You are on page 1of 12

Review

pubs.acs.org/EF

Factors and Mechanisms Governing Wettability Alteration by


Chemically Tuned Waterflooding: A Review
Prakash Purswani,# Miral S. Tawfik,# and Zuleima T. Karpyn*
John and Willie Leone Family Department of Energy and Mineral Engineering, The Pennsylvania State University, University Park,
Pennsylvania, United States

ABSTRACT: Enhanced oil recovery (EOR) techniques are aimed at improving the recovery efficiency of mature oil fields in
secondary and tertiary recovery modes. In particular, chemically tuned waterflooding (CTWF) has been a rising EOR technique
toward improving oil recovery from rocks that are difficult to produce due to their initial wetting state, e.g., oil-wet and
intermediate-wet. With an increasing oil-wetting affinity of a reservoir rock, extraction of oil becomes more challenging. As such,
wettability alteration has been identified as the primary mechanism for oil recovery from oil-wet and intermediate-wet rock types.
Recently, researchers have attempted to categorize the factors and mechanisms governing wettability alteration by CTWF.
Multiple studies have identified the importance of brine salinity and ion composition on promoting wettability alteration to a
more favorable water-wet state. Reservoir temperature, the surface charge of the rock, and the surface active components of crude
oil are also reported to influence wettability alteration and therefore oil recovery from waterflooding. In this paper, we present an
extensive literature review on the subject of wettability alteration, with an emphasis on experimental work conducted on
carbonate and sandstone rocks, as they constitute the majority of the oil reserves in the world. The purpose of this review paper is
to synthesize the current state of knowledge regarding the factors and mechanisms that govern wettability alteration by CTWF
and, through this exercise, set the platform to pose new research questions.

1. BACKGROUND ON WATERFLOODING of injected brine. Several laboratory studies have been


In 1865, the first waterflood occurred by an accidental water conducted to explain the mechanism by which LSWF improves
leak from shallow sands in the Pithole City area in oil recovery in sandstones.6−8 Zhang et al. (2006) further
Pennsylvania, USA.1 A significant increase in oil recovery was suggested that brine composition rather than salinity causes
observed and attributed to the flow of water through the oil IOR in carbonate rocks.9 This triggered numerous experimental
sands, which displaced the oil and increased the reservoir studies that focus primarily on identifying and optimizing the
pressure upon the depletion of the reservoir’s natural energy. key factors that affect the success of CTWF.
Since then, waterflooding has been slowly growing in Early on, the focus was only on understanding LSWF in
applicability until the 1950s, at which time it was recognized sandstone reservoirs due to its less complex structure, as well as
as the leading secondary oil recovery mechanism.1 the common understanding that IOR due to LSWF is due to
Reinjection of produced water has been the most common interactions of the injected brine primarily with clay minerals.6
practice in waterflooding to avoid formation damage, as well as On the contrary, carbonate rocks have minimal to no clay
to dispose and make use of produced water. However, volumes content,10 and hence, LSWF was not believed to be a viable
of produced water are rarely sufficient to replenish the reservoir enhanced oil recovery (EOR) method for carbonate reservoirs.
energy. In addition, the use of produced water requires more However, Bagci et al. (2001) first reported IOR due to LSWF
complex surface facilities to separate suspended solids, oil in carbonates.11 The recovery factors of carbonate reservoirs are
droplets, and dissolved hydrocarbons, as well as the corrosive lower than those of sandstone reservoirs.12 This is because of
components in order to avoid scaling and/or corrosion in the complex characteristics of carbonate rocks, which arise
injector wells.2 Surface facilities might not be readily available mainly due to their initial wetting state.13,14
on site, especially on offshore fields where space on platforms is Anderson (1986) defined wettability as “the ability of a fluid
limited. Hence, the injection of seawater has been deemed to spread or adhere to the rock surface in the presence of
more economically and technically convenient than treated another immiscible fluid.”15 While most sandstone reservoirs
produced water due to its abundance, lower cost, and
are strongly water-wet or mixed-wet systems, most carbonate
convenience for offshore use.2
reservoirs are either mixed-wet or completely oil-wet.13,14 The
It has been observed through several experimental and field
studies that the injection of seawater results in improved oil term “mixed wettability” was first coined by Salathiel (1973),
recovery (IOR), as compared to the injection of produced where the author suggested that reservoirs are primarily water-
water,3,4 which has been attributed to the lower salinity of wet in nature during formative stages, but, in later stages,
seawater compared to the higher salinity of produced water. organic components of oil come in contact with the rock
Martin (1959) suggested that reducing the salinity of water
could improve oil recovery in sandstone rocks.5 Chemically Received: April 17, 2017
tuned waterflooding (CTWF) initially emerged in the form of Revised: June 13, 2017
low salinity waterflooding (LSWF), which involved the dilution Published: July 10, 2017

© XXXX American Chemical Society A DOI: 10.1021/acs.energyfuels.7b01067


Energy Fuels XXXX, XXX, XXX−XXX
Energy & Fuels Review

Table 1. A Summary of Coreflooding Experiments Performed on Carbonate and Sandstone Cores for a Range of Experimental
Temperatures, Recovery Modes, and Brine Compositionsa

a
Sands are highlighted in grey.

B DOI: 10.1021/acs.energyfuels.7b01067
Energy Fuels XXXX, XXX, XXX−XXX
Energy & Fuels Review

