molecular Liquids

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

Wettability alteration in carbonate and sandstone rocks due to low


salinity surfactant flooding
Ali Aminian, Bahman ZareNezhad ⁎
Faculty of Chemical, Petroleum and Gas Engineering, Semnan University, PO Box 35195-363, Semnan, Iran

a r t i c l e i n f o a b s t r a c t

Article history: Commonly, desorption of crude oil from rock surfaces accomplished by the exposure to the decreasing salt con-
Received 23 June 2018 centrations of NaCl, Na2SO4, MgCl2 and CaCl2 solutions. However, adsorption and desorption of crude oil are the
Received in revised form 27 August 2018 functions of several attributing factors at undoubtedly low salinity brine solutions. For silica surfaces, both rock
Accepted 16 November 2018
and oil possess negative charges during low salinity injection that leads to the electrical double layer expansion
Available online 19 November 2018
and electrostatic repulsion between them. For carbonate rocks, rock dissolution and negatively charged surface at
Keywords:
higher pH could be regarded as the driving forces for oil desorption. Thus, adsorption/desorption studies of dif-
Enhanced oil recovery ferent oils over silica and carbonate rocks while considering salinity, ion composition and aging of the rocks
Low salinity surfactant would be sufficient to alter the wettability of the rocks. In the present paper, the effect of saline water containing
pH surfactant has been discussed on the performance of enhanced oil recovery (EOR). Wettability alteration trig-
Smart water gered by the change in interfacial tension (IFT) between crude oil and low salinity surfactant (LSS) is known to
Carbonate be the main factor regarding EOR processes. However, wettability depends on several factors, namely, multiple
Sandstone ion exchange (MIE), interfacial elasticity modulus, cation bridging, acid/base numbers of oil, rock composition
and structure, ratio of divalent/monovalent (M2+/M+) cations in the formation brine, aging time, the status of
zeta potential, surfactant structure and concentration, pH of the medium, T & P of reservoir, permeability and
pressure drop change. Therefore, we thoroughly examine the microscopic mechanisms of LSS flooding via study-
ing individual aforementioned factors that influence the contact angle change in sandstone and carbonate rocks.
© 2018 Elsevier B.V. All rights reserved.

1. Introduction [2]. On the other hand, the conditions under which oil recovery can be
attained by LSS flooding have been under question. Identifying the rela-
Waterflooding process cannot provide complete oil recovery from tionships between wettability and IFT is crucial to quantify the best pro-
porous media in reservoirs; therefore, alternative processes such as sur- cedure for oil recovery. This quantification is complicated due to the
factant or polymer flooding, known as tertiary recoveries, are of consid- strong interactions within crude oil/brine/rock systems. Interfacial me-
erable importance [1]. niscus development can also influence the relative permeability of
Re-injection of produced water from wells is useful for secondary in- water in a porous rock that originates from the wettability alteration
jection that does not produce formation damage similar to what we are of a rock toward water-wetness. Even though, initial wettability and
being observed in nanoparticles or surfactant flooding; however, the saturation history of a rock can also affect the wettability [2]. Therefore,
amount of produced water is limited and as a result, modified formation researchers are trying to monitor the wettability of a rock via contact
brines are coming to play role. angle measurements and use IFT to determine the interactions among
It is assumed that wettability alteration due to interfacial tension fluid phases (brine/surfactant/oil). Another approach is to calculate
change is responsible for the enhanced oil recovery, while there are sev- total disjoining pressure for a crude oil/brine/rock system that can tell
eral interaction forces behind it. Moreover, the high cost associated with us the extent of wettability alteration upon injection of low salinity
the surfactant flooding and possibility of damages to the well due to sur- brine solutions.
factant adsorption enforced researchers to put some criterions for the Oil recovery depends on the several factors, hence, brine composi-
concentration of surfactants and theirs structure in a brine solution tion & concentration and initial wetting condition of reservoir rock are
among the most important ones [3]. Frequently, reservoir rocks are ini-
⁎ Corresponding author.
tially oil-wetted, while attempts directed toward turning the oil-wet
E-mail addresses: ali_aminian@ymail.com (A. Aminian), bzarenezhad@semnan.ac.ir rock into a water-wet rock. Wetting alteration by destabilizing the oil
(B. ZareNezhad). layers adhered to reservoir rock through LSS flooding is a promising

https://doi.org/10.1016/j.molliq.2018.11.080
0167-7322/© 2018 Elsevier B.V. All rights reserved.
266 A. Aminian, B. ZareNezhad / Journal of Molecular Liquids 275 (2019) 265–280

EOR technology due to enhanced capillary number. Once capillary num- Another aspect of LSS flooding is the reduction in the differential pres-
ber increased, the higher permeability and lower pressure drop at the sure just at the start of the injection, which is especially important for low
pore entrance will result [4,5]. permeability reservoirs. The experiments showed that LSW could in-
Usually, LSS provides driving factors for: (a) the interfacial tension crease the differential pressure at the beginning of injection due to the oc-
reduction, (b) the change in three-phase contact angle among solid currence of emulsification, which is much lower in the presence of
rock-oil-water, and (c) instability in interface at the oil-water leading surfactants. Moreover, a decrease in the IFT to ultra-low values (i.e.
to oil-drop detachment due to the action of buoyancy force [6]. The suc- ~10−2 to 10−4 mN/m) greatly increased the oil displacement efficiency
cess of a LSS flooding heavily depends on the ionic strength and ion [22]. They claimed that emulsification could deteriorate displacement ef-
compositions. On the other hand, aging and dissolution of a carbonate ficiency, because IFT reduction accomplished by varying the salinity at
rock exposed to low salinity brines may contribute to the process of constant concentration of the ethoxylated fatty alcohol carboxylate
change in wettability and oil detachment [7,8]. (AEC) surfactant. They gradually increased NaCl and CaCl2 concentration
Performance of a successful LSS is strongly depends on the preflush at constant surfactant concentration of 0.2 wt% to reach an IFT about
of the reservoir with injecting the low or high salinity water (LSW or 1.44 × 10−4 mN/m and then flew the solutions over the initially water-
HSW). On the other hand, a LSS process possesses lower capillary num- wet quartz particles. It might be linked to the accumulation of salts at
ber because of its higher IFT than that of optimal salinity surfactant in- the interface between solid and brine that prevents the hydrogen bonding
jection primarily because of the stronger interaction exist between between nonionic components of oil and the solid surface; therefore, ions
rock and fluid in the case of a LSS. However, surfactant retention in electrostatically screen off from the interface and oil displacement in-
LSS process is lower than that of optimal salinity surfactant case, al- creases [22,23]. In both Refs. [22,23], oil components were nonionic that
though its IFT is higher than the IFT of the optimal salinity surfactant could not electrostatically adsorb onto the surface, and as a result high
[9–11]. The main advantage of LSS is the prominent wettability alter- concentration of salts deterred them from attaching onto the rock (Fig. 1).
ation of rocks in the presence of oil at which wettability increases to- Therefore, the following guidelines are presented when the LSS
ward water phase as the rock surface exposed to LSS. Also, there are a flooding is considered for oil displacement applications:
number of cases which show an increase in the ratio of divalent to
monovalent cations could greatly augment the reduction in the IFT of – The loss of surfactant due to adsorption on the solid rock surface in-
LSS without increasing the concentration of surfactant [12–14]. creases once its concentration increased in the solution.
For example, the vast amount of recent experiments show that low – The surfactant adsorption can be lowered by preflushing rock with a
concentration of monovalent sodium cation can reduce the contact LSW solution.
angle of brine droplet on the muscovite surface in n-decane + stearic – Further mobilization of oil after LSS flooding can be obtained by
acid solution, while decreasing the sodium concentration as calcium injecting HSW solutions other than surfactant.
concentration rises yield higher contact angle [15,16]. The acidic com- – IFT of a surfactant strongly depends on the pH. IFT between surfac-
ponents of oil can heavily influence both pH-state of the medium and tant solutions and oil droplets can be further lowered by increasing
the hydration state of ions that leads to a repulsive force. Specifically, pH. However, since IFT change due to change over the entire pH is
deporotonated acid components at higher pH bridged to the divalent very small it is worthwhile to focus on the LSS and its salinity to ad-
cations on the muscovite surface that makes the surface turned into just the IFT rather than adjusting PH.
more oil-wet. Since the hydration energy of sodium ion is much lower – Low salinity effect may be worked for brine compositions of up to
than that of Ca2+, the monovalent cation Na+ at lower pH will provide 5000 ppm.
conditions for the water bridging that leads to the oil release from the – Surfactant concentration should be kept as low as possible in order
surface as the bond between Ca2+ and oil breaks [17]. to increase the phase transition toward intermediate oil-water
In order to accurately examine the role of LSS in oil displacement in solubilization.
reservoir conditions, it is necessary to separate different forces among
different phases. One of the most important interactions is fluid-fluid in-
teractions that can be quantitatively verified by IFT measurements. This Referring to the fourth item of the above guideline, the presence of
kind of IFT refers to the interfacial tension between oil drop and sur- divalent Ca2+ cation in the surfactant solution can greatly adjust the
rounding fluid comprising surfactant dissolved into a brine solution amount of surfactant adsorbed on the solid rock surface. The phenome-
[18]. Surfactant concentration is often chosen to be low in the range of non called as cation bridging with the aid of divalent cations dictate the
0.1 to 0.5 wt% of brine solution. The other crucial interaction is between rate and the quantity of surfactant adsorption with no contributions
rock and fluid that literally refers to surfactant adsorption at the solid in- from Na+ monovalent cation [24]. Thus, decreasing the concentration
terface. The surfactant concentration at the effluent stream versus time of Ca2+ in the brine can readily reduce adsorption of surfactant onto
can be a representative of the amount adsorbed on the solid rock sur- the rock and it could trigger fines migration with oils attached to
face. The rock-fluid interactions is also characterized through wettabil- them. At LSS solutions with low ratio of Ca2+/Na+, pH could be raised
ity alteration measurement, which can be represented by contact via H+/Na+ exchange and as a result repulsion contribution to the
angle [19,20]. DLVO formulation exceeded the attraction one; leading to the expan-
The experiments are showing that after LSS flooding of a sion of electrical double layer and fines migration [25].
preflushed LSW rock, water relative permeability is improved com- Compared to LSS, tertiary flooding with HSW cannot overcome the
pared to that of LSW alone did [21]. This is because of the reduction energy barrier of the overlapped hydration shells of polar/acidic compo-
in IFT, which in turn increases the mobilization of oil, water satura- nents of oil and the cations adsorbed on the mineral surface of reser-
tion and finally increases the water relative permeability. In order voirs. However, in the low salinity flooding divalents are exchanged
to examine the role of salts in the wettability alteration of a rock sur- with the monovalents due to high exchange rate capacity of the me-
face, researchers carried out experiments on the state of being dium that makes the surface more water-wet as the electrical double
water-wet upon injection of brines with different ratios of divalent layer thickness increases [26]. Due to the complex nature of crude oil-
to monovalent ions. The results showed that LSS dissolved only in so- LSS-rock system, there is no clear comprehensive challenging view
dium chloride has better performance under initially water-wet con- over the influence of above-mentioned factors on the mechanisms of
ditions, while the opposite is true with calcium under initially less wettability alteration using LSS flooding. Thus, we summarize the indi-
water-wet conditions. A salinity shift after LSS to HSW can reduce vidual deciding factors that determine the wettability alteration of car-
IFT roughly an order of magnitude and this will intensify both in- bonate and sandstone rocks in the presence of surfactant in low salinity
situ oil and surfactant solution. brine solutions.
A. Aminian, B. ZareNezhad / Journal of Molecular Liquids 275 (2019) 265–280 267

