You are on page 1of 11

Journal of Colloid and Interface Science 563 (2020) 145–155

Contents lists available at ScienceDirect

Journal of Colloid and Interface Science


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

Effect of salinity, Mg2+ and SO2


4 on ‘‘smart water”-induced carbonate
wettability alteration in a model oil system
Jin Song a, Qing Wang a, Imran Shaik b, Maura Puerto a, Prem Bikkina b, Clint Aichele b, Sibani L. Biswal a,⇑,
George J. Hirasaki a,⇑
a
Rice University, 6100 Main St., MS-362, Department of Chemical and Biomolecular Engineering, Houston, TX 77005, USA
b
Oklahoma State University, 420 Engineering North, School of Chemical Engineering, Stillwater, OK 74078, USA

g r a p h i c a l a b s t r a c t

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

Article history: Hypothesis: We present a systematic study of the ‘‘smart water” induced wettability alteration. This pro-
Received 7 November 2019 cess is believed to be greatly affected by the brine salinity and the presence of Mg2+ and SO2
4 in the brine.
Revised 10 December 2019 Experiments and modelling: To characterize the wettability alteration, we perform spontaneous imbibition
Accepted 11 December 2019
measurement using Indiana limestone cores and a model oil with added naphthenic acid. Both single-
Available online 12 December 2019
electrolyte-based and seawater-based ‘‘smart water” are tested to investigate the effect of Mg2+, SO2 4
and salinity on wettability alteration. Rock/brine and oil/brine zeta potentials are measured, and the elec-
Keywords:
trostatic component of disjoining pressure is calculated to understand the role of electrostatics in the
Wettability alteration
Smart water
wettability alteration. The surface concentration of charged species on the limestone surface is analyzed
Low salinity water based on a natural carbonate surface complexation model (SCM).

⇑ Corresponding authors.
E-mail addresses: js110@rice.edu (J. Song), qw10@rice.edu (Q. Wang), imran.shaik@okstate.edu (I. Shaik), Maura@rice.edu (M. Puerto), prem.bikkina@okstate.edu
(P. Bikkina), clint.aichele@okstate.edu (C. Aichele), biswal@rice.edu (S.L. Biswal), gjh@rice.edu (G.J. Hirasaki).

https://doi.org/10.1016/j.jcis.2019.12.040
0021-9797/Ó 2019 Elsevier Inc. All rights reserved.
146 J. Song et al. / Journal of Colloid and Interface Science 563 (2020) 145–155

Enhanced oil recovery Findings: Both the reduction of Na+ and addition of SO2
4 are found to contribute to wettability alteration.
Spontaneous imbibition Mg2+ is found to be unfavorable for wettability alteration. Ca2+ is believed to facilitate SO2
4 with wetta-
bility alteration based on the comparison between the single-electrolyte-based and seawater-based
brines. The reduction of the Na+ surface complexation (>CaOH  Na+0.25) in low salinity brines is believed
to be a critical mechanism responsible for wettability alteration based on the SCM calculations.
Ó 2019 Elsevier Inc. All rights reserved.