surface, rendering some portions of the reservoir oil-wet.16 The 2. BRINE SALINITY AND ION COMPOSITION
degree of wettability of a rock is governed by the nature of the LSWF is considered a subset of CTWF as it involves the overall
crude oil−brine−rock system. The oil-wet nature of carbonate dilution of the injected brine while maintaining a constant ion
rocks can be attributed to the tendency of organic components composition. Early experimental studies focused on LSWF in
from the oil to adsorb onto the surface of the rock more sandstone and slowly progressed toward carbonates.6,10,11
strongly, making carbonate reservoirs harder to produce. One Several of these experiments concluded that decreasing the
of the earliest works on wettability modification was carried out salinity of injected water leads to an increase in oil recovery, as
by Wagner et al. (1959), where they performed a series of well as a corresponding increase in the water-wetness of the
waterflooding experiments on synthetic calcite and quartz cores rock in both carbonate and sandstone reservoirs.17,44−48 To
with brines of different salinities and pH to observe the positive illustrate the phenomenon of LSWF in carbonate reservoirs,
effects of wettability modifications for improving oil recovery.17 zeta potential measurements were conducted by Yousef et al.
Furthermore, the importance of ion composition present in (2012). Zeta potential, also known as “electrokinetic potential,”
seawater was recognized as a critical factor affecting the process is useful toward understanding the interactions among particles
of wettability alteration. This led to the use of tailored brine in a suspension and has been widely used for finding the surface
solutions made from seawater in the secondary and tertiary charge of a rock in a particular brine environment. The
recovery stages for improving oil recovery. Work has also been measurements showed that as the salinity of brine decreases,
done on seawater flooding to prove that wettability alteration zeta potential of the carbonate surface decreases, thereby
indeed plays an important role in improving oil recovery.9,18−23 suggesting that Ca2+ ions from the rock surface migrate toward
Many laboratory studies were conducted on carbonate and the low salinity brine to re-establish chemical equilibrium
sandstone reservoirs to verify whether the injection of brines of between the rock and the brine.49 This effect is commonly
different salinities and ion compositions could disturb the known as mineral dissolution and is considered one of the
chemical equilibrium between the rock and crude oil to possible mechanisms contributing to an improvement in oil
recovery by CTWF in carbonate rocks. Also, the decrease in the
facilitate oil recovery. Mg2+, Ca2+, and SO42− ions present in
positive charge on the carbonate surface, when exposed to low
seawater were experimentally proven to be the potential
salinity water, leads to the expansion of the electric double layer
determining ions responsible for incremental oil recov-
(EDL), which results in a thicker, more stable water film. EDL
ery.9,19,20,24,25 Further, the contribution of the nonparticipating expansion is also considered one of many mechanisms used for
ions was analyzed, where it was seen that Na+ and Cl− ions did explaining wettability alteration by LSWF that highly depends
not contribute toward oil recovery, and therefore, seawater on the electrostatic interaction between the brine/oil and
salinity and composition could be tuned by optimizing the brine/rock interfaces.50,51 A similar trend in the results was
potential ions at the expense of the nonparticipating ions.23,26 obtained by Alotaibi et al. (2011), where zeta potential values
Despite the numerous experimental, numerical, and field of sandstone rocks with differing clay contents were measured
studies on CTWF, there is still little consensus on the in brines of different salinities. It was observed that as the brine
mechanism(s) by which it improves oil recovery.27−30 The salinity was decreased, there was an decrease in zeta potential
mechanistic explanation for IOR by CTWF in carbonates is for all sandstone rock types, suggesting a thicker EDL with a
even less explored compared to sandstone reservoirs.30 This lower salinity brine.52
lack of understanding of the underlying mechanism(s) of With the advancement in technology, techniques like X-ray
CTWF leads to inconsistency in the reported results, where microcomputed tomography (micro-CT) have been used to
some experimental studies report an IOR by CTWF,31−33 while study the effect of LSWF. This pore scale methodology helps to
others report no additional recovery.34−36 The primary focus of track fluid flow behavior during waterflooding techniques in
this work is to document and analyze experimental work done porous media. In the study conducted by Khishvand et al.
thus far about CTWF as a potential EOR technique, provide a (2016), micro-CT image analysis was used to investigate in situ
comprehensive report of the key factors affecting the success of changes that occurred in fluid saturations via wettability
CTWF, and propose mechanisms by which CTWF improves alteration upon waterflooding with brines of different salinities
oil recovery in carbonate and sandstone reservoirs. Some of the in sandstone rocks.53 They reported higher oil recovery for the
most relevant experiments discussed in this literature review are brine of lower salinity, suggesting that wettability alteration,
presented in Table 1, as a comparative presentation of tested caused by mineral migration and/or multi-ion exchange (MIE)
conditions and reported gains in oil recovery. The effects of at the rock surface, potentially enabled the deeper invasion of
the brine into larger oil-filled pores to improve oil recovery.
surface properties of the rock, crude oil, and brine composition,
This was analyzed by measuring oil−water contact angles using
the salinity of the imbibing fluid, and reservoir temperature,
high resolution images captured during the experiments.53
together with characterizing techniques like zeta potential Although the majority of the experimental studies report a gain
measurement and acid number measurement, have been with LSWF, it is valuable to note that the injection of low
studied to understand the interactions that take place at the salinity brine, incompatible with the formation water, could
oil/rock/brine interface. This paper presents a comprehensive lead to formation damage and consequently a decrease in
literature review in the area of wettability alteration for production. Therefore, maintaining an optimum salinity level,
enhancing oil recovery, and it is intended to assist further in which the improvement in oil recovery outweighs formation
development of the existing and emerging mathematical models damage, is crucial.5,6
for predicting oil recovery through CTWF. In addition, we In 1999, the effect of injected brine composition rather than
explore a series of research questions and expose knowledge overall salinity was first experimentally investigated by
gaps that could guide further studies to improve our coreflooding and spontaneous imbibition experiments using
understanding of wettability alteration via CTWF. Berea sandstone cores.45 Further studies investigated the effect
C DOI: 10.1021/acs.energyfuels.7b01067
Energy Fuels XXXX, XXX, XXX−XXX
Energy & Fuels Review

of individual ions in seawater, including Ca2+, Mg2+, SO42−, groups (R-COO−) in crude oil, thus decreasing the chances of
Na+, and Cl−. Almost all experimental results concluded that readsorption of oil onto the carbonate surface.58
Ca2+, Mg2+, and SO42− are potential determining ions, owing to In the study by Gupta et al. (2011), coreflooding
the tendency of these ions to influence the rock surface charge experiments were performed to investigate the effect of
by altering the wettability to a more water-wet system.25,54,55 It replacing sulfate ions with other anions, including phosphate
is worth noting, however, that the overall salinity (ionic (PO42−) and borate (B(OH)4−).41 The coreflood experiments
strength) of the injected brine influences the activity of the were performed on several limestone and dolomite cores in
potential determining ions, which is in accordance with the tertiary mode, where the cores were first flooded with
extended Debye−Hückel model that states that the activity formation water, then with chemically tuned water. It was
coefficient of individual ions in a solution can be determined observed that adding four times the amount of SO42− ions in
using the following equation: seawater resulted in an additional recovery of 6% of the original
oil in place (OOIP). Replacing SO42− ions with B(OH)4− ions
Azi2 I caused an incremental recovery of 15.6% of OOIP, whereas
−log γi = replacing SO42− ions with PO42− ions led to an even higher
1 + aiB I
incremental recovery of 21.3% of OOIP.41
Connate water is also identified as a primary factor for a
where γi is the activity coefficient of ionic species i, zi is the
successful LSWF in sandstone reservoirs, without which there is
charge of ionic species i, I is the ionic strength of the solution, A
no additional oil recovery.45,59 This is because the connate
and B are constants depending on temperature and dielectric
water saturation history and ion composition dictate the
constant, and ai is the effective diameter of the ion. Hence, both
stability of the water film formation over the rock surface, and
ion composition and salinity are key parameters for engineering
therefore, has a direct effect on the initial wetting condition of
an optimum injection brine for maximum incremental oil
the rock. Most experimental studies pertaining to CTWF focus
recovery.56
on the effect of the salinity and composition of injected brine
Zeta potential measurements of rock particles suspended in
on oil recovery, whereas very few studies attempt to investigate
brine solutions have also been valuable in identifying these the effect of connate water properties on oil recovery. A study
potential determining ions responsible for improving oil conducted by Shehata et al. (2016) investigated the effect of
recovery. It was found that both Ca2+ ions and SO42− ions connate water salinity and concentrations of different cations
are able to change the surface charge of chalk particles when on the performance of LSWF in two types of cores: Bandera,
added to the chalk suspension.57 This was further verified in the and Buff Berea sandstone cores.46 The study did not focus on
work by Zhang and co-workers (2006, 2007), where the zeta the effect of the concentration of anions because connate water
potential of chalk suspension was measured at a fixed pH of 8.4 normally has a low concentration of sulfate ions. This is
by increasing the concentration of the potential determining partially due to the precipitation of anhydrite (CaSO4),
ion in consideration. They observed that increasing Mg2+ and especially at higher temperatures.39,60 The study concluded
Ca2+ ion concentrations increased the zeta potential of the rock that a higher salinity connate water results in a higher
sample, owing to the affinity of these ions toward the rock incremental recovery (7% of OOIP) by LSWF. Also, connate
surface. For SO42− ions, a decrease in zeta potential was water with a high concentration of cations, particularly, divalent
observed with an increasing SO42− ion concentration, owing to cations resulted in higher incremental oil recovery when a low
the possible adsorption of the negative ion onto the rock salinity brine is injected. These conclusions are consistent with
surface.9,54 the explanation provided earlier, where, in sandstones, higher
Tang and Morrow (1999) studied the effect of cation valency salinity and higher valence cations lead to thin unstable water
on the performance of low-salinity waterflooding (LSWF) in films on grain surfaces, and therefore, promotes the cationic
sandstones.45 Monovalent (Na+), divalent (Ca2+), and trivalent bridging of crude oil onto the clay surface, which in turn results
(Al3+) cations were used in different injection brines. The study in an increased initial oil-wetness and more trapped oil that can
suggested that, at relatively higher salinities, lower valence then be recovered by means of CTWF.
cations lead to a higher recovery by spontaneous imbibition. 2.1. Open Questions. Experimental studies conducted thus
The presence of higher valence cations promotes the far focus on the effect of potential determining ions on
adsorption of polar organic compounds onto the negatively recovery. The effect of other ions in seawater has not been
charged sandstone rock surface by “cationic bridging,” thus explicitly investigated, including HCO3− and K+. Also, the effect
decreasing its water-wetness, and consequently, trapping more of SO42− ions has been widely studied due to its presence in
oil and leading to lower recovery. However, at lower salinity, seawater. However, a high SO42− ion concentration increases
the effect of cation valency on the wetting condition and oil the risk of reservoir souring and scaling.25 Hence, the effect on
recovery was less significant.45 As opposed to the conclusion oil recovery of replacing SO42− ions with other anions such as
arrived at by Tang and Morrow (1999) for sandstones, borate and phosphate, and their behavior, needs to be further
coreflooding and spontaneous imbibition of oil from carbonate examined.41 Exploring the possibility of using anions other than
cores (particularly limestone) show that an increase in divalent SO42− to find the optimal injection brine composition is key.
cation concentrations (Ca2+ and/or Mg2+) in injected brine, Other anions such as nitrate (NO3−) can have lower associated
and a decrease in monovalent cations (Na+), lead to an increase risks and operational problems compared to SO42− ions, and
in oil recovery in the presence of SO42− ions.20,25,54,55 This can therefore, its ability to improve oil recovery needs to be
be explained by the adsorption of SO42− ions onto the investigated. In addition, only a few studies tackled the effect of
positively charged carbonate surface at the brine-rock interface, connate water on the performance of CTWF. The effects of
which, in turn, leads to the attraction of Ca2+ and Mg2+ closer salinity and cation valence have been addressed in the study by
toward the carbonate surface. In addition, at the brine−oil Shehata et al. (2016). However, the effect of different anions in
interface, divalent cations form bonds with the carboxylate the connate water on the performance of LSWF has not been
D DOI: 10.1021/acs.energyfuels.7b01067
Energy Fuels XXXX, XXX, XXX−XXX
Energy & Fuels Review