Fig. 1. MD simulations for nonionic α-D-glucopyranose monomers between calcite slabs terminated with (1014) plane with different concentration of brines. The atom colors correspond
to O: red, C: cyan, H: white, Na: blue, Cl: yellow, Ca: green and Mg: pink. The water molecules not showed [23].

2. Factors affecting contact angle the absence of divalent ions [28]. For the LSS flooding systems, three
main mechanisms have been suggested, in which water molecules pen-
Fig. 2 depicts a schematic of the contact angle of an oil droplet over a etrate to attach to the solid surface by destroying the organized layer of
rock surface in aqueous brine solution. As shown, increasing the contact oil molecules in initially oil-wet rocks:
angle leads to more oil-wet conditions.
A contact angle decrease about a few degrees can also trigger a con- 1- Hydrophobic interactions of the chains of surfactant with the oil de-
siderable increase in oil recovery with low salinity surfactant flooding in stroy the upper layer molecules and creates a way for the water to
penetrate,
2- DLVO and non-DLVO forces between water and solid surface will
pave the way for disturbing the remaining layers of oil, and
3- Once water reached on the top of the solid, hydrogen bonding breaks
the contact lines of oil-solid surface and creates water channel to
augment the detachment process of oil from solid surface known
as water bridging. Upon arrival on the silica solid surface, water mol-
ecules breakdown the bond between Si atoms and oxygen atoms to
establish hydroxylated silica surface under the force of hydration.

Churaev et al. [29] have shown that for a secure hydrophilic quartz
surface, it is required to use high concentration of CTAB solutions
(N10−3 M), which in turn led to creation of hydrophilic silica and oil sur-
faces. Thus, wetting of silica surface and preventing the oil droplets from
attaching to quartz need strong electrostatic and hydrophilic repulsion
that may occur upon high CTAB solutions. However, at low CTAB con-
centrations (b10−5 M), both surfaces became hydrophobic and strong
Fig. 2. Contact angle of an oil droplet over solid substrate in brine. The final contact angle is hydrophobic attraction persisted the attachment of oil droplets to silica
an average between the right and left contact angles [27]. surface due to hydrophobic attraction.
268 A. Aminian, B. ZareNezhad / Journal of Molecular Liquids 275 (2019) 265–280

Khanamiri et al. [30] revealed that tertiary low salinity surfactant so- OH groups at alumina silica surfaces compared to the silica allowed
lution with only sodium cations could act better than LSS with sodium more polar interactions between oil components and alumina silica
and calcium cations. The surfactant was a blend of alkyl benzene sulfo- surface.
nates with CMC of 0.059 wt%. As shown in Fig. 3, contact angle of oil/LSS/ Moreover, contact angle at three-phase contact line is depend on the
Berea sandstone increased in the presence of LSS aqueous solution con- temperature, composition of the brine (pH) and oil, and aging of the
taining only sodium chloride that lead to a wettability shift from core sample flooded by the low salinity brine solutions. An oil with
strongly water-wet to intermediate water-wet. However, for LSS- large value of acid number will raise contact angles as more oil compo-
NaCa, the presence of calcium cations enforce the surfactant molecules nents adsorbed on the mineral surfaces. Increasing the temperature and
to align more compactly to each other, and as a result interfacial elastic- aging time would increase the contact angle at three-phase contact line
ity increases and contact angle has smaller value. It is theoretically [37].
proved that monovalent cations like Na+ and K+ contribute positively
to the DLVO formulation through MIE and DLE mechanisms, which 2.1. Multiple ion exchange (MIE)
could increase the overall value of the disjoining pressure [31]. Also, it
is worth mentioning that different surfactant should be tested against Clays are layers of several tetrahedral and octahedral sheets that the
salinity to obtain a stable LSS solution. number of tetrahedral and octahedral sheets comprising a layer deter-
Nourani et al. [32] showed that upon exposing silica and alumina sil- mines the type of a clay, namely, 1:1, 2:1, and 2:1 with a hydroxide in-
ica surfaces to HSW, LSW and LSS aqueous solutions, contact angle al- terlayer clay minerals [38]. Depending on the structure of multi-layered
tered from partially water-wet to completely water-wet conditions as clay minerals, compensating cations can be exchanged with other cat-
the injected brine diluted in divalent concentrations and as surfactant ions present in the brine solutions providing cation-specific sites on
added into the medium [33–35]. LSW and LSS solutions added to the the mineral surfaces. Theses compensating cations that regulate the
core samples of silica and alumina silica surfaces as secondary and ter- charge equality of tetrahedral and octahedral sheets can be exchanged
tiary flooding modes, respectively, after the core initially flooded with with cations presented in the brine solution, which control oil desorp-
HSW solution. For both silica and alumina silica surfaces, LSS flooding tion or adsorption. For example, an initial structure of muscovite clay
over silica resulted in much more lower contact angles compared to mineral with different cations is shown in Fig. 4. If Na-muscovite ex-
those of alumina silica surfaces. The differences between contact angles changed with Ca2+ in the brine, then cation bridging can adsorb
between two mineral surfaces originated from the crystal structures, in deprotonated oil components onto the clay mineral and makes the sur-
which silanol (Si-OH) groups of silica surfaces showed higher hydrophi- face more oil-wet.
licity with respect to the less hydrophilic natures of the Brønsted acidic Furthermore, potential of mean force (PMF) and adsorption Gibbs
silanol and Lewis acidic aluminol (Al-OH) groups of alumina silica sur- energy can be used to investigate the occurrence of cation exchange be-
faces. The latter caused extra interactions with Lewis basic functional tween muscovite surface and porotonated and deporotonated decanoic
groups in the asphaltenes of the oil [36]. In addition, higher density of acid molecules [16]. The results showed that at very low pH (pH b 2),
uncharged acidic oil molecules have not any interaction with the Mus-
covite surface regardless of cation exchange condition of the surface.
The repulsion between the Muscovite surface and oil molecules ex-
plained through the magnitude of adsorption Gibbs energy and the sta-
tus of overlapped layers of hydration between –COOH group of oil
molecules and cations adsorbed on the surface of Muscovite. In contrast,
at higher pH where oil molecules are deporotonated, the PMF and ad-
sorption Gibbs energy curves revealed Muscovite surface strongly inter-
act with deporotonated oil molecules through cation exchange in the
order of Ca2+ N K+ N Mg2+ N Na+ (Fig. 5). In conclusion, the effect of cat-
ion type on oil-wetness of muscovite surface is larger at high pH. On the
other hand, low salinity surfactant could destabilize the cation bridging
and alters the wettability of rock surface to partially water-wet. Ca2+
and K+ interacted with oil molecules via cations bridging while Mg2+
and Na+ on the Muscovite did not directly interacted with oil molecules
as the higher hydration energy of later cations proved it. The study also
demonstrated that between Ca2+ and Mg2+ divalent cations only Ca2+
can alter the wettability of the Muscovite surface toward more oil-wet.
This behavior toward more oil-wet via cation bridging in the order of
Ca2+ N K+ N Mg2+ N Na+ were observed for illite, smectite and Berea
sandstone. For K+ and Ca2+, cation bridging occurs directly between
cation and deporotonated oil as this is the case in the bulk aqueous so-
lution, while for Mg2+ and Na+ water film bridges cations and
deporotonated oil molecules, which differs from that exist in the bulk
aqueous solution [16].
For carbonate rocks, MIE occurs if SO42− ions presented in the low sa-
linity surfactant solution; so, low salinity surfactant containing SO42− or
diluted seawater can be useful to change the wettability of carbonate
reservoirs toward more water-wet [39,40]. Ca2+cations on the calcite
surface can be exchanged with SO42− ions in the diluted surfactant solu-
tion and as result, the hydrophobic interactions between polar compo-
nents of oil and negatively charged calcite surface will repel each
Fig. 3. Contact angle (θ) of three-phase contact line of oil/LSS interface with rock surface
other [41]. This may also happen when Mg2+ cations substitute Ca2+
for both LSS-Na and LSS-NaCa. Surfactant: blend of alkyl benzene sulfonates, rock: Berea on the calcite surface and make the breakdown of bond between Ca2+
sandstone, cations: sodium and calcium [30]. and the polar components of oil. As the concentration of SO42− on the
A. Aminian, B. ZareNezhad / Journal of Molecular Liquids 275 (2019) 265–280 269

Fig. 4. Muscovite structure with different compensating cations. a) Na+, b) K+, c) Mg2+ and d) Ca2+ cations [16].