1. Introduction Yousef et al. [19] performed core flooding experiments with a


carbonate reservoir core and crude oil. They observed approxi-
Wettability is known to have a significant impact on the oil mately an additional 20% oil recovery of original oil in place (OOIP)
recovery in carbonate reservoirs. About 90% of carbonate reservoirs by diluted seawater injection. Al-Harrasi et al. [20] examined the
are neutral to preferentially oil-wet [1], which makes the oil recov- effect of low salinity brine (diluted seawater) by both spontaneous
ery very challenging due to the unfavorable capillary pressure dur- imbibition and core flood testing using the same core and crude oil.
ing brine injection. Injection of hydrophilic surfactants as the Significant additional oil recovery was observed in the sponta-
wettability modifier can successfully alter wettability [2–9] but neous imbibition test, but much less increment was observed in
increases the capital investment significantly. Another approach the core flooding. Mohanty and Chandrasekhar [21] studied the
that requires much lower cost is the injection of ‘‘smart water”, effect of modified seawater by core flooding and demonstrated that
which are the brines with designed salinity and ionic composition Mg2+ and SO2 4 have to be present for the modified seawater to
that can enhance oil recovery. ‘‘Smart water” has drawn increasing improve oil recovery. Gandomkar and Rahimpour [22] confirmed
attention from both the petroleum industry and the academia over that diluted seawater could result in higher oil recovery compared
the past two decades [10–12]. This method requires relatively low to seawater in the core flooding experiment with limestone cores
capital investment compared to other chemical enhanced oil and crude oils.
recovery (EOR) techniques and has been reported to successfully Besides core-level experiments, other efforts have been made to
increase oil recovery from lab scale to field scale [11]. However, decipher ‘‘smart water” EOR from different perspectives. Surface
there are many cases where ‘‘smart water” has not been successful complexation models have been developed to understand the sur-
[11,13,14]. Here we aim to understand the fundamental physico- face interactions between carbonate and the oil in different brines
chemical interactions governing ‘‘smart water” EOR. [12,23–25]. Disjoining pressure calculations have been performed
Significant research utilizing the core-scale experiments with to reveal the relationship between the intermolecular interactions
smart water have been reported in the literature. Strand et al. [1] and the wetting behavior [10,26–28]. Molecular dynamics simula-
and Zhang et al. [15] performed spontaneous imbibition tests with tion has also been applied to study the mechanism of Mg2+ and
chalk mineral and crude oil at several temperatures up to 130 °C SO2
4 induced wettability alteration by Sakuma et al. [29] and
and observed up to 40% additional oil recovery after introducing Andersson et al. [30] Multiple surface characterizations including
Ca2+, Mg2+, and SO2 4 into the modified seawater. With the assis- surface forces measurement have been performed to understand
tance of zeta potential measurement and effluent ionic composi- the change of calcite surface after being exposed to low salinity bri-
tion analysis, they proposed a mechanism based on electrostatic nes by Chen et al. [31] Moreover, contact angles of carbonate/oil/
interactions to explain the roles of Ca2+, Mg2+, and SO2 4 . In their brine system are utilized to characterize the rock wettability
hypothesis, SO24 favors the wettability alteration by lowering the before and after exposing to the ‘‘smart water” [21,32–34]. How-
positive surface charge of carbonate surface, thereby weakening ever, the consistency of contact angle data among different studies
the electrostatic attraction between the negatively charged crude is typically unsatisfactory.
oil and the carbonate surface. They also proposed that Ca2+ could Even though a large amount of work has been done in the crude
form a positively charged surface complex with the adsorbed oil system using modified seawater as the ‘‘smart water”, the
organic acids and result in acid desorption due to charge repulsion underlying mechanisms governing the oil-brine-rock interfaces
with the rock surface. Mg2+ was believed to substitute Ca2+ in the are still not conclusive. Here we utilize a model oil composed of
lattice at high temperature and take away the adsorbed organic naphthenic acids, one of the surface-active components in crude
acids with the released Ca2+. Fathi et al. [16] performed similar oil, to better understand these mechanisms. A model oil study
spontaneous imbibition tests using treated crude oil depleted in can provide valuable insights because most of the proposed mech-
water-soluble acids and treated crude oil with only water-soluble anisms for ‘‘smart water” are based on the interaction between
acids. They concluded that the water-soluble acids in the crude ions and naphthenic acids in the oil.
oil have dramatic impact on both the initial wetting state of the Moreover, diluted seawater is investigated as the ‘‘smart water”
rock as well as the effectiveness of low salinity water. In another in most of the literature work. Dilution of seawater lowers the
study, Fathi et al. [17] demonstrated that lowering salinity (deplet- salinity and the concentration of divalent ions (Mg2+ and SO2 4 )
ing NaCl in seawater) can alter wettability and improve oil recov- simultaneously, making it difficult to distinguish the individual
ery in spontaneous imbibition. Austad et al. [13,14] emphasized effects of lowering ionic strength and modifying divalent ions con-
the possible effect of anhydrite presence in the core in determining centration. Therefore, a systematic study of the roles of the brine
the effectiveness of low salinity water by core flooding experi- composition (potential determining ions and salinity) in well-
ments. They observed additional oil recovery in cores with initial characterized model systems is required and currently unavailable
anhydrite presence when switching formation brine flooding to in the literature. In this work, Indiana limestone cores and model
diluted seawater flooding, while insignificant oil recovery was oil with added naphthenic acid are tested by different series of
observed in clean cores without anhydrite presence. Shariatpanahi low salinity brines in spontaneous imbibition. Two groups of
et al. [18] examined the ability of seawater-based ‘‘smart water” to ‘‘smart water” are tested: one group is prepared by modifying 4x
alter wettability in dolomite cores. They found that the seawater diluted seawater; the other group is prepared with single elec-
could not significantly increase oil recovery while the 10-times- trolyte (NaCl/Na2SO4/MgSO4). Zeta potential, disjoining pressure
diluted seawater was a successful EOR fluid in a dolomite system. and surface complexation modeling are also used to assist in
J. Song et al. / Journal of Colloid and Interface Science 563 (2020) 145–155 147