studied because formation water has more cations than as well as stronger adhesion forces between dolomite and oil.63
anions.61 More importantly, the effects of salinity and the Also, contrary to outcrop limestones, outcrop dolomite
composition of connate water on incremental oil recovery via responds to chemically tuned water in the same way as
CTWF need to be quantified as connate water properties reservoir dolomite cores.43
significantly affect the optimum injection water design due to Sandstone reservoirs, on the other hand, are composed of
the inevitable injected and in situ brine mixing in the reservoir. different minerals, most commonly quartz and feldspar. These
minerals make up the matrix and are bound by secondary
3. MINERAL COMPOSITION OF THE ROCK minerals, known as “cement” that form after the deposition of
One important factor that is often neglected when comparing the sandstone matrix. There is a wide variety of minerals that
results from various studies is the fact that reservoirs are often make up cement, including anhydrite, dolomite, and clay
heterogeneous systems. Because of the reactive nature of minerals such as kaolinites, chlorites, and Illites, which are
CTWF, and its dependence on oil/rock/brine interactions, nonswelling clays, as well as montmorillonites, which are
understanding the effect of rock properties on the success of swelling clays.64 In sandstones, clays are believed to play a
CTWF is necessary for valid comparisons. Different rock types significant role in the success of CTWF due to their abundance
greatly differ in mineral composition, which directly impacts the and large specific surface area compared to matrix-forming
nature of the surface charge of the rock, and is therefore minerals.34 However, clays that swell and clays that have
important in deliberating the initial wetting state of the positive zeta potential are considered detrimental to oil
reservoir. recovery. Studies by Tang and Morrow (1999) and
Carbonate reservoirs are normally composed of a variety of Wickramathilaka et al. (2011) suggest that sandstone core
minerals, primarily calcite (CaCO3), dolomite (CaMg(CO3)2), samples with higher clay content exhibit higher incremental
anhydrite (CaSO4), and quartz (SiO4). The degree of recovery. However, none of these studies reported which
improvement in oil recovery has been observed to vary based type(s) of clay minerals have a greater impact on the low-
on the mineral composition of carbonate rocks.58 According to salinity effect.8,65 According to Rezaei Gomari et al. (2015),
Austad et al.,62 injecting diluted seawater into chalk cores both, the amount and the type of clay play a significant role in
resulted in a drastic decrease in oil recovery compared to LSWF.66 Some clays act as cation exchangers, whereby clays
injecting seawater. This showed that low salinity waterflooding tend to swap or exchange cations with cations from the fluids in
is not the reason for IOR in chalk. In the same study, coreflood the pore space due to the deficiency in their positive charge,
experiments were performed on limestone core sample- which must be countered by cations from the surrounding
s―one that did not contain an anhydrite mineral as part formation water if the clay particles are to remain electrically
of its composition, and on another limestone sample that did. neutral.67 Cation exchange capacity (CEC) is a measure of the
The study concluded that LSWF effect could only be seen in excess negative charge on the clay particles. CEC varies by clay
the presence of anhydrite, indicating that the presence of type in the following descending order: montmorillonites,
anhydrite plays an important role toward wettability alter- Illites, and kaolinites.68 According to Austad et al. (2010), clays
ation.62 Injection of low-salinity water into carbonate cores in with higher CEC, coupled with the presence of divalent cations
the presence of anhydrite promotes the dissolution of anhydrite in formation water, and coadsorption of these cations and the
due to the low concentration of Ca2+ in the injected brine polar oil components on the clay surface, are more favorable for
compared to that of formation water (common ion effect). observing low-salinity effects.38 Additionally, it was noted in
Anhydrite dissolution leads to the in situ generation of free this study that for clayey sandstones, the main parameter that
SO42− ions, which adsorb onto the positively charged carbonate needs to be optimized in the injected brine is the concentration
surface and catalyzes desorption of carboxylic groups, altering of divalent cations, which should be low. An experimental study
the wettability to a more water-wet state and thereby improving conducted by Cissokho et al. (2010) confirmed these findings,
oil recovery. On the other hand, the injection of the same low- where an incremental oil recovery of 10% was observed when a
salinity water into limestone cores in the absence of anhydrite sandstone core sample with 9.2% clay content was flooded with
shows no improvement in oil recovery. This shows that the low-salinity brine.69 The clays included chlorite, muscovite, and
presence of anhydrite is necessary for IOR in carbonate rocks, Illite, which indicates that the low-salinity effect can be
where the injected brine contains minimal SO42− ions. It has witnessed in the absence of kaolinite. On the other hand,
been experimentally observed that the dissolution of anhydrite other studies suggest that kaolinite is considered a requirement
is dependent on temperature, connate water salinity, and for effective LSWF, due to its oil-wet nature and high surface
concentration of Ca2+ ions in the injected brine.39 charge density, which makes its surface electric effect much
In the study conducted by Shariatpanahi et al. (2016), a greater when compared to other clay minerals.70 The study
comparison between the low salinity effect in Silurian outcrop concluded that a variation in kaolinite content from 5 to 30%
dolomite cores, and in chalk and limestone from previous results in a change in the overall cation exchange capacity
studies, was presented.43 The study concluded that similar to (CEC) of 47−270 eq/m3, which translates to ∼2.65%
limestone rocks, the presence of sulfate ions (whether in the incremental recovery at 30% kaolinite content. The importance
injected brine or as a result of dissolution of anhydrite) is of kaolinite mineral was also investigated in the pore scale study
critical to observing wettability alteration, as well as an conducted by Lebedeva et al. (2011). The authors used X-ray
improvement in oil recovery. However, the extent of adsorption micro-CT imaging to quantify the effect of LSWF via
of sulfate ions onto the dolomite surface is observed to be wettability alteration on sand packs coated with or without
weaker compared to chalk and limestone cores. This was also kaolinite mineral.71 Micro-CT image analysis helped track oil-
confirmed in the study performed by Mahani et al. (2015), blob occupancy changes that occurred during waterflooding
where CTWF in dolomites showed a smaller or no contact experiments. They found that, for the sand packs coated with
angle change compared to limestones (which are mainly kaolinite mineral, low salinity brine injection improved oil
composed of calcite), implying fewer wettability modifications, recovery as opposed to the high salinity brine, suggesting that
E DOI: 10.1021/acs.energyfuels.7b01067
Energy Fuels XXXX, XXX, XXX−XXX
Energy & Fuels Review