calcite surface increased, it released complexes of Ca2+-polar compo- could not improve oil recovery (injection after 6 PV, Fig. 6). Thus, low sa-
nents of oil and hindered adsorption of further oil components due to linity water should contains some SO42− ions when rock is depleted
the creation of electrostatic repulsion between them [42]. In addition from anhydrites [46]. A diluted seawater with or without surfactant is
to SO42−, BO32−and PO43−are also useful to improve oil recovery from car- recommended in this case.
bonate rocks [43]. Even though, cores containing anhydrite produce On the other hand, in sandstone rocks, dissolution of silicate min-
SO42− ions upon rock dissolution via LSS injection. Thus, LSS can be erals is too slow that can be neglected. However, as salinity of formation
used instead of diluted seawater in such cases. brine decreases the accumulation of negative charges increased on the
Haagh et al. [44] proposed that for the sandstone rock, the two pos- silica surface and upon electrostatic repulsion between negatively
sible mechanisms for wettability alteration are: double layer expansion charged fines and silica surface, they migrate from the surface and oils
and multicomponent ion exchange (MIE). It was declared that DLE has attached to them will be released. In some situations, throat plugging
minor effect on the wettability alteration, while MIE has a pronounced of the core may occur because of fines migration leading to the increase
effect on water-wetness of sandstone surface by cation bridging and in pressure drop. Also, production of oil-free kaolinite fines in the water
complexion mechanism. Therefore, they proposed that overlap of the or fines adhered to the oil as a Pickering emulsion observed [47,48].
hydration layers of ions/surfaces and MIE play much more important
roles than double layer expansion. Since DLVO formulation fails in thin 2.3. Cation bridging
film, short-range forces like overlapped layers of hydration (non-
DLVO forces) is the key for the success of LSS flooding. By referring to Fig. 7, only divalent cations can adhere to the polar
The main advantage of LSS flooding with decreasing concentration components of oil when they exchanged with the compensating cations
of divalents is the occurrence of MIE in which divalents cannot bridge on the clay minerals. As a result, they bridged between clay surfaces and
surfactant to the rock surface. Another reason pertains to the polar components of oil and surface becomes oil-wet. For carbonate res-
screened-off ions from the rock surface that help the compact arrange- ervoirs, specifically at lower pH and higher acid numbers of oil, the min-
ment of surfactants at the oil/brine interface, in turn, disturbance in oil eral surface becomes abundant in positive charges and therefore Ca2+
layers because of lower IFT resulted. cation can readily adsorb negatively charged components of oil.
In a study, Allen et al. [49] investigated monovalent cesium cation
2.2. Rock dissolution bridging between mica and AOT anionic surfactant. It was under-
stood that anionic surfactants can adsorb to mica through Ca2+cat-
PHREEQC software is being used in order to determine the potential tion bridging, but this was not the case for monovalent ions Na+ .
effect of mineral dissolution upon low salinity water/surfactant injec- Therefore, Na+ of AOT surfactant cannot bridge mica to the surfac-
tion. The modeling results showed that Ca dissolution for the calcite tant. The studies show that metal cations have the affinity toward
rock can be ignored in the presence of low salinity water, hence, core mica in the order K+, Rb+ , Cs+ N Na + N Li+ [7,8]. On the other
plugs containing anhydrites (CaSO4) could trigger SO42− ions into the hand, for smectites/montmorillonite clay surfaces the binding
aqueous solution. Since Ca2+ dissolution in formation brine solution is strength is K+ N Na+ N Ca2+ N Cs+ N Ba2+. Lower hydration energy
limited, therefore, mineral dissolution has not anything to do with low of Cs+ monovalent ion than that of Na+ makes the cation bridging
salinity water/surfactant injection [36,45]. between negative AOT and the negative mica surface (Fig. 7).
In another calculation, formation water and 100-times diluted for- Generally, at low PH, cation bridging effect is ignored because of the
mation water injected onto the limestone core containing anhydrites overall repulsive force that originates from the overlapping of the hy-
in secondary and tertiary modes, respectively [46]. The simulations re- dration layers of mineral surface and acidic groups of oil. Whereas at
vealed that in the case of core plugs with anhydrites oil recovery in- high pH, ionized state of acidic groups tend to adhere to the counter
creased for formation brine and injection of diluted formation brine ions on the mineral surface. The adsorption Gibbs energy in this case
(after 6 PV). However, in another simulation where anhydrites removed is negative showing the tendency of oil molecules to adsorb on the min-
(simulation B), no oil recovery observed proving that even diluted brine eral surface.
270 A. Aminian, B. ZareNezhad / Journal of Molecular Liquids 275 (2019) 265–280

Fig. 6. Simulations for oil recovery in cores with anhydrite dissolution (simulation A) and
without anhydrite dissolution (simulation B). FW: formation water, rock: limestone with
or without anhydrite, CATIONS: NaCl solution with equal amounts of Ca2+ and Mg2+ [ 46].

at the interface. Upon adsorption of second layer, zeta potential starting


to decrease since columbic repulsion overcomes the hydrophobic
attraction force. Furthermore, in the absence of surfactant, divalent cat-
ions are more likely to decrease the value of zeta potential as more pos-
itive charges accumulate at the interface. When surfactant and Ca2+
cations are both present, zeta potential increases suggesting that Ca2+
is responsible for the adsorption of surfactant via cation bridging,
while monovalent cations have a little contribution to the bridging ef-
fect. Increasing the concentration of Ca2+ beyond the formation of sur-
factant monolayer would not have any significant effect on the previous
recorded zeta potential and the formation of second layer is less likely to
occur. Generally, Ca2+ facilitates the adsorption of surfactant to the
rock; however, low salinity injection and layering phenomena along
with ions confinement inhibit the surfactant adsorption that enhances
the repulsive hydration forces between rock/brine surfaces.
Divalent cations stabilize solid by lowering the zeta potential, and as
a result the repulsive force lowered and higher amounts of charge
added to the surface. Therefore, after secondary flooding with high sa-
linity water and through multiple ion exchange, zeta potential de-
creases resulting in a new equilibrium that established between oil-

Fig. 5. Cation bridging for Ca2+ and K+ and water bridging for Mg2+ and Na+ on the
muscovite surface [16].

Gerold and Henry [50] investigated surfactant adsorption on a solid


rock surface in the presence of metal cations through zeta potential
measurements. For a pair of surfactant-rock system with the same
charge, adsorption being started as the hydrophobic interaction over-
comes the columbic repulsion. Once an adsorbed monolayer formed, Fig. 7. Cation bridging between AOT anionic surfactant and negatively charged mica
zeta potential increases because of the accumulation of negative charges surface. Cation: Cs+ [49].
A. Aminian, B. ZareNezhad / Journal of Molecular Liquids 275 (2019) 265–280 271

brine and brine-rock interfaces. Depending on the initial wetness of the bridging effect between Ca2+ and carboxylate group of DHNA− on kao-
rock and upon low salinity surfactant injection the following events linite and montmorillonite surfaces (Fig. 9).
may occur: the strongest cation exchanges with the weakest cation, However, hydrophobic pyrophyllite surface had the strongest ad-
acid/base components of the oil bond with the solid surface or break sorption of oil molecules than neutral kaolinite and hydrophilic
down may happen, surfactant molecules bridge with new cations, inter- montmorillonite surfaces, primarily because of the hydrophobic in-
facial oil-brine decreases, interfacial elastic modulus changes and over- teractions (in the absence of hydrogen bonding or cations) that did
all result would be the alteration in wettability of the solid surface. exist between oil molecules and basal hydrophobic surface [52].
Underwood et al. [51] proposed a twofold description for the low- Neutral oil molecules and protonated DHNA through hydrophobic
salinity flooding over Montmorillonite clay surfaces. Under initially interactions strongly adsorbed onto the basal pyrophyllite surface, and
oil-wet clay surface, cation bridging is present between oil-wet clay therefore higher ionic strength resulted in lesser adsorption of DHNA
and organic compounds, while for the initially water-wet conditions on the pyrophyllite. This is because of the electrostatic repulsion be-
cation exchange occurs through replacement of divalents by monova- tween adsorbed counter ions and hydrophobic DHNA. On the other
lents. The first reason for this phenomenon is described by the level of hand, deprotonated DHNA− adsorption on the hydrophilic surfaces of
PH and the extent of protonation/charge of organic oil molecules. The kaolinite and montmorillonite was not significant, because hydrophobic
level of CO2 concentration dissolved in the formation water, generation DHNA− had not noticeable affinity toward hydrophilic surface [52].
of hydroxyl ions from a clay‑calcium sites, the existence of multicompo-
nent ion exchange mechanisms through which oil bond to the clay sur- 2.4. pH
face and acid number of crude oil would affect the overall pH level of the
fluid adjacent to the rock surface. The second reason is the ratio of diva- At moderate pH, the cation bridging between cations of brine solu-
lent to the monovalent cations present in the low salinity injection. The tion and acid components of oil make the oil attachment to the solid
lower ratios enhance the thickness of electrical double layer in which surface because ionized acid groups attract to the counter ions, and con-
replacement of divalents by monovalents occurs. Also, lower hydration sequently surface charge at the interfaces decrease. In order to maintain
of Na+ along with higher bonding strength compared to those of Ca2+ the water-wetness of a surface in this condition, it is preferred to use LSS
and Mg2+ divalents boost water-wetness of the clay surfaces in which containing monovalent cations because these ions have lower affinity to
cation bridging between divalents and oil molecules breaks and the sur- interact with deporotonated acid components in the oil and have lower
face wettability alters from oil-wet to water-wet. Fig. 8 shows the state hydration state with respect to divalent ones either.
of cation bridging between clay and organic compounds and cation Sandstone and carbonate minerals dissolution may increase the
bridging among identical molecules via MD simulations [51]. medium's pH. This effect can be viewed from ion exchange between
Binding energies was used in order to predict the interactions be- Na+ and H+, as follows:
tween oil organic species and clay minerals [52]. For a crude oil mixture
comprising hexane, cyclohexane, toluene and decahydro-2-naphthoic NNa þ Hþ ⟷NH þ Naþ
acid (DHNA), the DFT calculation results indicated that clay minerals
had higher affinity toward deprotonated anionic DHNA− resin and Ca2 where N denotes adsorbed ion. The uptake of H+ in this ion exchange
+
presented in the brine solution. Also, binding energies showed that reaction will raise pH. Specifically for sandstone rocks at low level of
adsorption of the oil organic molecules on the mineral surface enhanced Ca2+in low salinity aqueous surfactant solution, the above ion exchange
with Ca2+ and Na+, while the binding energy for the later was smaller reaction increases pH and as pH soars the magnitude of total disjoining
than the Ca2+ cation. For instance, toluene binding energy onto the ka- pressure will augment and the height of the barrier energy occurs at
olinite clay surface enhanced 363 kcal/mol with the aid of Ca2+. lower distances between rock/brine interfaces that leads to higher elec-
Negatively-charged montmorillonite with exchangeable Ca2+cations trostatic repulsion.
slightly intensified adsorption of oil organic molecules compared to The point at which silica surfaces are isoelectric (neutral charge) is
the uncharged kaolinite surface. This can be explained via cation pH ~ 2. Thus, at pH N 2, silica surfaces demonstrate negative charge