analyzing the effects of salinity and divalent ions, Mg2+ and SO2
4 , Single-electrolyte brines are also tested. This series of brines were
on wettability alteration. chosen to examine the effect of the change in the electrostatic
double-layer and correlate the electrostatic interactions to individ-
2. Experimental section ual potential determining ions.
All the brines were pre-equilibrated with calcite rock powder
2.1. Core material and AN1.5 MO (vol/vol = 1:1) for a week at ambient conditions
until the solution pH remained constant. During the pre-
Indiana limestone was chosen as the representative of carbon- equilibration process, the rock was not in direct contact with the
ate minerals. Seven Indiana limestone cores (Kocurek Industries) oil phase; therefore, the naphthenic acid in the oil would not be
of 1-inch length, 1-inch diameter, were selected for the sponta- depleted by adsorption onto the rock minerals. The purpose of
neous imbibition tests. The permeability was approximately pre-equilibration was to avoid rock dissolution and acid desorption
50mD. The mineralogy of the limestone was characterized by a from the rock surface to the brine. Moreover, the pH was constant
JEOL JXA 8530F Hyperprobe electron probe microanalyzer (EPMA) during the imbibition test if the brine was pre-equilibrated. All
with energy dispersive X-ray spectroscopy (EDX). The composition brine samples were prepared without air contact and equilibration
is shown in the supplementary material Table S1. The Indiana was done in a closed vessel.
limestone is primary CaCO3 (93% by molar fraction). Anhydrite Due to the dissolution of calcite, Ca2+ was present in all the
presence in the rock sample is expected to be negligible. This equilibrated brines even though no calcium salt is added in most
was confirmed by equilibrating 20 mL 1 wt% NaCl solution with of the brines. Ca2+ is one of the most important potential determin-
0.5 g crushed Indiana limestone powder for a week and then test- ing ions in a carbonate mineral system. Therefore, the Ca concen-
ing the supernatant by inductively coupled plasma atomic emis- tration was measured by ICP-OES to monitor possible variation of
sion spectroscopy (ICP-OES). Sulfate was found to be below the Ca level after switching to a different brine. The equilibrated brines
detection limit in the brine. are filtered by 0.22 mm PTFE filters and diluted in 1% nitric acid for
the ICP-OES measurement.
2.2. Model oils
2.4. Core preparation
A model oil composed of 87% (vol%) dodecane (Sigma-Aldrich,
99%) and 13% (vol%) toluene (Sigma-Aldrich, 99%), with added
The Indiana limestone cores were first dried to constant weight
naphthenic acid (Sigma-Aldrich, AN = 230) was used to model
in an oven. Then the cores were vacuumed and saturated with the
the crude oil. The dodecane and toluene represents the long chain
initial brine (5 M NaCl) at room temperature. The porosity and pore
saturates and aromatic compounds, respectively. The volume ratio
volume were calculated based on the change of core weight and
of dodecane/toluene was chosen based on lab experience to mimic
the density of the initial brine (1.1805 g/cm3 for 5 M NaCl). After-
the phase behavior of a crude oil. Naphthenic acid was added to
wards, the cores were placed in glass centrifuge tubes and
obtain a model oil with total acid number (TAN) = 1.5 mg KOH/g.
immersed in the oil for water drainage using a high-speed ultra-
The model oil is abbreviated as ‘‘AN1.5MO”. TAN is measured by
centrifuge at 11,000 rpm for at least 24 h. The equivalent capillary
a modified ASTM titration method [35] proposed by Fan and Buck-
pressure was calculated to be 80psi. The displaced brine volume
ley and confirmed to be 1.5 mg KOH/g oil. Moreover, model oil
was measured, and the initial oil saturation was calculated. After
(‘‘MO”) without any added acid was also tested as the control oil.
the oil saturation, the cores were aged for 2 weeks at 120 °C to
obtain oil-wet surfaces.
2.3. Brines

A 5 M NaCl solution was chosen as the representative formation 2.5. Procedure for spontaneous imbibition tests
brine. This ionic strength (I) is characteristic of carbonate reser-
voirs and is used as the initial brine to saturate the cores before The oil-wet cores were first placed into dry Amott imbibition
oil saturation. Neither Ca2+ nor Mg2+ is added to the formation cells. Then the cells were filled with the initial brine (5 M NaCl)
brine so that the individual role of the divalent ions and salinity and transferred into a water bath at 90 °C. The top narrow neck
can be investigated separately in the diluted brines. A series of of each Amott cell was not immersed in the water bath and was
low salinity brines are prepared, as shown in Table 1. Diluted elongated by connecting it to a glass pipette of the same OD as
seawater-based brines include the following: 4x diluted seawater the top of the cell. This extension functioned as a condenser. No
(‘‘4dSW”), 4x diluted seawater with 3x concentrated SO2 4 noticeable oil loss due to evaporation was observed during the
(‘‘4dSW3S”), 4x diluted seawater depleted in SO2 4 (‘‘4dSW0S”), experiment. The recovered oil was read daily through the gradu-
and 4x diluted seawater depleted in Mg2+ (‘‘4dSW0Mg”). ated section of the Amott cell.

Table 1
Recipe of the tested brines (before equilibration) and their equilibrium pH.