the oil-wetness caused by the presence of kaolinite mineral was charged.73 The variety of experimental work on zeta potential
altered to a favorable water wetting state, facilitating oil measurements clearly showcases the variation in the surface
recovery. Because of the lower resolution of the micro-CT charge response of the rock samples with different injected
images (sufficient for distinguishing oil, brine, and rock entities brine properties, indicating that rock properties have a
to suitably quantify fluid saturations), wettability analysis was significant effect on the performance of CTWF.
performed using receding and advancing oil-drop contact angle 3.1. Open Questions. The inconsistencies observed in
measurements conducted on quartz slices that mimicked the experimental results using sandstone cores show difficulty in
rock used for waterflooding experiments. These confirmed the generating repeatable and reproducible data due to the
process of wettability alteration during LSWF.71 heterogeneity of the sandstone rock mineralogy, as well as
Carbonate rock surfaces are mostly positively charged, the wide variety of clays involved. The contradictions in the
whereas sandstone rock surfaces are negatively charged. The experimental studies on sandstones can also be attributed to the
positive charge on the carbonate surface can be attributed to interplay of several pore-scale mechanisms, contributing to the
the hydration of the calcite surface to form >CaOH2+,58 overall low-salinity effect. Controlled experiments have been
whereas the negative surface charge in the case of sandstones is conducted on carbonates to assess the effect of particular
attributed to the presence of negatively charged clay particles as minerals such as calcite, dolomite, and anhydrite on the
described previously.15,58 The surface charge of the rock is performance of CTWF. However, more controlled experiments
sensitive to changes in the surrounding environment, including are yet to be conducted to map the effect of different minerals
the ionic concentration and pH of the brine around the rock, and clay types in sandstones.
which in turn affects the wetting state of the rock.15 Therefore, In addition, some studies suggest that zeta potential varies
efforts have been made to carefully measure the rock surface with particle size;75 however, there has been little work focusing
charge in different brine environments to more accurately study on the sensitivity of zeta potential to particle size. Further, most
the interactions of oil and brine with the rock surface. One conventional methods of measuring zeta potential at a
approach of estimating the surface charge of the rock is through laboratory scale, like the electrophoretic mobility measurements
zeta potential measurements as described earlier. In the work or the electroacoustic measurements, assume that zeta potential
by Vdovic and Biscan (1998), zeta potential values of synthetic does not depend on particle size.76 However, Nakatuka et al.
and natural calcite in simple NaCl brine environment were (2015) studied the importance of particle size distribution on
investigated. They found that the zeta potential value of zeta potential by the application of the liquid sedimentation
synthetic calcite was positive at lower pH due to the prevalence method. They discovered that the negative zeta potential values
of positive surface species like Ca2+, CaHCO3+, whereas natural for the silica particles do indeed depend on the particle size,
calcite showed negative zeta potential, which was attributed to such that the zeta potential values increased with the decrease
the adsorption of organic material onto the calcite surface.72 As in particle size. The authors attributed the increase in the zeta
CTWF gained momentum in the oil industry, zeta potential potential values to particles with smaller size, due to the
measurements were performed in more complex brine increased random movements of smaller particles rather than
environments. Experiments conducted by Mahani et al. larger particles.76 Therefore, the reliability of zeta potential
(2015) focused on the importance of ion composition and measurements at the laboratory scale, and their true
pH of the brine system on the oil/rock interface.30 Zeta interpretation at the field scale, remain an area of investigation.
potentials of carbonate rock samples and oil droplets were
measured as a function of pH, in different brine environments, 4. CRUDE OIL COMPOSITION
namely, formation water, seawater, and diluted seawater. An Heavier components of crude oil, like asphaltenes or resins, are
increase in the zeta potential values was seen for both dolomite believed to carry compounds bearing oxygen, nitrogen, and
and limestone rock particles with an increasing pH for all brine sulfur atoms.77 These compounds are responsible for the
systems, suggesting that the surface of the rock acted more acidity and/or basicity of crude oils. Tests done by Seifert and
positively charged at higher pH ranges, whereas, at lower pH Howell (1969) confirmed the presence of carboxylic groups in
ranges, the rock particles showed a negative zeta potential California crude, which, the authors suggested, were active at a
owing to the increased adsorption of SO42− ions in seawater.30 particular pH of the system.78 Fathi et al. (2011) further
In a different set of experiments performed by Karimi et al. reinforced that the wetting properties of crude oil were affected
(2015), the zeta potential of carbonate rock samples was by the chemical properties of the carboxylic material present in
measured both before and after aging,73 which is the the crude oil.79 The presence of carboxylic acids in the crude oil
accelerated process of exposing the core sample to reservoir was further confirmed by the electrokinetic study conducted by
conditions in order to assist in the development of an initially Mahani et al. (2015). They measured the zeta potential of oil
oil-wet sample.74 This helped in finding the change in the droplets, as suspended particles, in different brine environments
nature of the surface charge, which is necessary for under- and observed that the oil particles showed a negative zeta
standing the surface wetness. They found that the zeta potential potential over the range of pH measurements in different brine
of the rock sample changed from +29.6 mV to −33.4 mV for salinities, owing to the presence of negative carboxyl groups in
preaged and postaged calcite samples, respectively, at a pH of 8. the crude oil.63
The positive nature of the surface charge before aging was It is therefore understood that the acidic and basic
attributed to the prevalence of positive divalent Ca2+ and Mg2+ components present in the crude oil are crucial in under-
ions present in the calcite sample below pH of 10. The authors standing the extent of the adsorption of hydrocarbon
suggested that at a higher pH, the calcite surface acts more components on the rock surface, as well as the initial wetting
negatively charged due to the possible adsorption of CO32− characteristics of rocks.15,80 According to Zhang et al. (2007),
ions. The negative nature of the surface charge postaging was in order to obtain a satisfactory oil recovery from carbonate
attributed to the adsorption of polar oil components onto the reservoirs with low water wetness, initial wetting conditions
rock surface, making the rock surface more negatively must be altered.17 For carbonate rocks, where the surface is
F DOI: 10.1021/acs.energyfuels.7b01067
Energy Fuels XXXX, XXX, XXX−XXX
Energy & Fuels Review

primarily positively charged, the adsorption of the negatively and molecular structure is more important than measuring the
charged acidic components from the crude oil plays a key role amount of acids in regard to interfacial and wettability
in determining the potential for oil recovery from these properties.83 Hence, additional experimental studies are needed
reservoirs. Therefore, the quantification of the acidic content of for a more accurate quantification of the effects of different
crude oil has gained importance in the recent past. types and structures of polar components that contribute to the
One way of quantifying the amount of acidic or basic initial oil-wetting extent of the reservoir rocks.
components in crude oil is through acid and base number
measurements. The widely accepted method for estimating 5. EFFECT OF TEMPERATURE ON SURFACE
these measurements is by potentiometric titration. Standardized REACTIONS
workflows were developed by the American Society for Testing Because of the reactive nature of CTWF, temperature is a
and Materials (ASTM) and have seen several revisions over crucial factor that affects the wettability alteration process. The
time. The most updated revisions are ASTM D664 (2011) for role of temperature is 2-fold, as it affects the activity of the ionic
acid number measurements81 and ASTM D2896 (2015) for species present in the brine, as well as the interaction of the
base number measurements.82 In recent years, there have been polar organic compounds in crude oil with the rock surface.
improvements in the application of these methods. One widely Together, these can affect the type of chemical reactions that
accepted improvement in the measurement of acid number was take place at the oil−water and fluid−rock interfaces.
proposed by Fan and Buckley (2007). They suggested that the In the experiments conducted by Zhang et al. (2006) and
application of a small quantity of spiking solution, prepared by Austad et al. (2005), an increase in oil recovery was observed
the addition of a small concentration of stearic acid, could assist along with an increase in temperature through spontaneous
in achieving clear inflection points. This method also led to a imbibition from chalk.3,54 This improvement was attributed to
much lower consumption of valuable crude oil.80 the higher affinity of SO42− ions to the carbonate surface at
With the importance of the acid number of crude oil toward higher temperatures, which results in the displacement of the
improving oil recovery, together with the ease of acid number negatively charged carboxylate groups (R-COO−) of the sulfate
measurements, more research has been performed to explore ions onto the positively charged carbonate surface, freeing the
this domain. In the work by Stadness and Austad (2000), it is oil from the rock surface, and altering the wettability of the rock
seen that the oil recovery was about 70% when the acid number to a more water-wet state.3,25 This relation between the
was 0 mg KOH/g, implying that the oil had no acidic polar improvement in oil recovery and the increased adsorption of
components to adsorb onto the rock surface, indicating a water- sulfate ions onto the carbonate surface at elevated temperatures
wet system. On the other hand, when the acid number has been further verified using the chromatographic wettability
increased up to 1.73 mg KOH/g, oil recovery decreased test developed by Strand et al. (2004), where an increase in the
dramatically.18 Similar observations were found in the work degree of water-wetness was simultaneously observed along
conducted by Fathi and co-workers (2010, 2010, 2011).21−23 with a decrease in the concentration of sulfate ions in the
Several spontaneous imbibition experiments were performed effluent at higher temperatures.84
on cores aged with oils of different acid number, where the acid In the absence of sulfate ions in the injected fluid, as
number of each crude oil was tailored by extracting the acidic discussed in section 3, an improvement in oil recovery due to
components from the oil. It was observed that the oil with the LSWF can still be observed in carbonates, but only in the
maximum quantity of acidic components showed the least presence of anhydrite minerals. However, the dissolution of
recovery, as well as the lowest water-wet fraction at the end of anhydrite is adversely affected by an increase in temperature,
the recovery process.22,79 This was also validated in the surface where higher temperature results in a lower dissolution
complexation model developed by Qiao et al. (2015), which rate.39,60 Additionally, the efficiency of wettability alteration
predicted a lower recovery potential by increasing the acid increases as temperature increases.62 Hence, a reduced
number of the crude oil.58 These studies suggest that the acid dissolution rate can be compensated by a higher wettability
number of the crude oil plays a critical role in determining the alteration efficiency, and vice versa. It is therefore critical to
incremental oil recovery and whether a reservoir could be understand what type of chemical reactions are taking place in
considered a good candidate for CTWF. the system based on the ionic composition of the brine and the
4.1. Open Question. The investigation of acid number rock mineralogy. Further, it is also important to see which of
emphasizes the importance of organic acids present in crude oil these reactions are more active, and at which particular
toward enhancing oil recovery. However, there is little temperature.
understanding about the origin of the acidic groups in the In the electrokinetic study conducted by Alotaibi et al.
crude oil, which could be attributed to the oxidation process (2011), they measured the zeta potential of the limestone and
that occurs during the developmental stages of hydrocarbon dolomite particles in different brine environments, and at
formation. It is understood that underground water plays a key different temperature values. They observed that in a seawater
role in transporting the crude oil to its producing zone in the environment at a pH of 7 at 25 °C, the zeta potential of
reservoir. It can be inferred that in the process of this migration limestone was about +6.8 mV, suggesting a positive charge of
and interaction with water molecules the hydrocarbons are the limestone surface. Further, when the temperature was
oxidized and begin to possess the acidic groups. However, little increased to 50 °C, they observed a decrease in zeta potential,
experimental work supports this argument. In addition, acid owing to the increased dissolution of Ca2+ ions from the
number measurement is not an accurate way of identifying the limestone surface due to the higher activity of these ions at the
amount of acidic components in oil that are reactive with, and elevated temperature. Similar results were obtained for
have an affinity to, the rock surface (i.e., acidic components dolomite particles at higher temperatures.52
with surface active species). It is rather a measure of the total The activity of cations in injected brine is highly affected by
acidic components in oil, both reactive and nonreactive. temperature. In the study conducted by Zhang et al. (2007),
According to Hoeiland et al. (2001), identifying the acid type adding Mg2+ ions resulted in a higher incremental oil recovery
G DOI: 10.1021/acs.energyfuels.7b01067
Energy Fuels XXXX, XXX, XXX−XXX
Energy & Fuels Review