Fig. 8. a) Cation bridging between like-charged clay and oil organic molecules by a cation. b) Cation bridging among identical molecules by cations [51].
272 A. Aminian, B. ZareNezhad / Journal of Molecular Liquids 275 (2019) 265–280

Fig. 9. Optimized structures of cation bridging effect through DFT calculations, (a) Ca2+-DHNA− on kaolinite, and (b) Ca2+-DHNA− on montmorillonite [52].

and the charge density increases with pH. Therefore, by using LSS de- charges on the calcite surface, because concentration of SO42− ions de-
pleted in Ca2+, and as result of Na+/H+ ion exchange pH will increase creases as seawater diluted largely.
that makes negatively-charged silica surfaces repel negatively-charged As shown in Fig. 11, for calcite surface, increasing pH above 8 will re-
polar components of oil. The experiments showed that by using a low sult in the decrease of zeta potential as more CO32−, HCO3−, and OH−
salinity solution free from divalent cations, it is possible to have very dissolution occurs. However, for instance at pH 8, lowering the salinity
low contact angle, while a small increase in contact angle increases ad- of the solution makes Ca2+ and Mg2+ cations screened off from the sur-
sorption of acid components of oil onto the silica surfaces. face, which led to the expansion of double layer and the accumulation of
The negative adsorption Gibbs energy reveals the affinity of oil to- negative charges on the surface. Hence, at ultra-low salinity, polar com-
ward silica surfaces at higher pH. Specifically, divalent cations such as ponents of oil enter into the aqueous phase from the interface and
Ca2+ makes the surface more oil-wet at high pH, whereas at very low higher IFT at oil-brine and lower negative charges at rock-oil interfaces
pH (pH b 2), a low salinity aqueous solution will result a silica surface happen. At conditions where oil acid number is high, the positively
with no affinity toward oil regardless of its nature of cation exchange. charged of calcite surface can strongly adsorb negatively charged acid
(See Fig. 10.) components, while adsorption decreases as the acid number lowered.
For calcite surfaces, the surface charge is positive due to the high By switching to low salinity, zeta potential of calcite surface decreases
concentration of Ca2+ in the crystal structure. Also, at high pH (i.e. pH and deprotonated acid components repelled from the calcite surface
N 8), the increase in concentration of CO32−, HCO3− and OH− is in line due to the electrostatic repulsion between calcite and acid components
with pH that leads to a less positive surface. Also, dissolution of carbon- [54].
ate rocks increases the pH and shifting to low salinity intensifies this Since pH range prevailed in most seawaters is between 5.7 and 6.77,
trend toward higher pH in such a way that more and more divalent- the effect of pH increase on IFT is strongest at low pH values (pH b 5),
oil complexes released from the calcite surface upon replacement by and therefore increasing pH due to low salinity slightly increases the
SO42− ions. Therefore, switching to low salinity seawater/surfactant- IFT values at the oil/brine interface created by the surfactant [55]. So,
containing SO42− can screen off counter ions and the electrical double it is worthwhile to adjust the salinity of the solution to affect the IFT
layer expansion will increase the height of energy barrier against oppo- rather than focusing on the pH, because ions decrease the electrostatic
sitely charged oil components. However, very diluted seawater (i.e., 50- repulsion among surfactants and arrange them more compactly at the
times diluted seawater) losses its efficiency for increasing the negative interface.

2.5. Zeta potential

Through zeta potential measurements it is possible to determine the


magnitude of surface potential at rock-brine and brine-oil interfaces,
which denotes the amount of charge accumulated at the interfaces.
For instance, when brine is deionized water, zeta potential at calcite-
deionized water interface is always positive as a sign of the presence
of positive charges on the calcite surface over wide ranges of pH
(Fig. 12). Increasing pH results in decrease in zeta potential for calcite
minerals [54].
In Fig. 12, the dissolution of ions of the calcite lattice in deionized
water makes an increase in pH and more negative charges accumulate
at the surface. The equilibrium reactions during dissolution of calcite
particles are as follow:

CaCO3 þ Hþ ↔Ca2þ þ HCO3− @pHb3:5 ð1Þ

Fig. 10. Adsorption Gibbs energy for the adsorption of protonated C9H19COOH and CaCO3 þ 2Hþ ↔Ca2þ þ H2 CO3 ð2Þ
deprotonated C9H19COO− oil molecule. The inset figure represents contact angle versus
pH of the medium, while increasing the concentration of electrolytes will result in
higher contact angles [16,53].
CaCO3 þ H2 CO3 ↔Ca2þ þ 2HCO3− @3:5bpHb7 ð3Þ
A. Aminian, B. ZareNezhad / Journal of Molecular Liquids 275 (2019) 265–280 273

Fig. 11. Switching from high salinity to low salinity effects on the averaged zeta potential of calcite rock/brine at pH 8. FW: formation water, SW: seawater, 10d: 10 times diluted [54].

CaCO3 þ H2 O↔Ca2þ þ HCO3− þ OH− @pH N7 ð4Þ When surfactant adsorbed onto the rock surface, indeed, it can mod-
ify the rate of changing in the zeta potential at both the oil/brine and the
rock/brine interfaces. For instance, cationic surfactant adsorption at the
If zeta potential of brine/rock and oil/brine has the same polarity, rock surface increases the zeta-potential toward more positive values
then they repel each other, and as a result the rock surface would be and an ionic surfactant has the opposite impact. While the concentra-
more water-wet. Therefore, contact angle decreases as sum of brine/ tion of ions is high, the accumulation of these ions near the solid wall
rock and oil/brine zeta potentials increases: and competition for the adsorption decreased the adsorption of surfac-
As shown in Fig. 13, interface at oil/brine is usually charged posi- tants and therefore, the rate of increasing in zeta potential slowed [59].
tively at low pH, however, the quantity of positive charges decrease as For the silica surface, the presence of Ca2+ is essential to bridge between
pH increased and at sufficiently high pH, the interface becomes nega- surfactant and the rock surface, while Na+ has not anything to do in this
tive. Thus, as the pH increases, the same polarity between brine/rock regard.
and oil/brine interfaces makes the rock surface more water-wet.
In sandstone rocks this situation is much more better since surface 2.6. Disjoining pressure: DLVO and non-DLVO forces
zeta potential is always negative at low and high (i.e., 54,680 mg/L)
ionic strengths. Brines with high ionic strength, such as seawater, Total disjoining pressure can be summed up over three main terms
strongly shrink the extent of electrical double layer as the concentration including, long-range electrostatic repulsion, short-range Van der
of divalent cations increases, and consequently zeta potential moves to- Waals attraction and structural forces [60–62]. Van der Waals forces
ward positive charge. contribute negatively to the disjoining pressure because they represent
As can be seen from Fig. 14, for both quartz and calcite minerals, high the attraction between two surfaces, while electrostatic forces
salinity brines will result in near-zero and even positive zeta potentials, strengthen total disjoining pressure as Debye length increases with
while divalent cations speed up the rate of increasing zeta potential be- the expansion of electrical double layer due to low saline water [63].
cause the charge density at the interface much more influences in this Structural forces are hydrophobic and solvation forces, which cannot
case. Also, temperature has not any significant thing to do with the be explained through Van der Waals or electrostatic forces [64,65]. Hy-
zeta potential. A similar trend to what we saw in Fig. 14 can be ex- drophobic surfaces are generating hydrophobic attraction forces and
plained for the oil/brine interfaces in such a way that increasing the sa- adsorption of a surfactant can change the hydrophobicity/hydrophilicity
linity of the brine solution raises the accumulation of cations at the of a surface, hence, it is possible to calculate the magnitude of
interface, and therefore the interface becomes less negative and even
positive at higher concentrations. The large negative value of zeta po-
tential at low salinity brine is due to the adsorption of hydroxyl ions
and oil's asphaltene and resins, however, at high salinity brine, adsorp-
tion of Na+ and Mg2+ reduce the large values of negative charge at the
interface (Fig. 15) [58].