Name Ionic strength(M) Na+(mM) Ca2+(mM) Mg2+(mM) Cl(mM) HCO


3 (mM) SO2
4 (mM) pHa Group
5 M NaCl 5.000 5000 0 0 5000 0 0 8.0 Initial brine

4dSW 0.164 115.0 3.25 11.25 131.5 0.5 6 7.5 Seawater based
4dSW3S 102.5 3.25 11.25 95.0 0.5 18 7.4
4dSW0S 121.0 3.25 11.25 149.5 0.5 0 7.4
4dSW0Mg 148.8 3.25 0 142.8 0.5 6 7.6

LowSal NaCl 164 0 0 164 0 0 7.8 Single electrolyte


LowSal Na2SO4 109 0 0 0 0 55 7.7
LowSal MgSO4 0 0 41 0 0 41 7.7
a
The brine pH was measured after the brine was equilibrated with calcite and AN1.5MO (1:1 vol/vol). The same equilibrated brines were used in the imbibition test.
148 J. Song et al. / Journal of Colloid and Interface Science 563 (2020) 145–155

The initial brine was switched to the designed low salinity bri- sonication, the large particles were allowed to settle for 1 h. Then
nes after the oil reading was stable for at least 5 days. Nine cores the supernatant was extracted for the measurement. Five samples
(A ~ I) were tested with different brine series. The porosity (/), ini- from each bottle was measured to obtain the mean value and stan-
tial oil saturation (Soi) and the testing brine sequence were sum- dard deviation.
marized in Table 2. AN1.5 model oil (25 mL) and equilibrated brine (25 mL) were
mixed together to prepare the dispersion for oil/brine zeta poten-
2.6. Amott-Harvey wettability index (IA-H) measurement tial measurement. A sonifier (Branson) was used to sonicate the
sample for 30 s to facilitate oil dispersion. After sonication, a period
To evaluate the initial wettability of Indiana limestone core and of 1 h was given to allow the oil droplet density to decrease to a
its wettability after oil aging in AN1.5MO, the Amott-Harvey wet- proper level for the zeta potential measurements. The supernatant
tability index (IA-H) was measured for two Indiana limestone cores was extracted, and each sample was measured 5 times.
(J and K) using 5 M NaCl brine. Core J was saturated with MO with-
out any oil aging step. Core K was saturated with AN1.5MO and 2.8. Surface complexation modeling
then aged in AN1.5MO for 2 weeks at 120 °C, the same as core A ~ I.
There were four sequential steps for IA-H measurement: (1) To probe the concentration variation of surface charged species
brine spontaneous imbibition; (2) brine forced imbibition; (3) oil in different brines, the carbonate surface complexation model
spontaneous drainage; and (4) oil forced drainage. IA-H was calcu- (SCM) previously proposed by Song, et al.[36] was adopted to
lated based on the definition: model the Indiana limestone rock. Based on the model assump-
V osp V wsp tions, the surface charge of limestone is a result of ion complexa-
IAH ¼  tion with the two major surface sites: >CO3H+0.75 and >CaOH0.75.
V osp þ V of V wsp þ V wf
Surface coverage of organic impurities (>CO3HA5.25), as well as
Vosp and Vof were the volume of displaced oil in the (1) and (2) the presence of silica impurity (>SiOH) are both considered to
steps, respectively. Vwsp and Vwf were the volume of displaced brine account for the surface charge difference between Indiana lime-
in the (3) and (4) steps, respectively. All steps were done at room stone and ideal pure calcite. Model parameters from this previous
temperature and the forced imbibition/drainage are achieved by work were used directly without re-fitting (shown in Table 3). A
centrifuging the core at 10,000 rpm for 48 h in total (after first geochemical database, PHREEQC Version 3.3.9 (database phreeqc.-
24 h, the core was flipped to eliminate capillary end effect). The dat), was used for the model calculation. The modeling procedure
tested oil was MO for core J and was AN1.5MO for core K. can be found in the reference by Song, et al. [36]. For the experi-
ments, the zeta potentials were measured in brines pre-
2.7. Zeta potential measurement equilibrated with the acidic AN1.5MO in order to represent the
actual brine chemistry in the spontaneous imbibition experiments.
To analyze the role of electrostatic interactions in determining Therefore, the solubility of naphthenic acid in the brines must be
the rock wettability, both oil/brine and rock/brine zeta potential considered in the SCM. The dissociation constant (pKa) for naph-
were measured in various brines. The brines used to prepare the thenic acids can vary depending on their origins. Previous litera-
rock/brine and oil/brine dispersion were first allowed to equili- ture report pKa between 3.9 and 5.0 for naphthenic acids based
brate with the AN1.5MO and the Indiana limestone rock (3 phases on titration curves [37,38]. Therefore, a model naphthenic acid
equilibration) for 1 week until the pH of the brine was stable. The whose pKa = 4.8 was assumed in the calculation to account for
container for the sample can be treated as a closed system because the impact of acid solubility on the pH of the equilibrated brines.
the brine is filled up to the brim, leaving negligible air space in the Small quantities of naphthenic acid were added to the brine in
bottle. Then the brine was extracted and filtered through a 0.5-mm the model to ensure that the brine pH after equilibration was equal
stainless steel filter. The same equilibrated brine was used for to the measured pH (7.7–8.0). Based on the calculations,
preparing both rock/brine and oil/brine dispersion. A cartoon 0.8–3.3 mM of the model naphthenic acid was needed to account
demonstrating the equilibration method is provided in the supple- for the naphthenic acid partitioning for a given brine. Note that
mentary material Fig. S1. The zeta potential was measured by a the calculation is not sensitive to pKa of naphthenic acid as long
light scattering analyzer (DelsaMax PRO) based on electrophoretic as pKa < 7.7 because the added acids will almost completely disso-
phase analysis light scattering (PALS) method. ciate in the tested brines. The calculated zeta potential and calcium
Ground Indiana limestone powder (0.2 g) and equilibrated brine concentration from SCM were compared to the actual measure-
(25 mL) were mixed to obtain 0.8%wt. sample for rock/brine zeta ments to check the validity of model. Finally, the concentrations
potential measurements. A sonifier (Branson) was used to sonicate of surface charged species were analyzed to provide insights on
the sample for 1 min to facilitate powder dispersion. After wettability alteration.