than by adding Ca2+ ions at 100 °C.9 This was attributed to the 6. PROPOSED MECHANISMS FOR WETTABILITY
higher activity of Mg2+ ions as compared to Ca2+ ions at higher ALTERATION IN SANDSTONE RESERVOIRS
temperatures. This effect was more pronounced as the It has been well accepted that wettability alteration in oil-wet
temperature was further increased. The increase in the activity and mixed-wet systems is the primary means for IOR by
of Mg2+ observed at 130 °C overshadowed the effect of adding CTWF. Several authors have proposed different physicochem-
twice the amount of SO42− ions to an injected brine that had ical mechanisms to validate wettability alteration as the primary
Ca2+ ions instead of Mg2+ ions, where a larger incremental oil cause for IOR by CTWF. These include MIE,9,18,24 salting-
recovery (∼20% of OOIP) was observed,9 as shown in Figure in,10,89 fines migration,8 EDL expansion,50,90 mineral dissolu-
1. tion,7,31 pH modification,8,62 and in situ emulsification. Some of
these mechanisms were proposed for sandstone rocks and
others for carbonate rocks. Authors have also tried to establish
various similarities and differences among these mechanisms.89
There still exists a lack of consensus as to which mechanism is
more prevalent, and under what conditions. In this section, we
elaborate on the proposed mechanisms that are relevant toward
wettability alteration in sandstone reservoirs, while section 7
discusses mechanisms relevant toward wettability alteration in
carbonate reservoirs.
6.1. Wettability Alteration by “Salting-In”. Austad et al.
(2010) propose a low salinity mechanism in sandstone
reservoirs, where they suggest that clay particles act as cation
exchangers on the sandstone surface.38 Initially, in the natural
state of the reservoir, both basic and acidic materials are
adsorbed onto the clay surface, owing to the negative surface
charge of clay particles and their large specific surface area
Figure 1. Spontaneous imbibition of different brines into chalk at 70, compared to quartz. As shown in Figure 2, at initial conditions,
100, and 130 °C to study the effect of temperature on the activity of
ions affecting the process of oil recovery.9

Another factor where temperature plays a role in CTWF is


the dependence of the oil-wet characteristic of carbonate rocks
on temperature. This property is mostly attributed to the
adsorption of carboxylic materials onto the rock surface, due to
their high affinity to carbonate surfaces compared to other polar
components that naturally exist in crude oil.85 Hence,
decarboxylation of carboxylic materials can lead to the
wettability alteration of carbonate surfaces to a more water- Figure 2. Schematic representation of the “salting-in” mechanism of
wet state. Thermal decarboxylation at higher temperatures is wettability alteration in sandstones, showing the liberation of acidic
possible, especially on a carbonate surface, where CaCO3(s) and basic components attached to the sandstone surface upon low
salinity water injection (used with permission, Austad et al. 2010).38
acts as a catalyst for the decarboxylation reaction.86
5.1. Open Questions. In the simulation work by Khorsandi
et al. (2016), the authors modeled the effects of low salinity
waterflooding together with polymer flooding based on the clay particles are exposed to an environment of formation brine,
concepts of reactive transport, wettability alteration, reaction which has high salinity levels. The divalent ions in the rock
network, and multiphase fluid flow.87 They modeled the effects samples establish a chemical equilibrium with the surrounding
of preflushing with low salinity water because of its ability to high salinity formation brine. After the low salinity water is
prevent the degradation of polymer viscosity. The model injected, the chemical equilibrium is disturbed and the divalent
successfully accounted for the degradation in viscosity caused ions from the rock surface tend to free themselves to re-
by a high salinity brine and was validated against the establish the chemical equilibrium. Water molecules in the
experimental work performed by Shiran and Skauge (2013),88 neighborhood help facilitate this process by dissociating into
showing a good match.87 These mathematical models are H+ and OH− ions. H+ ions from the water molecules adsorb
successful in predicting oil recoveries; however, they do not onto the clay surface, due to having the highest affinity to the
consider all of the factors that may affect the process of clay surface,67 and OH− ions remain in solution, resulting in a
wettability alteration, and in particular, the importance of local increase in pH near the clay surface.38 These OH− ions
temperature in changing the activity of the ionic species that react with the acidic/basic material adsorbed onto the rock
affect recovery rates. surface, releasing them from the surface. This is described as
Further, there is limited understanding in the literature as to the “salting-in” mechanism by Rezaeidoust et al. (2009).89
why Ca2+ ions show higher activity at lower temperature ranges The following reactions describe the salting-in mechanism in
as compared to Mg2+ ions. Also, to the best of our knowledge, sandstones:
no studies have been performed to identify the temperature Clay − Ca 2 + + H 2O = Clay − H+ + Ca 2 + + OH−
threshold at which a shift in the activity of Ca2+ and Mg2+ ions
takes place, and whether this threshold changes with rock type. For the basic component binding to the rock surface
H DOI: 10.1021/acs.energyfuels.7b01067
Energy Fuels XXXX, XXX, XXX−XXX
Energy & Fuels Review