Fig. 13. Effect of the sum of brine/rock and oil/brine zeta potentials on the contact angle
over carbonate rocks. Red-line from [36], green-line from [56], blue, black, yellow, and
Fig. 12. pH-dependent zeta potential at calcite-deionized water interface [54]. orange-lines from [57].
274 A. Aminian, B. ZareNezhad / Journal of Molecular Liquids 275 (2019) 265–280

Fig. 14. Experimental values of zeta potential at rock/brine interfaces at different temperatures and salinity [58].

hydrophobic force of a surface, while surfactant changes the wettability ψ0 illustrates the potential of the isolated surfaces; h is the film thick-
of the surface: [66] ness and κ is: [69]

Πhydrophobic ¼ −C1 e−h=λ1 ð5Þ 1=2


κ−1 ¼ εε0 kT=2ρ∞ e2 ð9Þ
 
3 2
Πvdw ¼ ð−Hð15:96h=λ þ 2ÞÞ= 12πh ð1 þ 5:32h=λÞ
where ρ∞ is the charge density in the bulk, e stands for the electron
3
 −H=12πh ðfor thin filmsÞ ð6Þ charge 1.602 × 10−19C, ε0 symbolizes the vacuum permittivity
8.854 × 10−12 C2 /J·m, ε is the dielectric constant of the solution,
where H is Hamaker constant usually taken as 1 × 10−20 J and h denotes and T is 298 K [70,71].
film tickness [67]. For calcite surface with its (1014) cleavage plane exposed to low sa-
   linity brine solution comprising Ca2+, Na+ and Cl−, the atomic force mi-
2 croscopy revealed that a tightly-arranged layer of water molecules
Πelectst ¼ nkT 2ψr1 ψr2 coshðκhÞ þ ψr1 2 þ ψr2 2 = sinh ðκhÞ ð7Þ
formed adjacent to the calcite surface in which hydrated ions should
consume energy in order to overcome the energy barrier to leave the
where n is the number of ions (cations or anions) per unit volume, k
hydration shell and close themselves to the calcite surface [72]. The
represents Boltzmann constant, 1.381 × 10−23 J/K, ψr1 and ψr2 are the
MD simulations showed that Na+ is the closest one on the top of the
potentials at oil/water and water/solid interfaces, respectively: [68]
water film; Cl− and Ca2+ are arranged next to the sodium ions far
from the water film. Na+ cations have 6 water molecules in their hydra-
ψh ¼ ψ0  e−κh ð8Þ tion shell, while this number is 8 for Ca2+; therefore Ca2+ cations should

Fig. 15. Experimental values of zeta potential at oil/brine interfaces at different temperatures and salinity [58].
A. Aminian, B. ZareNezhad / Journal of Molecular Liquids 275 (2019) 265–280 275

Fig. 18. Prediction of total disjoining pressure in sodium sulfate aqueous solutions of
10 mM over mica surface. Blue squares represent experimental data of CFM
Fig. 16. Density profiles (mol/L) of MD simulations for different ions (and water measurements, red circle are predictions with considering hydration forces and green
molecules) perpendicular to the calcite surface exposed to low salinitybrine. The density triangles denote calculations in the absence of hydration forces [70].
of the water atoms at the bottom. Red: oxygen, blue: hydrogen [72].

an oil-wet rock. Also, brine solution of NaCl showed a little amount of


pay larger energy to leave its hydration shell and consequently it stays adhesion, however, water film did not interrupted significantly
farther from calcite surface and sodium cations (Fig. 16) [73]. (Fig. 17) [77].
This hydration repulsion force exerted between two oppositely Therefore, in the absence of the calculation of hydration forces the
charged interfaces at short ranges can be calculated through a two or net total disjoining pressure may not accurately predict the nature of in-
more exponentially-decayed terms as follow: [70,74,75] teraction among solid/brine/oil interfaces. For instance, total disjoining
pressure for sodium sulfate brine aqueous solutions without considerig
Πhydration ¼ C2 e−h=λ2 þ C3 e−h=λ3 ð10Þ the hydration forces of ions largely deviates from the real experimental
data.
where C2, λ2, C3 and λ3 are the coefficients and decay lengths of short- As we can see in Fig. 18, total disjoining pressures are all positive
range and long-range forces, respectively. through the entire water film demonstrating the stability of water film
Or, over the mica surface that enhances the water-wetness of the surface.
When surfactants are present in the aqueous solution, hydration
Πhydration−extended ¼ C4 e−ðh−b4Þ2=ðλ4Þ2 þ C5 e−ðh−b5Þ2=ðλ5Þ2 þ … ð11Þ forces between monovalents and the charged rock surface on one
hand, and hydrated head group interaction with the rock on the other
where C4, λ4, C5 and λ5 are the coefficients and decay lengths of short- hand add the total value of disjoining pressure. Monovalents are more
range and long-range forces, respectively; b4 and b5 are constants; and prone to leave their hydration shell and close themselves to the rock
h corresponds to the film thickness [70,76]. surface, while hydrated cations do stay away hydrated head group
Through a chemical force microscopy (CFM) with a functional tip from attaching to the rock surface. Thus, repulsive hydration force
near mica surface that resembles solid/oil interactions, different ions must be took into account because of the ions confinement between
injected in order to quantify the magnitude of interfacial forces contrib- rock and surfactant that produces layering phenomena controlling the
uting to the total disjoining pressure. The CFM results demonstrated interaction between hydrated surfaces in the aqueous solution.
that in the brine solution containing Na2SO4 the least adhesion force be-
tween water film and functional tip of CFM apparatus obtained, which is 2.7. Interfacial elasticity moduli
a sign of a stable water film. Whereas, brine containing CaCl2 revealed
the maximum adhesion due to the interruption of water layers and The two predominant mechanisms in water injection system are
break down of H-bonds between film and mica surface, which produced well known to be sweep-like and snap-off oil displacements. After

Fig. 17. Functional tip interaction with mica surfaces in the presence of a) sodium chloride, b) calcium chloride, and c) sodium sulfate brine aqueous solutions (dark blue balls denoting
water molecules) [70].
276 A. Aminian, B. ZareNezhad / Journal of Molecular Liquids 275 (2019) 265–280

Fig. 19. Oil phase snap-off through pore throat in low and high salinity brines [80].
Fig. 21. Elastic modulus (G′) and viscous modulus (G″) vs NaCl concentration at the oil-
brine interface without and with 100 ppm of DEM surfactant [81].
LSW injection, snap-off of oil decreases and amount of entrapped oil re-
covery improves. However, in the case of HSW, structural forces as a result, oil's polar components accumulate at the interface [81].
weakens against increasing rate of capillary forces, hence, addition of However, it was mentioned that surfactant molecules were predomi-
surfactants could greatly reduce interfacial surface tension and increase nant at intermediate concentration of brines. At high salinity brine solu-
interfacial viscoelasticity (Fig. 19) [78,79]. tions, pressure fluctuations due to snap-off of the oil phase observed
An increase in interfacial elastic and viscous moduli elucidated as the after water breakthrough. On the other hand, adding nonionic surfac-
salt concentration decreases and with further decrease in salt concen- tants to slightly higher salinity brine could develop interface viscoelas-
tration, there was a decline. It was also shown that adsorption of surfac- ticity, while pressure fluctuations and pressure drop were much lower
tant at the oil-brine interface strongly affected its viscoelasticity [81]. than those of high salinity water and, furthermore, addition of nonionic
Surfactants may prevent the interaction between the oil's polar compo- surfactant canceled the effect of salt concentration by providing a stable
nents and the ions in the aqueous phase or even could avoid oil/brine interface that is not affected too much with the salt concentra-
asphaltenes from forming aggregations as depicted in Fig. 20. tion (Fig. 21) [81].
Experimental measurements of oil recovery in microfluidic device Khanamiri et al. [30] pointed out that interfacial rheology is the main
showed that low salinity surfactant solution was superior to the slightly cause for the wettability alteration in LSS oil recovery processes. The
high salinity water solution, in which a maximum of 15% increase in oil higher ratio of Ca2+/Na+ is responsible for the lower performance of
recovery obtained compared to the high salinity water [81]. The in- oil recovery in the presence of a surfactant. Generally, Ca2+ changes
crease in oil recovery suggested to be due to the intensification of elas- the interfacial shear viscosity at the three-phase contact line. Ca2+ cat-
ticity of the brine-oil interface. They reported an increase in interfacial ion tightens surfactant assembly at low salinity solution/oil interface,
elastic and viscous moduli at lower to intermediate concentration of which is produced a gel-like film that does not possess the power to
NaCl and MgCl2 brine solutions, because diffuse layer is broader and, sweep the oil effectively. However, in the absence of Ca2+cation in LSS

Fig. 20. Effect of surfactant at the brine-oil interface [81].


A. Aminian, B. ZareNezhad / Journal of Molecular Liquids 275 (2019) 265–280 277

solution and the initially water wet rock surfaces, a large amount of fine
oil clusters with lower interfacial dilational elastic modulus compared
to LSS with Ca2+/Na+ pushed the oil away from the rock instead of wet-
ting the rock or adhering to it. Fig. 22 compares the formation size of the
oil clusters in low salinity surfactant with and without Ca [30].
On the other hand, the initially oil-wet solid surfaces are more prone
to produce higher amount of oil with LSS-Ca2+/Na+. In addition, this is
also true for secondary low salinity water flooding, in which LSW con-
taining Ca2+/Na+ is more efficient than LSW without Ca2+. The forma-
tion of gel-like layer in LSW-Ca2+/Na+ adjacent to the rock surface
pushed the oil away from the rock instead of just wetting the rock sur-
face [30,82,83].

2.8. Surfactant structure

To prevent the formation of undesirable phases or instability of sur-


factant at reservoir conditions, surfactant-brine system usually supple-
mented with alkali and co-solvents. For instance, alkyl benzene
sulfonate anionic surfactants are often combined with alkyl ethoxy car-
boxylate in order to mitigate the precipitation of the surfactant at high
salinity environments by forming hydrogen bonding with water leading
to the change in the polarity of the surfactant [84]. Also, studies show
that alkyl chain of surfactants have played critical role in the reduction
of interfacial tension of oil-water, which defined as the extent of misci-
bility of the hydrophobic tail of the surfactant in the oil phase. For exam-
ple, aromatic substitution (with shorter alkyl chain) to the head of
surfactant led to the lower CMC, decrease in surface excess concentra-
tion and decline in the interfacial tension compared to that of mono-
alkyl benzene sulfonate at the same surfactant concentration (See
Fig. 23) [85]. Regularly, the head group of an ionic surfactant solvates
into water, while its alkyl chain attracts to the oil. Additional aromatic
substitution caused the higher tendency toward water than oil, and
therefore water tries to maximize the solvation of head group ring. In
the presence of cations, the electrical double layer creates as the cations
attracted toward the ionized water molecules and the length of the EDL
could reach the interface causing more water attracted in the three-
point contact line. Also, the alkyl chain of the para anionic surfactants
completely solvate into the oil phase, while the alkyl chain of the meta
Fig. 22. The largest residual oil clusters in a) LSS-NaCa and b) LSS-Na. The images are in the anionic surfactants partially solvates into the oil phase, therefore meta
same scale [30]. substitution of the primary alkyl chains increase the efficacy of the sur-
factants [85]. Once the surfactant forms a continuous layer at the inter-
face of oil-water, it can entirely separate the contact line between them