Table 2
Summary of experimental design of the spontaneous imbibition tests.

# /(%) Soi (%) Initial brine Oil Testing brine sequence Purpose
A 16 81 5 M NaCl AN1.5 MO 5 M NaCl ? 4dSW ? 4dSW3S Effect of Mg2+and SO42 (seawater based)
B 17 83 5 M NaCl AN1.5 MO 5 M NaCl ? 4dSW0S
C 18 85 5 M NaCl AN1.5 MO 5 M NaCl ? 4dSW0Mg

D 16 78 5 M NaCl AN1.5 MO 5 M NaCl ? LowSal NaCl ? LowSal Na2SO4 Effect of Na+, Mg2+and SO2
4 (single electrolyte)
E 17 81 5 M NaCl AN1.5 MO 5 M NaCl ? LowSal Na2SO4
F 17 86 5 M NaCl AN1.5 MO 5 M NaCl ? LowSal MgSO4

G 17 79 5 M NaCl AN1.5 MO 5 M NaCl ? LowSal NaCl Reproducibility (repeat)


H 17 78 5 M NaCl AN1.5 MO 5 M NaCl ? I = 0.4 M NaCl ? LowSal NaCl
I 17 85 5 M NaCl AN1.5 MO 5 M NaCl ? LowSal MgSO4

J 17 83 5 M NaCl MO 5 M NaCl Amott-Harvey wettability index


K 16 81 5 M NaCl AN1.5 MO 5 M NaCl
J. Song et al. / Journal of Colloid and Interface Science 563 (2020) 145–155 149

Table 3 additional oil is recovered in 4dSW3S in the tertiary recovery


Model parameters for the surface complexation model of the Indiana limestone from mode. From the comparison, the presence of SO2 is confirmed
4
Song et al. [36].
to be beneficial for oil recovery while Mg2+ is not.
Surface reaction Log10K Surface composition Spontaneous imbibition using single-electrolyte solutions was
>CaOH0.75 + H+
> CaOH+0.25
2 0.30 CaCO3 (96%) also tested to further compare the individual effect of Mg2+, SO2 4
>CO3H+0.75 + OH
> CO0.25
3 + H2O 0.50 >CaOH0.75 4.95 nm2 and reduction in ionic strength (Fig. 1(b)). Low salinity NaCl brine
0.75
>CaOH + Ca
> CaOH  Ca+1.25
2+
1.98 >CO3H+0.75 4.80 nm2 without any added divalent ions is found to also trigger wettability
0.75
>CaOH + Mg
> CaOH  Mg+1.25
2+
1.66 >CO3HA5.25 0.15 nm2
>CO3H+0.75 + CO2 1.25 alteration and results in 17% additional oil recovery. Higher oil
3
> CO3H  CO3 1.99 SiO2 (4%)
>CO3H+0.75 + SO2
4
> CO3H  SO4
1.25
1.00 >SiOH 2.3 nm2 recovery is observed in the two sulfate-containing brines: 25% in
>CO3H+0.75 + HCO 3
> CO3H  HCO3
0.25
0.25 LowSal Na2SO4 and 21% in LowSal MgSO4. Additionally, LowSal
>CaOH0.75 + Na+
> CaOH  Na+0.25 0.09 Na2SO4 is also tested in a tertiary mode following the LowSal NaCl
>CO3H+0.75 + Cl
> CO3H  Cl0.25 0.50
brine for Core D. The final oil recovery in LowSal Na2SO4 is almost
>SiOH
> SiO + H+ 4.0 [39]
>SiOH + Ca2+
> SiOCa+ + H+ 9.7 [40] identical as that for Core E, which is tested in LowSal Na2SO4 in sec-
ondary mode. This comparison shows the good reproducibility of
the imbibition test.
Reproducibility of the oil recovery in LowSal NaCl and LowSal
3. Results and discussion
MgSO4 is examined by duplicating the experiments. 4dSW is also
re-tested, but in tertiary mode, following LowSal MgSO4 brine.
3.1. Wettability after oil aging
Good reproducibility is observed for all the repeated cases. The fig-
ures of the oil recovery history of the repeated cases are provided
The wettability index (IA-H) of Indiana limestone cores were
in the supplementary material. A summary of all the final oil recov-
measured to investigate the initial wettability and the wettability
ery results is provided in Fig. 2. The effectiveness of low salinity
after AN1.5MO oil aging. The wettability index IA-H is 0.4 for the
‘‘smart water” with the same ionic strength ranks as: 4dSW3S 
core J, which was saturated with the non-surface-active MO and
4dSW0Mg > LowSal Na2SO4  4dSW > LowSal MgSO4 > 4dSW0S
tested without oil aging. IA-H is positive when the brine can more
 LowSal NaCl. For both seawater-based brines and single-
favorably displace the oil comparing to the oil displacing the brine
V V
electrolyte solutions, SO24 is confirmed to be beneficial and Mg
2+