Clay − NHR+3 + OH− = Clay + R3N + H 2O negative zeta potential value means a thicker EDL, which is
indicative of a higher likelihood of water-wetting conditions.
For the acidic component binding to the rock surface Lee et al. (2010) reported that the water layer thickness of
Clay − RCOOH + OH− = Clay + RCOO− + H 2O the sandstone rock surface increased from 10.8 to 11.8 Å as
salinity was decreased from 0.1 to 0.001 M (6000 ppm to 60
6.2. Electric Double Layer (EDL) Expansion. An EDL, ppm) by changing the amount of NaCl dissolved in brine.
first proposed by Von Helmholtz, is a structure that appears on However, when decreasing the salinity of brine by altering the
portions of rock minerals that are exposed to fluids in the pores, concentration of MgCl2, the water layer thickness increased
as well as at rock−fluid and fluid−fluid interfaces. It refers to from 8.1 to 14.8 Å, therefore suggesting the importance of both
two parallel layers of charge. The first layer is referred to as the salinity and ion type toward the change in the thickness of the
surface charge, which comprises oppositely charged ions double layer.90
adsorbed onto the surface, whereas the second layer, known As lower salinity water is injected, reducing multivalent
as the dif f use layer, is composed of ions that are loosely cations, the EDL expands both at the oil/brine interface and at
attracted to the surface charge.34 EDL expansion is dependent the rock/brine interface. This expansion causes an overlap
on the electrical surface charge and is a function of the pH and between both EDLs, which in turn leads to an increase in the
salinity of the brine, as well as the cation type surrounding the repulsive forces between them. Once the repulsive forces
clay or sandstone particle. Zeta potential, as shown in Figure 3, exceed the binding forces involved in cationic bridging of acidic
groups in oil to the clay surface, the water layer between the oil
and clay surface expands, and oil particles desorb from the clay
surface, increasing water-wetness and improving oil recov-
ery.51,64 Also, decreasing the salinity of injected brine leads to
more negative zeta potentials at both the oil−brine and clay-
brine interfaces, which results in a stronger water-wet state and
IOR.51 However, the trend of increased production as salinity
decreases is not indefinite. If salinity is further lowered,
repulsive forces within clay minerals can exceed the binding
forces that keep clay particles intact. This results in
deflocculation of clay minerals and formation damage.91
Hence, an optimum injection brine salinity has to be identified
to avoid hindering oil flow and jeopardizing oil recovery.
6.3. Fines Migration. Early experiments conducted by
Bernard (1967) suggest that improved sweep efficiency can be
attributed to fines migration and clay swelling as a result of
freshwater injection.6 This was considered the main mechanism
for improving oil recovery and was further confirmed by the
study conducted by Tang and Morrow (1999).8 These authors
proposed an interplay between the mechanical capillary forces
that bind the oil particles to the fines and the viscous forces of
the low salinity waterflood. Further, the authors commented on
Figure 3. Schematic representation of the electric double layer (EDL)
the stability of the fines particle in the suspension, suggesting
around a charged particle in a brine suspension.50
that low-salinity water helps expand the electrical double
layer,50,90 which assists in the stripping of fines and increased
is the potential measured at the shear plane of the EDL and is oil recovery, as shown in Figure 4.8 This was also seen in the
useful toward the estimation of the double layer thickness. High coreflood experiments by Pu et al. (2010), where an increase in
pH, low salinity brines result in a thicker EDL.50 A more pressure drop was found to assist in oil recovery. The authors

Figure 4. Schematic representation of “fines migration” mechanism of wettability alteration in sandstones showing the liberation of fines attached to
the rock surface during low salinity waterflooding (recreated from Tang and Morrow (1999), with permission).8

I DOI: 10.1021/acs.energyfuels.7b01067
Energy Fuels XXXX, XXX, XXX−XXX
Energy & Fuels Review

Figure 5. A schematic representation of the “MIE” mechanism of wettability alteration in carbonate reservoirs showing the sequential process of
liberation of the crude oil particles from the carbonate surface by a cumulative participation of the PDIs. (a) Low temperature ranges below 70 °C.
(b) high temperature ranges of above 100 °C (adapted from Zhang et al. 2007).9

attributed this to the cumulative effect of anhydrite dissolution rock surface. This releases the organometallic complexes
and fines migration in the Tensleep reservoir in Wyoming.31 (RCOO-M, where M is the multivalent cation) from the
However, some other authors reported little to no generation rock.10 A similar mechanism for carbonate rocks was proposed
of fines or decrease in permeability changes during low salinity by Zhang et al. (2007).9 The cumulative interplay of the
brine injection and proposed for a mechanism based on the divalent ions from the brine, together with the rock and crude
specific interactions between the rock mineral and the organic oil, impacts the process of wettability alteration. The
components in the crude oil. This was termed as the MIE mechanism is a sequential process that can be understood as
mechanism and is described in the following section.10 follows: Owing to the positive surface charge of the carbonate
surface, SO42− ions are attracted toward the surface, lowering
7. PROPOSED MECHANISMS FOR WETTABILITY the surface potential and subsequently attracting the divalent
ALTERATION IN CARBONATE RESERVOIRS positive ions closer to the surface. As discussed earlier,
temperature plays an important role toward the activity of
7.1. Wettability Alteration by “Salting-In”. In the these ionic species. Hence, the mechanism is divided into two
experiments conducted by Yousef et al. (2012)49 with cases with respect to temperature. At lower temperature ranges
carbonate rocks and low salinity water injection, the authors of about 70 °C, the activity of Ca2+ ions is higher, and
concluded that wettability alteration takes place due to rock consequently, more Ca2+ ions are seen close to the surface than
dissolution. They proposed that a chemical equilibrium is set Mg2+ ions. This is reversed at temperatures above 100 °C,
up in the predisturbed state when the rock is exposed to high where Mg2+ ions show higher activity. During the interaction of
saline formation water. Once the system is exposed to a low the positive ions with the SO42− ions, the positive ions interact
salinity environment, the divalent ions (Ca2+, Mg2+) present in with the negative carboxylic end of the crude oil attached to the
the rocks tend to displace toward the brine to re-establish carbonate surface. More positive ions in the solution form a
equilibrium. During this process, the polar ends of the crude oil stronger interaction with the carboxylic end, as opposed to the
attached to the carbonate surface are liberated, in a way, attraction between crude oil and carbonate surface releasing the
increasing the solubility of crude oil in the surrounding brine, oil particles off the surface, resulting in IOR.9 The mechanism is
thereby improving oil recovery.49,89 This is similar to the summarized in the schematic shown in Figure 5.
“salting-in” mechanism described in the previous section for The following chemical reactions help explain the mecha-
sandstone reservoirs. Some authors describe this as “rock nism:89,92
dissolution” mechanism, which is a consequence of the process
of wettability alteration as the divalent ions migrate from the RCOO− − Ca − CaCO3(s) + Ca 2 + + SO24 −
rock surface toward the brine.
= RCOO − Ca 2 + + Ca − CaCO3(s) + SO24 −
7.2. Multi Ion Exchange (MIE). Lager et al. (2008)
proposed this mechanism based on the different interactions,
including, cation exchange, anion exchange, ligand exchange, RCOO− − Ca − CaCO3(s) + Mg 2 + + SO24 −
protonation, cation bridging, hydrogen bonding, van der Waals = RCOO − Ca 2 + + Mg − CaCO3(s) + SO24 −
interaction, or water bridging, that may happen between the
rock mineral and the various organic functional groups present At high temperatures, Mg2+ ions, due to their higher activity,
in the crude oil. Upon the injection of brine, rich in multivalent interact with the surface Ca2+ ions and replace them with a
ions, stronger ligand bonding between the carboxylic material process called dolomitization.9 These interactions release the
and the metal cation in the brine overcomes the weaker cation Ca2+ ions from the surface which react with the crude oil at the
bridging associations between the carboxylic material and the surface releasing the carboxylic material.
J DOI: 10.1021/acs.energyfuels.7b01067
Energy Fuels XXXX, XXX, XXX−XXX
Energy & Fuels Review