Fig. 23. Shifting the primary alkyl chain of the surfactant from the para to the meta position and the aromatic substitution on the head of the surfactant [85].
278 A. Aminian, B. ZareNezhad / Journal of Molecular Liquids 275 (2019) 265–280

and by reducing the interfacial tension it could eventually lead to oil de- can tolerate the hardness of seawater (presence of divalent cations)
tachment and formation of oil-water emulsions. by binding Ca2+ to their aggregates that originates from their lower
Surfactant's formulation should withstand high temperature and CMC and special aggregation morphologies (two anionic head groups)
pressure prevailed in the reservoirs in order to be soluble in brines compared to those of conventional anionic surfactants [93].
and show proper phase behavior. Thus, cloud point measurements are Saline water containing monovalent and/or divalent cations can suc-
performed to quantify the suitable surfactant's formulation. Usually cessfully hinder aggregation of anionic surfactants via binding to the
cloud point of a surfactant should exceed that of reservoir temperature, ionic head groups that leads to the reduction in the electrostatic interac-
proofing the stability of surfactant at high salinity and high temperature tion among them. Through salinity scan tests, it is possible to select a
conditions of the reservoir. However, this method cannot be applied for surfactant that is suitable for a specific type of oil as a function of salinity
the ionic surfactant because of its sensitivity toward temperature [86]. [93,94]. For instance, a surfactant that gives higher volumes for micro
Phase behavior tests can be accomplished at interesting tempera- emulsions of water in oil at lower salinities is preferred for LSS flooding
tures to examine the phase separation of surfactant solution over (Fig. 24). Also, there are some empirical correlations that relate the sur-
time. Nonionic surfactant is useful at high salinity and temperature sit- factant formulation to the salinity, temperature, alcohol type and con-
uations and can act as a wettability altering agent [87]. Lu et al. [88] centration and type of the surfactant [95]. However, these correlations
showed that ether sulfate surfactants including Guerbetalkoxy sulfates are based on the trial-error fittings and may not respond well to the se-
have 4 to 6 years stability at 85 °C and ultralow IFT. vere change in the salinity and oil type.
It is revealed that linear alkyl benzene sulfonate (ABS) surfactants
are not appropriate at high sanity conditions, whereas ABS surfactants
3. Conclusions
combined with an alkyl ethoxy carboxylate could endure high salinities
of seawater as the solubility improved. A formulation of ABS/alkyl eth-
LSS flooding can significantly modify the wetting behavior of oils at
oxy carboxylate of 40/60 is useful to create a stable phase with ultralow
the oil/rock/brine interfaces. The adsorption of surfactant at oil/brine in-
IFT [84].
terface permits the spreading coefficient to be changed over a wide
Moreover, sacrificial agents like sodium polyacrylate,
range. Use of smart water containing proper types of monovalent and
polyethyleneglycol, sodium carbonate, sodium metaborate, glycolic
divalent cations can control the angle of LSS/oil/rock contact line.
acid, glyceric acid and disodium EDTA can be utilized to mitigate the ad-
Aging time increases the contact angle to an equilibrium value, and con-
sorption of anionic surfactants on the carbonate, sandstone and clay
tact angle intensifies in the case of low salinity water with increasing
minerals. The presence of sacrificial agents reduce the adsorption of pri-
concentration of divalent cations. On the other hand, LSS works better
mary surfactant at high salinity environment by screening off the diva-
with a small increase in concentration of divalent, primarily due to the
lent ions from the solid/brine interface that hinder the direct contact
increase in the interfacial elasticity, because of compact structures of
between surfactant/divalents and solid/divalent [89].
surfactant at the oil/brine interface in the presence of Ca2+. Oils with
Interestingly, Gemini surfactants (a surfactant with more than one
large number of acid or base numbers adhere to the rock surface at
hydrophilic head group) are reported to be promising for low salinity
high pH thanks to the presence of divalent cations; hence, low salinity
injection in such a way that it could provide ultralow IFT and very
brine containing higher concentration of monovalents would provide
good interfacial rheology characteristics, which nominate them for
conditions for water bridging instead of cation bridging that leads to
EoR applications [90]. They are comprising greater than one hydrophilic
the oil release from the rock. At the same ionic strength for Berea sand-
head group, which can be zwitterionic, anionic, cationic or non-ionic;
stone, brines containing CaCl2 facilitates the oil adsorption much more
and a hydrophobic tail group that linked by a spacer close to the head
than NaCl can do. LSS not only can lower the IFT, but also it can hinder
groups. The hydrophobic tail may be short or long and the spacer
snap-off of the oil via increasing the interfacial elasticity. The fines mi-
group contains either polyether, aliphatic or aromatic, benzene or meth-
gration due to low salinity injection is a controversial issue; hence, if it
ylene [91,92]. Shorter length of a spacer leads to the high rheology of the
could play role, the oil desorption from the rock assists. Electrical double
surfactant that helps improvements in EoR performance. Gemini surfac-
layer expansion along with multiple ion exchange should be studied for
tants often possess a proper value for HLB of around 7 to 13, because
desorption studies of low salinity brines, because if EDL expansion alone
very high values for HLB adversely affects the IFT. Gemini surfactants
cause the detachment of oil, then diluted seawater should work with
sandstone rocks. Silica surface exposed to low salinity brine with
Na2SO4 shows less instability in water thickness than low salinity
brine containing NaCl. pH increase for both sandstone and carbonate
rocks aids oil release from the mineral surface when decreasing salinity
of brine formations used. Anhydrite, a catalyst and essential for the wet-
tability alteration in carbonate rocks, solubility in brine formation de-
creases as temperature increases and its precipitation is along with
the temperature; however, a temperature gradient from the point of in-
jection to the well-head (40 to 130 °C) is useful for the LSS, because it
can maintain the concentration of essential ions at the maximum any-
time. Finally, the changes in pH, salinity, ion composition, type of surfac-
tant, oil composition, and type of rock have profound effect on the
wettability of the rock surface. Therefore, by carefully examining the
surfactant stability in the reservoir condition and selection of appropri-
ate types of salts and a thorough screening test, it can be possible to alter
the wettability of a rock.

References
[1] K.A. Elraies, I.M. Tan, M. Awang, M.T. Fathaddin, A new approach to low-cost, high
performance chemical flooding system, Presented at the SPE Production and Oper-
Fig. 24. A proper diagram for relative phase volume as a function of salinity for a special ation Conference and Exhibition held in Tunis, Tunisia, 8–10 June, Paper SPE
surfactant [96]. 133004, 2010.
A. Aminian, B. ZareNezhad / Journal of Molecular Liquids 275 (2019) 265–280 279