(V osposp
þV
> V wspwsp
þV
). Therefore, a positive IA-H indicates a water-wet is unfavorable for wettability alteration at 90 °C.
of wf

state of the Indiana limestone core before oil aging. However, IA- A similar benefit of SO24 has also been observed in the sponta-
H is 0.1 for the core K, which was saturated with the acidic neous imbibition test with crude oils [1,15]. Strand et al. [1] pro-
AN1.5MO and tested after oil aging at 120 °C for 2 weeks. The neg- posed a two-step mechanism explaining the effect of SO2 4 on
ative IA-H shows the limestone core becomes preferentially oil-wet wettability alteration: (1) SO2 4 adsorption reduces the positive
after oil aging. The reversal of the IA-H comparing core K and core J charge of the carbonate surface so that the co-adsorption of Ca2+
demonstrates the effect of naphthenic acid presence and oil aging. is increased; (2) Ca2+ and adsorbed carboxylic acids (RCOO) form
The adsorbed naphthenic acid on the limestone surface after the oil complexes (RCOO-Ca+), and the positively charged complexes
aging process is responsible for the oil-wetness. (RCOO-Ca+) are released from the positively charged rock surface
The oil recovery of Core A and Core J during spontaneous imbi- due to electrostatic repulsion. This proposed synergetic effect
bition experiments are also compared in the supplementary mate- between Ca2+ and SO2 4 also explains the observation that a higher
rial Figs. S1, S2 to show the effect of wettability. The water-wet SO24 concentration does not always result in a higher oil recovery.
Core J has 27% oil recovery (of OOIP) in 5 M NaCl at room temper- For instance, 4dSW0Mg (6 mM SO2 4 ) is more effective than LowSal
ature. However, slightly oil-wet Core A has only 6% oil recovery in Na2SO4 (55 mM SO2 4 ) on wettability alteration. The two brines
5 M NaCl at 90 °C. The oil recovery for Core A is very limited, even have similar Na+ concentration (109 mM and 149 mM) and neither
though the higher temperature is a favorable condition for oil of them contains Mg2+. However, 4dSW0Mg brine contains added
recovery. This difference demonstrates that the capillary imbibi- Ca2+ while LowSal Na2SO4 brine only has Ca2+ as a result of calcite
tion of the brine will be significantly inhibited if the core becomes dissolution. It is possible that the limiting factor for the wettability
preferentially oil-wet. alteration in LowSal Na2SO4 is the low Ca2+ concentration.
To analyze the effect of Ca2+, the calcium concentrations were
measured for some of the brines after equilibrating them with cal-
3.2. Effect of Mg2+, SO2
4 and salinity cite rock and AN1.5MO (Fig. 3). Note that even though there are no
added calcium salts in the single-electrolyte brines, Ca2+ still exist
Wettability alteration in the ‘‘smart water” is quantified by the as a result of calcite dissolution. Naphthenic acid partitioning from
additional oil recovery in the spontaneous imbibition experiment the AN1.5MO further promotes calcite dissolution, and results in
after replacing the initial brine with the ‘‘smart water”. Capillary an even higher Ca2+ concentration. The calcium level (including
imbibition is inhibited when the core is preferentially oil-wet. Ca2+ and all other Ca2+-complexes) was found to be roughly
Therefore, additional oil recovery in spontaneous imbibition test 8 mM in seawater-based brines and 5 mM in single-electrolyte
is assumed to indicate wettability alteration. brines. The seawater-based brines have 3 mM added Ca2+ (shown
Core A–C were all first tested in 5 M NaCl and then immersed in in Table 1), which is exactly the increment compared to the
4dSW, 4dSW0S and 4dSW0Mg respectively to investigate the role single-electrolyte brines. Based on the hypothesis by Strand et al.
of Mg2+ and SO2 4 in seawater-based brines (Fig. 1(a)). The oil [1], SO2
4 is only effective when Ca
2+
is present because the release
recovery in the high salinity 5 M NaCl brine is very limited in all of carboxylic acids from mineral is due to Ca-complexes formation.
the cases (3–6%). A significant amount of oil is recovered after The additional 3 mM Ca2+ in seawater-based brines can indeed
switching to low salinity brines, clearly indicating wettability contribute significantly to the wettability alteration if the Ca2+ is
alteration. An additional 28% (of OOIP) oil was observed for Core the limiting ion while SO2 4 is abundant. In that case, 55 mM
C in 4dSW0Mg, which is 4x diluted seawater depleted in Mg2+. SO2 2
4 may be the same as 6 mM SO4 because the wettability alter-
Depletion of SO2 4 in the diluted seawater results in the lowest ation is limited by relatively low Ca2+ concentration.
additional oil recovery (20% of OOIP) among the three experiments. The presence of Mg2+ is unfavorable because the depletion of
For Core A, the additional oil recovery in 4dSW is 24%. A further 5% Mg2+ (4dSW0Mg) results in higher additional oil recovery when
150 J. Song et al. / Journal of Colloid and Interface Science 563 (2020) 145–155