The simulation model developed by Qiao et al. (2015) is CEC = cation exchange capacity
based on this mechanism for altering wettability and for CTWF = chemically tuned waterflooding
predicting incremental oil recovery from mixed wet carbonate EDL = electric double layer
reservoirs.93 The model accounted for the effect of the surface EOR = enhanced oil recovery
active components in the crude oil, the surface charge of IOR = improved oil recovery
carbonate rock and temperature. It was successfully validated LSWF = low-salinity waterflooding
against the coreflood experiments performed by Fathi et al. MIE = multi-ion exchange
(2010).22 micro-CT = microcomputed tomography
OOIP = original oil in place
8. CONCLUDING REMARKS PDI = potential determining ion
CTWF is an expanding technique in EOR, with the particular
attribute of enabling wettability alteration of oil- and
intermediate-wet rock systems in order to assist in oil
■ REFERENCES
(1) History of Petroleum Engineering; American Petroleum Institute:
production. The success of a chemically tuned waterflood is Dallas, Texas, 1961.
highly determined by the interplay of many factors, such as (2) Bader, M. S. H. Desalination 2007, 208 (1−3), 159−168.
crude oil properties, brine salinity, brine composition, rock (3) Austad, T.; Strand, S.; Høgnesen, E. J.; Zhang, P. In SPE
mineral composition, and reservoir temperature. Proper International Symposium on Oilfield Chemistry; Houston, Texas, 2005;
integration of surface analytical techniques like zeta potential, pp 1−10.
acid number, X-ray diffraction, etc. can be useful for the (4) Høgnesen, E. J.; Strand, S.; Austad, T.; Stavanger, U. In SPE
characterization of rock and crude oil compositions for Europec/EAGE Annual Conference; Madrid, Spain, 2005; pp 1−9.
quantification of mineral content in the rock and acid/base (5) Martin, J. In 3rd Annual Meeting, Society of Petroleum Engineers;
content in the crude oil. These analyses can be key in assessing AIME; Caracas, Venezuela, 1959; pp 1−24.
the optimum chemical tuning of injected brine for the most (6) Bernard, G. G. In Annual California Regional Meeting of the Society
of Petroleum Engineers of AIME; California, Los Angeles, 1967; pp 1−8.
effective waterflood scheme. Reservoir temperature is also (7) Tang, G. Q.; Morrow, N. R. SPE Reservoir Eng. 1997, 12
recognized to be a critical factor impacting the activity of ionic (November), 269−276.
species as is evident from the “MIE” mechanism for wettability (8) Tang, G. Q.; Morrow, N. R. J. Pet. Sci. Eng. 1999, 24 (2−4), 99−
alteration in carbonate rocks. Several other mechanisms 111.
including “rock/mineral dissolution”, “fines migration”, “elec- (9) Zhang, P.; Tweheyo, M. T.; Austad, T. Colloids Surf., A 2007, 301
trical double layer expansion”, etc., have been proposed for (1−3), 199−208.
understanding wettability alteration in carbonate and sandstone (10) Lager, A.; Webb, K. J.; Black, C. J. J.; Singleton, M.; Sorbie, K. S.
reservoirs. These mechanisms display both similarities and Petrophysics 2008, 49 (1), 28−35.
disparities in the two rock types. For instance, low salinity water (11) Bagci, S.; Kok, M. V; Turksoy, U.; Eng, N. G.; East, M. In SPE
injection helps to disturb the chemical balance between the International Symposium on Oilfield Chemistry; Houston, Texas, 2001;
rock and the brine, leading to the release of oil particles in both pp 1−11.
(12) Farouq-Ali, S. M.; Stahl, C. D. Earth Miner. Sci. (United States)
carbonates and sandstones, with the exception of those
1970, 39, 4.
carbonate rocks devoid of anhydrite mineral. Although some (13) Treiber, L. E.; Owens, W. W. SPEJ, Soc. Pet. Eng. J. 1972, 6 (12),
studies indicate the importance of low salinity brine injection in 531−540.
carbonate reservoirs, most point toward the role of potential (14) Chilingar, G. V.; Yen, T. F. Energy Sources 1983, 7 (1), 67−75.
determining ions for improving oil recovery. A holistic (15) Anderson, W. G. JPT, J. Pet. Technol. 1986, 38 (10), 1125−1144.
understanding of these factors affecting wettability alteration, (16) Salathiel, R. A. JPT, J. Pet. Technol. 1973, 25 (10), 1216−1224.
and their interplay, can help improve the current understanding (17) Wagner, O. R.; Leach, R. O. Trans. Soc. Pet. Eng. 1959, 216 (1),
of the simultaneous chemical reactions that occur at the rock 65−72.
surface. These can include salt precipitation, mineral dissoci- (18) Standnes, D. C.; Austad, T. J. Pet. Sci. Eng. 2000, 28 (3), 111−
ation, crude oil desorption/solubilization, etc. Their appropriate 121.
identification according to the crude oil/rock/brine system can (19) Puntervold, T.; Strand, S.; Austad, T. Energy Fuels 2007, 21 (6),
lead to the development of more robust mathematical models 3425−3430.
(20) Strand, S.; Austad, T.; Puntervold, T.; Høgnesen, E. J.; Olsen,
that are better equipped to optimize waterflooding practices
M.; Barstad, S. M. F. Energy Fuels 2008, 22 (5), 3126−3133.
and oil recovery estimations.


(21) Fathi, S. J.; Austad, T.; Strand, S.; Puntervold, T. Energy Fuels
2010, 24 (5), 2974−2979.
AUTHOR INFORMATION (22) Fathi, S. J.; Austad, T.; Strand, S. Energy Fuels 2010, 24 (4),
Corresponding Author 2514−2519.
*Phone: (814) 863-2273. Fax: (814) 865-3248. E-mail: (23) Fathi, S. J.; Austad, T.; Strand, S. Energy Fuels 2011, 25 (11),
ZKarpyn@psu.edu. 5173−5179.
(24) Shariatpanahi, S. F.; Strand, S.; Austad, T. Energy Fuels 2010, 24
ORCID
(11), 5997−6008.
Zuleima T. Karpyn: 0000-0002-2418-7653 (25) Strand, S.; Høgnesen, E. J.; Austad, T. Colloids Surf., A 2006,
Author Contributions 275, 1−10.
# (26) Puntervold, T.; Strand, S.; Ellouz, R.; Austad, T. J. Pet. Sci. Eng.
P.P. and M.S.T. contributed equally.
Notes 2015, 133, 440−443.
(27) Zahid, A.; Stenby, E. H.; Shapiro, A. A. In EAGE Annual
The authors declare no competing financial interest.


Conference & Exhibition; Copenhagen, Denmark, 2012.
(28) Emadi, A.; Sohrabi, M. In SPE Annual Technical Conference and
ABBREVIATIONS Exhibition; New Orleans, Lousiana, USA, 2013; pp 1−15.
ASTM = American Society for Testing and Materials (29) Sheng, J. J. J. Pet. Sci. Eng. 2014, 120, 216−224.

K DOI: 10.1021/acs.energyfuels.7b01067
Energy Fuels XXXX, XXX, XXX−XXX
Energy & Fuels Review