[2] B. Bera, Ion and Surfactant Induced Wetting Transition, Ph.D. Thesis University of [31] M.E.J. Haagh, I. Siretanu, M.H.G. Duits, F. Mugele, Salinity-dependent contact angle
Twente, 2016. alteration in oil/brine/silicate systems: the critical role of divalent cations, Langmuir
[3] J.J. Sheng, Critical review of low-salinity waterflooding, J. Pet. Sci. Eng. 120 (2014) 33 (14) (2017) 3349–3357.
216–224. [32] M. Nourani, T. Tichelkamp, B. Gaweł, G. Øye, Desorption of crude oil components
[4] M.D. Jackson, J. Vinogradov, G. Hamon, M. Chamerois, Evidence, mechanisms and from silica and aluminosilicate surfaces upon exposure to aqueous low salinity
improved understanding of controlled salinity waterflooding part 1: sandstones, and surfactant solutions, Fuel 180 (1–8) (2016).
Fuel 185 (2016) 772–793. [33] J.J. Adams, Asphaltene adsorption: a literature review, Energy Fuel 28 (5) (2014)
[5] P.C. Myint, A. Firoozabadi, Thin liquid films in improved oil recovery from low- 2831–2856.
salinity brine, Curr. Opin. Colloid Interface Sci. 20 (2015) 105–114. [34] L.T. Zhuravlev, Concentration of hydroxyl groups on the surface of amorphous silica,
[6] P.S. Hammond, E. Unsal, Forced and spontaneous imbibition of surfactant solution Langmuir 3 (3) (1987) 316–318.
into an oil-wet capillary: the effects of surfactant diffusion ahead of the advancing [35] A. Edin, A. Skauge, Combined low salinity brine injection and surfactant flooding in
meniscus, Langmuir 26 (2010) 6206–6221. mixed-wet sandstone CODRs, Energy Fuel 24 (2010) 3551–3559.
[7] M. Sedghi, M. Piri, L. Goual, Atomistic molecular dynamics simulations of crude [36] A. Sari, Q. Xie, Y. Chen, A. Saeedi, E. Pooryousefy, Drivers of low salinity effect in car-
oil/brine displacement in calcite mesopores, Langmuir 32 (14) (2016) bonate reservoirs, Energy Fuel 31 (9) (2017) 8951–8958.
3375–3384. [37] M. Alotaibi, R. Nasralla, H. Nasr-El-Din, Wettability studies using low-salinity water
[8] M. Sedghi, M. Piri, L. Goual, Molecular dynamics of wetting layer formation and in sandstone reservoirs, SPE Reserv. Eval. Eng. 14 (2011) 713–725.
forced water invasion in angular nanopores with mixed wettability, J. Chem. Phys. [38] Q. Wang, C. Zhu, J. Yun, G. Yang, Isomorphic substitutions in clay materials and ad-
141 (2014) 194703–194712. sorption of metal ions onto external surfaces: a DFT investigation, J. Phys. Chem. C
[9] S. Chandrasekhar, H. Sharma, K.K. Mohanty, Dependence of wettability on brine 121 (48) (2017) 26722–26732.
composition in high temperature carbonate rocks, Fuel 225 (2018) 573–587. [39] M. Adeel Sohal, G. Thyne, E.G. Søgaard, Review of recovery mechanisms of
[10] B. Yuan, D.A. Wood, A comprehensive review of formation damage during enhanced ionically modified waterflood in carbonate reservoirs, Energy Fuel 30 (3)
oil recovery, J. Pet. Sci. Eng. 167 (2018) 287–299. (2016) 1904–1914.
[11] N. Kumar Maurya, A. Mandal, Investigation of synergistic effect of nanoparticle and [40] M.A. Ahmadi, S.R. Shadizadeh, Spotlight on the new natural surfactant flooding in
surfactant in macro emulsion based EOR application in oil reservoirs, Chem. Eng. carbonate rock samples in low salinity condition, Nature 8 (2018), 10985.
Res. Des. 132 (2018) 370–384. [41] P. Purswani, M.S. Tawfik, Z.T. Karpyn, Factors and mechanisms governing wettability
[12] J.J. Sheng, Optimum phase type and optimum salinity profile in surfactant flooding, alteration by chemically tuned water flooding: a review, Energy Fuel 31 (8) (2017)
J. Pet. Sci. Eng. 75 (2010) 143–153. 7734–7745.
[13] S. Strand, T. Puntervold, T. Austad, Water based EOR from clastic oil reservoirs by [42] H. Sharma, K.K. Mohanty, An experimental and modeling study to investigate brine-
wettability alteration: a review of chemical aspects, J. Pet. Sci. Eng. 146 (2016) rock interactions during low salinity water flooding in carbonates, J. Pet. Sci. Eng.
1079–1091. 165 (2018) 1021–1039.
[14] S.C. Ayirala, A.A. Yousef, L. Zuoli, Z. Xu, Coalescence of crude oil droplets in brine sys- [43] R. Gupta, G.G. Smith, L. Hu, T. Willingham, M. Lo Cascio, J.J. Shyeh, C.R. Harris, En-
tems: effect of individual electrolytes, Energy Fuel 32 (5) (2018) 5763–5771. hanced waterflood for carbonate reservoirs—impact of injection water composition,
[15] J. Tang, Z. Qu, J. Luo, L. He, P. Wang, P. Zhang, X. Tang, Y. Pei, B. Ding, B. Peng, Y. SPE Middle East Oil and Gas Show and Conference, Manama, Bahrain, 25–28 Sep-
Huang, Molecular dynamics simulations of the oil-detachment from the hydroxyl- tember, 2011 , SPE-142668.
ated silica surface: effects of surfactants, electrostatic interactions, and water flows [44] M.E.J. Haagh, I. Siretanu, M.H.G. Duits, F. Mugele, Salinity-dependent contact angle
on the water molecular channel formation, J. Phys. Chem. B 122 (6) (2018) alteration in oil/brine/silicate systems: the critical role of divalent cations, Langmuir
1905–1918. 33 (14) (2017) 3349–3357.
[16] K. Kobayashi, Y. Liang, S. Murata, T. Matsuoka, S. Takahashi, K-i. Amano, N. Nishi, T. [45] S. Tavassoli, A. Kazemi Nia Korrani, G.A. Pope, K. Sepehrnoori, Low-salinity surfac-
Sakka, Stability evaluation of cation bridging on muscovite surface for improved de- tant flooding—a multimechanistic enhanced-oil-recovery method, SPE J. 21 (2016)
scription of ion-specific wettability alteration, J. Phys. Chem. C 121 (17) (2017) 744–760 SPE-173801-PA.
9273–9281. [46] Ch. Qiao, R.T. Johns, L. Li, Modeling low salinity water flooding in chalk and lime-
[17] H. Ding, S. Rahman, Experimental and theoretical study of wettability alteration dur- stone reservoirs, Energy Fuel 30 (2) (2016) 884–895.
ing low salinity water flooding—a state of the art review, Colloids Surf. A [47] G.Q. Tang, N.R. Morrow, Influence of brine composition and fines migration on
Physicochem. Eng. Asp. 520 (2017) 622–639. crude oil brine rock interactions and oil recovery, J. Pet. Sci. Technol. 24 (1999)
[18] A. Kumar, A. Mandal, Characterization of rock-fluid and fluid-fluid interactions 99–111.
in presence of a family of synthesized zwitterionic surfactants for application in [48] A. Fogden, M. Kumar, N. Morrow, J.S. Buckley, Mobilization of fine particles during
enhanced oil recovery, Colloids Surf. A Physicochem. Eng. Asp. 549 (2018) flooding of sandstones and possible relations to enhanced oil recovery, Energy
1–12. Fuel 25 (2011) 1605–1616.
[19] V. Kedar, S.S. Bhagwat, Effect of salinity on the IFT between aqueous surfactant so- [49] F.J. Allen, L. Griffin, R.M. Alloway, P. Gutfreund, S.Y. Lee, C.L. Truscott, R.J.L. Welbourn,
lution and crude oil, Pet. Sci. Technol. 36 (12) (2018) 835–842. M.H. Wood, S.M. Clarke, An anionic surfactants on an anionic substrate: monovalent
[20] C.D. Yuan, W.-F. Pu, X.-C. Wang, L. Sun, Y.-C. Zhang, S. Cheng, Effects of interfacial cation binding, Langmuir 33 (32) (2017) 7881–7888.
tension, emulsification, and surfactant concentration on oil recovery in surfactant [50] C. Gerold, C.S. Henry, Observation of dynamic surfactant adsorption facilitated by di-
flooding process for high temperature and high salinity reservoirs, Energy Fuel 29 valent cation bridging, Langmuir 34 (4) (2018) 1550–1556.
(10) (2015) 6165–6176. [51] T. Underwood, V. Erastova, P. Cubillas, H.C. Greenwell, Molecular dynamic simula-
[21] F.-Y. Jin, S. Wang, W.-F. Pu, X.-L. Liu, C.-D. Yuan, S. Zhao, Y.-L. Tang, L. Dou, The effects tions of montmorillonite-organic interactions under varying salinity: an insight
of interfacial tension, injection rate, and permeability on oil recovery in dilute sur- into enhanced oil recovery, J. Phys. Chem. C 119 (2015) 7282–7294.
factant flooding, Pet. Sci. Technol. 34 (16) (2016) 1490–1495. [52] J.A. Greathouse, R.T. Cygan, J.T. Fredrich, G.R. Jerauld, Adsorption of aqueous crude
[22] W. Pu, Ch. Yuan, W. Hu, T. Tan, J. Hui, Sh. Zhao, S. Wang, Y. Tang, Effects of interfacial oil components on the basal surfaces of clay minerals: molecular simulations includ-
tension and emulsification on displacement efficiency in dilute surfactant flooding, ing salinity and temperature effects, J. Phys. Chem. C 121 (41) (2017) 22773–22786.
RSC Adv. 6 (56) (2016) 50640–50649. [53] F. Mugele, B. Bera, A. Cavalli, I. Siretanu, A. Maestro, M. Duits, M. Cohen-Stuart, D.
[23] H. Chen, A.Z. Panagiotopoulos, E.P. Giannelis, Atomistic molecular dynamics simula- Van den Ende, E. Stocker, I. Collins, Ion adsorption-induced wetting transition in
tions of carbohydrate-calcite interactions in concentrated brine, Langmuir 31 (8) oil-water-mineral systems, Sci. Rep. 5 (2015), 10519.
(2015) 2407–2413. [54] A. Al-Khafaj, A. Neville, M. Wilson, D. Wen, Effect of low salinity on the oil desorp-
[24] B. Wei, L. Lu, Q. Li, H. Li, X. Ning, Mechanistic study of oil/brine/solid interfacial be- tion efficiency from calcite and silica surfaces, Energy Fuel 31 (11) (2017)
haviors during low-salinity waterflooding using visual and quantitative methods, 11892–11901.
Energy Fuel 31 (6) (2017) 6615–6624. [55] T. Tichelkamp, Y. Vu, M. Nourani, G. Øye, Interfacial tension between low salinity so-
[25] J. Yang, Z. Dong, M. Dong, Z. Yang, M. Lin, J. Zhang, C. Chen, Wettability alteration lutions of sulfonate surfactants and crude and model oils, Energy Fuel 28 (2014)
during low-salinity waterflooding and the relevance of divalent ions in this process, 2408–2414.
Energy Fuel 30 (1) (2016) 72–79. [56] M.J. Alshakhs, A.R. Kovscek, An experimental study of the impact of injection water
[26] N. Kumar Jha, S. Iglauer, J.S. Sangwai, Effect of monovalent and divalent salts on the composition on oil recovery from carbonate rocks, SPE Annual Technical Conference
interfacial tension of n-heptane against aqueous anionic surfactant solutions, J. and Exhibition, Houston, Texas, USA, 28–30 September, SPE-175147, , 2015.
Chem. Eng. Data (2017)https://doi.org/10.1021/acs.jced.7b00640 Article ASAP. [57] H. Mahani, A.L. Keya, S. Berg, W.B. Bartels, R. Nasralla, W.R. Rossen, Insights into the
[27] A. Kakati, J.S. Sangwai, Wettability alteration of mineral surface during low salinity mechanism of wettability alteration by low-salinity flooding (LSF) in carbonates,
water flooding: role of salt type, pure alkanes and model oils containing polar com- Energy Fuel 29 (3) (2015) 1352–1367.
ponents, Energy Fuel 32 (3) (2018) 3127–3137. [58] Y. Lu, N. Fathinajafabadi, A. Firoozabadi, Effects of temperature on wettability of oil/
[28] F. Jiménez-Ángeles, A. Firoozabadi, Contact angle, liquid film, and liquid-liquid and brine/rock systems, Energy Fuel 31 (5) (2017) 4989–4995.
liquid-solid interfaces in model oil-brine-substrate systems, J. Phys. Chem. C 120 [59] L. Chena, G. Zhanga, L. Wang, W. Wub, J. Ge, Zeta potential of limestone in a large
(2016) 11910–11917. range of salinity, Colloids Surf. A Physicochem. Eng. Asp. 450 (2014) 1–8.
[29] N.V. Churaev, A.P. Ershov, N.E. Esipova, G.A. Iskandarjan, E.A. Madjarova, I.P. [60] J. Morag, M. Dishon, U. Sivan, The governing role of surface hydration in ion specific
Sergeeva, V.D. Sobolev, T.F. Svitova, M.A. Zakharova, Z.M. Zorin, J.-E. Poirier, Interac- adsorption to silica: an AFM-based account of the Hofmeister universality and its re-
tion of oil droplets with silica surfaces, Colloids Surf. A Physicochem. Eng. Asp. 91 versal, Langmuir 29 (2013) 6317–6322.
(1994) 97–112. [61] N. Schwierz, D. Horinek, R.R. Netz, Anionic and cationic Hofmeister effects on hydro-
[30] H. Hosseinzade Khanamiri, O. Torsæter, J. Ågestensen, Effect of calcium in pore scale phobic and hydrophilic surfaces, Langmuir 29 (2013) 2602–2614.
oil trapping by low salinity water and surfactant EOR at strongly water wet condi- [62] M. Dishon, O. Zohar, U. Sivan, Effect of cation size and charge on the interaction be-
tions: in situ imaging by X-ray microtomography, Energy Fuel 30 (10) (2016) tween silica surfaces in 1:1, 2:1, and 3:1 aqueous electrolytes, Langmuir 27 (2011)
8114–8124. 12977–12984.
280 A. Aminian, B. ZareNezhad / Journal of Molecular Liquids 275 (2019) 265–280