Fig. 1. Oil recovery history for core A–F in the spontaneous imbibition of (a) seawater-based brines and (b) single-electrolyte brines at 90 °C. Dashed lines mark the date when
brines were replaced.

Fig. 2. Summary of additional oil recovery compared to 5 M NaCl for all the seawater-based and single-electrolyte brines. All tests were performed with AN1.5MO and
Indiana limestone cores at 90 °C.
J. Song et al. / Journal of Colloid and Interface Science 563 (2020) 145–155 151

Fig. 3. ICP-OES results of calcium concentration in most of the tested brines after equilibrating with calcite and AN1.5MO.

compared with the 4dSW case, and the addition of Mg2+ (MgSO4)
results in lower additional oil recovery when compared to the Na2-
SO4 test. These observations are contradictory to the results by
Zhang et al. [15] with crude oil spontaneous imbibition results
and the density function theory (DFT) calculations by Sakuma
et al. [29], who showed that substitution of MgSO4 ion pairs onto
calcite can increase the water affinity to calcite surface, while the
substitution of SO2
4 alone does not. We hypothesize that the effect
of Mg2+ is a function of the rock/oil/brine interaction in a specific
system. Zhang et al. utilized a system with crude oils from the
North Sea and chalk while a limestone/model oil system is used
in our study. The working mechanism of Mg2+ may associate with
the specific system of Zhang et al.’s work. On the other hand, our
observation agrees with the contact angle investigation by Shaik
et al. [41] They observed higher water advancing contact angles
(ACA) in MgSO4 brine (I = 0.164 M) compared to Na2SO4 brine
(I = 0.164 M) for oil-wet calcite plates. The ACA is initially 128°
in 5 M NaCl for the oil-aged Iceland spar plates. After the ‘‘smart
water” treatment at 120 °C, the ACA reduces to 26° in Lowsal
Na2SO4 and it only reduces to 49° in Lowsal MgSO4. The tested
brine and oil in the work by Shaik et al. have the same composition Fig. 4. Summary of zeta potential measurements and additional oil recoveries from
spontaneous imbibition experiments for all tested brines.
as that studied in this work.