(30) Mahani, H.; Keya, A. L.; Berg, S.; Bartels, W.-B.; Nasralla, R.; (63) Mahani, H.; Keya, A. L.; Berg, S.; Bartels, W. B.; Nasralla, R.;
Rossen, W. Europec 2015 2015, 1−27. Rossen, W. R. Energy Fuels 2015, 29 (3), 1352−1367.
(31) Pu, H.; Xie, X.; Yin, P.; Morrow, N. R. In SPE Annual Technical (64) Hilner, E.; Andersson, M. P.; Hassenkam, T.; Matthiesen, J.;
Conference and Exhibition; Florence, Italy, 2010; pp 1−8. Salino, P. A.; Stipp, S. L. Sci. Rep. 2015, 5, 9933.
(32) Romanuka, J.; Hofman, J. P.; Ligthelm, D. J.; Suijkerbuijk, B. M. (65) Wickramathilaka, S.; J. Howard, J.; R. Morrow, N.; Buckley, J. In
J. M.; Marcelis, a H. M.; Oedai, S.; Brussee, N. J. Eighteenth SPE IOR-2011−16th European Symposium on Improved Oil Recovery;
Improv. Oil Recover. Symp., 2012; pp 1−16. Cambridge, United Kingdom, 2011.
(33) Awolayo, A.; Sarma, H.; AlSumaiti, A. Transp. Porous Media (66) Rezaei-gomari, S.; Andrade, A. D.; Soltani, B. Int. J. Eng. Res.
2016, 111 (3), 649−668. Appl. 2015, 5 (7), 16−22.
(34) Zhang, Y.; Morrow, N. R. In SPE/DOE Symposium on Improved (67) Lake, L. W.; Johns, R. T.; Rossen, W. R.; Gary, P. A.
Oil Recovery; Oklahoma, USA, 2006; pp 1−15. Fundamentals of Enhanced Oil Recovery, 2nd ed.; Society of Petroleum
(35) Hamouda, A. A.; Valderhaug, O. M.; Munaev, R.; Stangeland, H. Engineers, 2014.
In SPE Improved Oil Recovery Symposium; Oklahoma, USA, 2014; pp (68) Grim, R. E. Clay Mineralogy; McGraw-Hill: New York, 1968.
1−13. (69) Cissokho, M.; Boussour, S.; Cordier, P.; Bertin, H.; Hamon, G.
(36) Sandengen, K.; Tweheyo, M. T.; Raphaug, M.; Kjølhamar, A.; Petrophysics 2009, 51 (5), 305−313.
Crescente, C.; Kippe, V. International Symposium of the Society of Core (70) Law, S.; McDonald, A.; Fellows, S.; Reed, J.; Sutcliffe, P. G. In
Analysts, 2011; pp 112. SPE Offshore Europe Conference and Exhibition; Aberdeen, Scotland,
(37) Puntervold, T.; Strand, S.; Austad, T. Energy Fuels 2009, 23 (5), UK, 2015; pp 1−11.
2527−2536. (71) Lebedeva, E. V; Fogden, A. Energy Fuels 2011, 25, 5683−5694.
(38) Austad, T.; Rezaeidoust, A.; Puntervold, T. In SPE Improved Oil (72) Vdović, N.; Bišcá n, J. Colloids Surf., A 1998, 137, 7−14.
Recovery Symposium; Tulsa, Oklahoma, USA, 2010; pp 1−17. (73) Karimi, M.; Al-Maamari, R. S.; Ayatollahi, S.; Mehranbod, N.
(39) Shariatpanahi, S. F.; Strand, S.; Austad, T. Energy Fuels 2011, 25 Colloids Surf., A 2015, 482, 403−415.
(7), 3021−3028. (74) Dandekar, A. Y. Petroleum Reservoir Rock and Fluid Properties,
(40) Yousef, A. A.; Al-Saleh, S.; Al-Kaabi, A.; Al-Jawfi, M. SPE Reserv. 2nd ed.; CRC Press: Boca Raton, FL, 2010.
Eval. Eng. 2011, 14 (5), 578−593. (75) James, M.; Hunter, R.; O’Brien, R. Langmuir 1992, 8 (8), 420−
(41) Gupta, R.; Smith, P. G., Jr.; Hu, L.; Willingham, T. W.; Lo 423.
Cascio, M.; Shyeh, J. J.; Harris, C. R. In SPE Middle East Oil and Gas (76) Nakatuka, Y.; Yoshida, H.; Fukui, K.; Matuzawa, M. Adv. Powder
Show and Conference; Manama, Bahrain, 2011; pp 1−21. Technol. 2015, 26 (2), 650−656.
(42) Austad, T.; Shariatpanahi, S. F.; Strand, S.; Aksulu, H.; (77) Denekas, M. O.; Mattax, C. C.; Davis, G. T. Pet. Trans. AIME
Puntervold, T. Energy Fuels 2015, 29 (11), 6903−6911. 1960, 216, 330−333.
(43) Shariatpanahi, S. F.; Hopkins, P.; Aksulu, H.; Strand, S.; (78) Seifert, W. K.; Howells, W. G. Anal. Chem. 1969, 41 (4), 554−
Puntervold, T.; Austad, T. Energy Fuels 2016, 30 (1), 180−187. 562.
(44) Agbalaka, C. C.; Dandekar, A. Y.; Patil, S. L.; Khataniar, S.; (79) Fathi, S. J.; Austad, T.; Strand, S. Energy Fuels 2011, 25 (6),
Hemsath, J. R. Transp. Porous Media 2009, 76 (1), 77−94. 2587−2592.
(45) Tang, G.; Morrow, N. R. Soc. Core Anal. 1999, 1−12. (80) Fan, T.; Buckley, J. SPE J. 2007, 12 (4), 22−26.
(46) Shehata, A. M.; Nasr-el-din, H. A.; Texas, A. SPE Reserv. Eval. (81) ASTM Standard Test Method D664-11 ; American Society for
Eng. 2016, 59−76. Testing And Materials, 2011.
(47) Yousef, A. A.; Al-Saleh, S.; Al-Jawfi, M. In SPE EUROPEC/ (82) ASTM Standard Test Method D2896-15; American Society for
EAGE Annual Conference and Exhibition; Vienna, Austria, 2011; Vol. 4, Testing And Materials, 2015.
pp 2814−2830. (83) Hoeiland, S.; Barth, T.; Blokhus, A. M.; Skauge, A. J. Pet. Sci.
(48) Yousef, A. A.; Al-Saleh, S.; Al-Kaabi, A. U.; Al-Jawfi, M. S. In Eng. 2001, 30 (2), 91−103.
Canadian Unconventional Resources & International Petroleum Confer- (84) Strand, S.; Standnes, D. C.; Austad, T. J. Pet. Sci. Eng. 2006, 52,
ence; Calgary, Canada, 2010; pp 1−35. 187−197.
(49) Yousef, A. A.; Al-Saleh, S.; Al-Jawfi, M. In Eighteenth SPE (85) Thomas, M. M.; Clouse, J. A.; Longo, J. M. Chem. Geol. 1993,
Improved Oil Recovery Symposium; Tulsa, Oklahoma, USA, 2012; pp 109, 201−213.
1−18. (86) Shimoyama, A.; Johns, W. D. Geochim. Cosmochim. Acta 1972,
(50) Nasralla, R. a; Nasr-el-din, H. a. SPE Reserv. Eval. Eng. 2012, 36 (1), 87−91.
49−59. (87) Khorsandi, S.; Qiao, C.; Johns, R. T. SPE J. 2016, No. 2014, 1−
(51) Myint, P. C.; Firoozabadi, A. Curr. Opin. Colloid Interface Sci. 14.
(88) Shiran, B. S.; Skauge, A. Energy Fuels 2013, 27, 1223−1235.
2015, 20 (2), 105−114.
(89) Rezaeidoust, A.; Puntervold, T.; Strand, S.; Austad, T. Energy
(52) Alotaibi, M.; Nasralla, R.; Nasr-El-Din, H. SPE Reserv. Eval. Eng.
Fuels 2009, 23 (9), 4479−4485.
2011, 14 (6), 3−6.
(90) Lee, S. Y.; Webb, K. J.; Collins, I. R.; Lager, A.; Clarke, S. M.;
(53) Khishvand, M.; Alizadeh, A. H.; Kohshour, I. O.; Piri, M. 2016,
Sullivan, M. O.; Routh, A. F.; Wang, X. In SPE Improved Oil Recovery
No. August, 1−12.
Symposium; Tulsa, Oklahoma, USA, 2010; pp 1−11.
(54) Zhang, P.; Austad, T. Colloids Surf., A 2006, 279, 179−187.
(91) Ligthelm, D. J.; Gronsveld, J.; Hofman, J. P.; Brussee, N. J.;
(55) Strand, S.; Standnes, D. C.; Austad, T. Energy Fuels 2003, 17
Marcelis, F.; Van Der Linde, H. a. In SPE EUROPEC/EAGE Annual
(5), 1133−1144.
Conference and Exhibition; Amsterdam, The Netherlands, 2009; pp 1−
(56) Fetter, C. W. Applied Hydrogeology; Prentice Hall: Upper Saddle
22.
River, NJ, 1988.
(92) Al-shalabi, E. W.; Sepehrnoori, K. J. Pet. Sci. Eng. 2016, 139,
(57) Pierre, A.; Lamarche, J. M.; Mercier, R.; Foissy, A.; Persello, J. J.
137−161.
Dispersion Sci. Technol. 1990, 11 (6), 611−635. (93) Qiao, C.; Li, L.; Johns, R. T.; Xu, J. SPE J. 2015, 20, 767−783.
(58) Qiao, C.; Johns, R.; Li, L. Energy Fuels 2016, 30 (2), 884−895.
(59) Filoco, P.; Sharma, M. SPE J. 1998, 5 (3), 293−300.
(60) Austad, T. Enhanc. Oil Recover. F. Cases 2013, 301−332.
(61) Pourhaji, P.; Arab, Z.; Foroughizadeh, P. In The International
Conference on New Research in Chemistry and Chemical engineering;
2015; pp 1−15.
(62) Austad, T.; Shariatpanahi, S. F.; Strand, S.; Black, C. J. J.; Webb,
K. J. Energy Fuels 2012, 26, 569−575.

L DOI: 10.1021/acs.energyfuels.7b01067
Energy Fuels XXXX, XXX, XXX−XXX

You might also like