[63] F. Mugele, I. Siretanu, N. Kumar, B. Bera, L. Wang, R. de Ruiter, A. Maestro, M. Duits, [81] T.E. Chavez-Miyauchi, A. Firoozabadi, G.G. Fuller, Nonmonotonic elasticity of the
D. van den Ende, I. Collins, Insights from ion adsorption and contact angle alteration crude oil-brine Interface in relation to improved oil recovery, Langmuir 32 (9)
at mineral surfaces for low-salinity water flooding, SPE J. 21 (4) (2016) 38–48. (2016) 2192–2198.
[64] G. Brekke-Svaland, F. Bresme, Interactions between hydrated calcium carbonate [82] V. Alvarado, M.M. Bidhendi, G. García-Olvera, B. Morin, J.S. Oakey, Interfacial visco-
surfaces at nanoconfinement conditions, J. Phys. Chem. C 122 (13) (2018) elasticity of crude oil-brine: an alternative EOR mechanism in smart waterflooding,
7321–7330. SPE Improved Oil Recovery Symposium, Tulsa, Oklahoma, USA, 12–16 April, SPE-
[65] A. Røyne, K.N. Dalby, T. Hassenkam, Repulsive hydration forces between calcite sur- 169127, , 2014.
faces and their effect on the brittle strength of calcite-bearing rocks, Geophys. Res. [83] L. Rosenfeld, G.G. Fuller, Consequences of interfacial viscoelasticity on thin film sta-
Lett. 42 (2015) 4786–4794. bility, Langmuir 28 (2012) 14238–14244.
[66] J. Israelachvili, R.M. Pashley, Measurement of the hydrophobic interaction between [84] C. Negin, S. Ali, Q. Xie, Most common surfactants employed in chemical enhanced oil
two hydrophobic surfaces in aqueous electrolyte solutions, J. Colloid Interface Sci. 98 recovery, Petroleum 3 (2) (2017) 197–211.
(2) (1984) 500–514. [85] J.S.J. Tan, L. Zhang, F.C.H. Lim, D.W. Cheong, Interfacial properties and monolayer col-
[67] P.V. Brady, N.R. Morrow, A. Fogden, V. Deniz, N. Loahardjo, Winoto, Electrostatics lapse of alkyl benzene sulfonate surfactant monolayers at the decane-water inter-
and the low salinity effect in sandstone reservoirs, Energy Fuel 29 (2) (2015) face from molecular dynamics simulations, Langmuir 33 (18) (2017) 4461–4476.
666–677. [86] P.D. Patil, N. Rohilla, A. Katiyar, W. Yu, S. Falcone, Ch. Nelson, P. Rozowski, Surfactant
[68] J.N. Israelachvili, Chapter 14 — electrostatic forces between surfaces in liquids, in: based EOR for tight oil reservoirs through wettability alteration: novel surfactant
J.N. Israelachvili (Ed.), Intermolecular and Surface Forces, 3rd ed.Academic Press, formulations and their efficacy to induce spontaneous imbibition, SPE EOR Confer-
San Diego 2011, pp. 291–340. ence at Oil and Gas West Asia, Society of Petroleum Engineers SPE EOR Conference
[69] H.J. Butt, Measuring electrostatic, van der Waals, and hydration forces in electrolyte at Oil and Gas West Asia-Muscat, 2018.
solutions with an atomic force microscope, Biophys. J. 60 (6) (1991) 1438–1444. [87] Y. Muhammad, A. Al Sowaidi, A. Al Obeidli, T.W. Willingham, C. Britton, S. Adkins, M.
[70] J. Wu, F. Liu, G. Chen, X. Wu, D. Ma, Q. Liu, Sh. Xu, Sh. Huang, T. Chen, W. Zhang, H. Delshad, R. Xiao, G.-F. Teletzke, Chemical formulation design in high salinity, high
Yang, J. Wang, Effect of ionic strength on the interfacial forces between oil/brine/ temperature carbonate reservoir for a super giant offshore field in Middle East,
rock interfaces: a chemical force microscopy study, Energy Fuel 30 (2016) 273–280. SPE Abu Dhabi International Petroleum Exhibition, Society of Petroleum Engineers,
[71] R. Nasralla, H. Nasr-El-Din, R. Nasralla, Double-layer expansion: is it a primary Abu Dhabi, UAE, 2017.
mechanism of improved oil recovery by low-salinity waterflooding? SPE Reserv. [88] J. Lu, P.J. Liyanage, S. Solairaj, S. Adkins, G.P. Arachchilage, D.H. Kim, C. Britton, U.
Eval. Eng. 17 (1) (2014) 49–59. Weerasooriya, G.A. Pope, New surfactant developments for chemical enhanced oil
[72] M. Ricci, P. Spijker, F. Stellacci, J.-F. Molinari, K. Voïtchovsky, Direct visualization of recovery, J. Pet. Sci. Eng. 120 (2014) 94–101.
single ions in the stern layer of calcite, Langmuir 29 (2013) 2207–2216. [89] H. Shamsijazeyi, G. Hirasaki, R. Verduzco, Sacrificial agent for reducing adsorption of
[73] S. Kerisit, S.C. Parker, Free energy of adsorption of water and metal ions on the {101 anionic surfactants, SPE International Symposium on Oilfield Chemistry-Society of
4} calcite surface, J. Am. Chem. Soc. 126 (2004) 10152–10161. Petroleum Engineers. The Woodlands, Texas, USA, 2013.
[74] G. Yang, T. Chen, J. Zhao, D. Yu, F. Liu, D. Wang, M. Fan, W. Chen, J. Zhang, H. Yang, J. [90] Z. Dong, Y. Zheng, J. Zhao, Synthesis, physico-chemical properties and enhanced oil
Wang, Desorption mechanism of asphaltenes in the presence of electrolyte and the recovery flooding evaluation of novel zwitterionic gemini surfactants, J. Surfactant
extended Derjaguin-Landau-Verwey-Overbeek theory, Energy Fuel 29 (7) (2015) Deterg. 17 (6) (2014) 1213–1222.
4272–4280. [91] S.F. Tang, X.D. Hu, X.N. Ouyang, S.X. Yan, S.C. Wen, Y.L. Lai, Experimental study of an-
[75] D. Yu, H. Yang, H. Wang, Y. Cui, G. Yang, J. Zhang, J. Wang, Interactions between col- ionic gemini surfactant enhancing waterflooding recovery ratio, Adv. Mater. Res.
loidal particles in the presence of an ultrahighly charged amphiphilic polyelectro- 361 (2012) 469–472.
lyte, Langmuir 30 (48) (2014) 14512–14521. [92] Z. Bi, W. Liao, L. Qi, Wettability alteration by CTAB adsorption at surfaces of SiO2 film
[76] H. Guo, A simple algorithm for fitting a Gaussian function, IEEE Signal Process. Mag. or silica gel powder and mimic oil recovery, Appl. Surf. Sci. 221 (2004) 25–31.
28 (5) (2011) 134–137. [93] R. Aparna, C.P. Thomas, Y. Bian, J.T. Kwan, M. Salehi, G.J. Hirasaki, M.C. Puerto, C.
[77] Z.M. Aman, A. Haber, N.N.A. Ling, A. Thornton, M.L. Johns, E.F. May, Effect of brine sa- Maura, C.A. Miller, Laboratory studies for surfactant flood in low-temperature,
linity on the stability of hydrate-in-oil dispersions and water-in-oil emulsions, En- low-salinity fractured carbonate reservoir, SPE International Symposium on Oilfield
ergy Fuel 29 (12) (2015) 7948–7955. Chemistry, The Woodlands, Texas, USA, 2013.
[78] C.G. Quintero, C. Noïk, C. Dalmazzone, J.L. Grossiord, Formation kinetics and visco- [94] S. Lglauer, Y. Wu, P. Shuler, Y. Tang, W.A. Goddard III, New surfactant classes for en-
elastic properties of water/crude oil interfacial films, Oil Gas Sci. Technol. 64 (5) hanced oil recovery and their tertiary oil recovery potential, J. Pet. Sci. Eng. 71
(2009) 607–616. (2010) 23–29.
[79] E.L. Nordgård, S. Simon, J. Sjöblom, Interfacial shear rheology of calcium [95] R.E. Anton, J.M. Anderez, C. Bracho, F. Vejar, J.L. Salager, Practical surfactant mixing
naphthenate at the oil/water Interface and the influence of pH, calcium, and in pres- rules based on the attainment of microemulsion-oil-water three-phase behavior
ence of a model monoacid, J. Dispers. Sci. Technol. 33 (7) (2012) 1083–1092. systems, Adv. Polym. Sci. 218 (2008) 83–113.
[80] B. Wei, R. Wu, L. Lu, X. Ning, X. Xu, C. Wood, Y. Yang, Influence of individual ions on [96] A. Bera, K. Ojha, A. Mandal, T. Kumar, Interfacial tension and phase behavior of
oil/brine/rock interfacial interactions and oil water flow behaviors in porous media, surfactant-brine-oil system, Colloids Surf. A Physicochem. Eng. Asp. 383 (2011)
Energy Fuel 31 (11) (2017) 12035–12045. 114–119.

You might also like