3.3. Role of electrostatics: Surface complexation modeling and


electrostatic component of disjoining pressure (Pelectrostatics) is then
disjoining pressure
calculated based on the zeta potential measurements to quantify
the electrostatic interaction. A model of two flat charged plates is
Electrostatic interactions have been generally considered a key
assumed. In the cases where the two charged plates have varying
factor in wettability alteration process triggered by low salinity
surface potentials of the same sign or one has a zero potential,
water. If there is sufficient electrostatic repulsion between rock/
the assumption of constant surface charge or constant surface
brine and oil/brine interfaces, the disjoining pressure may stabilize
potential can be unreasonable at small separation (less than one
the brine film between the rock and oil, resulting in a water-wet
Debye length) [44]. Therefore, we choose to use the linear superpo-
condition [27,42]. Fig. 4 summarizes the rock/brine and oil/brine
sition approximation (LSA) to calculate the electrostatic compo-
zeta potential in all tested brines along with the additional oil
nent of the disjoining pressure (Eq. (1)), which has been shown
recovery results. The zeta potentials are measured at 25 °C, which
to provide reasonable approximation lying between the results of
may deviate from the actual value at 90 °C due to the temperature
constant surface charge and constant surface potential cases [44].
effect on the ion binding to the limestone surface. However, the
P(d) is the electrostatic component of disjoining pressure as a func-
values at ambient temperature are an important indicator of the
tion of separation, d, n is the number of cations (or anions) per unit
charge interaction between rock/oil/brine during the imbibition
volume (/m3), k is Boltzmann’s constant, and T is the system tem-
test. Also, the trend of zeta potential variation is not expected to
perature in degrees Kelvin. The reduced potential is defined as
change dramatically at different temperatures [43]. The oil/brine
zeta potential with AN1.5MO remains strongly negative in all
ci ¼ zew
kT
, wherew is the surface potential. The w can be calculated
tested brines, while the Indiana limestone/brine zeta potential is from the zeta potential by assuming a slip plane distance
small in magnitude with either positive or negative sign. The inten- xs ¼ 0:0013
I0:5
[45], where I is the ionic strength of the brine. j is the
sity of electrostatic interaction cannot be directly compared quan- reciprocal Debye length, which is also a function of the ionic
titatively among different cases because both the rock/brine and strength I. Equation (1) is derived for calculating the interaction
oil/brine zeta potentials vary in different brines. Therefore, the in z:z symmetrical electrolytes such as NaCl and MgSO4. For the
152 J. Song et al. / Journal of Colloid and Interface Science 563 (2020) 145–155

case of Na2SO4, z = 1 is assumed because using the expression for a the brine is accounted for by assuming a certain concentration of
1:1 electrolyte to calculate the interaction in a 1:2 electrolyte solu- model naphthenic acid (pKa = 4.8) [37,38] in the brine so that the
tion has been shown to be acceptable in the literature [46]. More- calculated brine pH after phase equilibrium matches with the mea-
over, 4dSW brine is also considered as a 1:1 electrolyte solution sured pH.
because it primarily contains NaCl. The calculated Pelectrostatics The experimental and SCM-predicted values of the calcium con-
curves are summarized in Fig. 5. centration and zeta potential for four tested brines are summarized
in Fig. 6. The general trends for both calcium concentration and
PðdÞ ¼ 64nkTc1 c2 expðjdÞ ð1Þ
zeta potential are successfully captured by the model when com-
The extent of wettability alteration aligns with the intensity of paring these tested brines. However, calcium concentration is
rock/oil electrostatic repulsion if comparing within a group always under-predicted by roughly 2–3 mM and the rock/brine
(Na2SO4 > MgSO4 > NaCl and 4dSW0Mg > 4dSW > 4dSW0S). How- zeta potentials are over-predicted by 4 and 7 mV for 4dSW and
ever, the trend is inconsistent when comparing the single- LowSal MgSO4, respectively. The more positive zeta potential cal-
electrolyte group and the seawater-based group. For example, culated from the SCM compared with experimental results is likely
there is unfavorable rock/oil electrostatic attraction in 4dSW because naphthenic acid adsorption on the rock surface is
because the Indiana limestone is positively charged. However, neglected in the SCM. The impact of naphthenic acid is only consid-
the additional oil recovery in 4dSW is even higher than that in ered in the bulk brine phase by including a model naphthenic acid
LowSal MgSO4, where the rock is negatively charged and has elec- in the SCM, but it is not considered at the interface. Adsorption of
trostatic repulsion to the oil. Moreover, the rock/oil electrostatic the naphthenic acid on limestone surface will result in a more neg-
repulsion should be stronger in LowSal Na2SO4 than 4dSW0Mg ative rock/brine zeta potential.
according to the zeta potentials. But the additional oil recovery Since the SCM successfully predicts the qualitative trend of cal-
in 4dSW0Mg is higher than that in LowSal Na2SO4. cium concentration and zeta potential, it is further used to calcu-
The inconsistency between electrostatic repulsion and wettabil- late the surface concentration of charged species on Indiana
ity alteration indicates the positive effect of Ca2+. The Indiana lime- limestone surface (Fig. 7). Surface charge density contribution (C/
stone rock has more positive zeta potentials in the seawater-based m2) from various species is calculated by multiplying the surface
brines because of the 3 mM added Ca2+. The added Ca2+ facilitates concentration (mol/m2) of a specie and its charge in coulomb
removing the adsorbed carboxylic acids on the rock due to RCOO- (C/mol). The summation of all the positive and negative contribu-
Ca+ complexation [1], even though it is unfavorable in terms of tions yields the net surface charge density of the rock. Although the
electrostatics. The effect of Ca2+ has also been discussed in previous zeta potential in 5 M NaCl cannot be measured due to the electrical
section when comparing the oil recovery results in 4dSW0Mg and conductivity limitation, the SCM prediction in 5 M NaCl is provided
LowSal Na2SO4. The higher Ca2+ concentration is responsible for to help understand the effect of salinity. >CO3H  A5.25 and >SiO
the better outcome in 4dSW0Mg, even though SO2 4 concentration represent the organic and inorganic impurities on Indiana lime-
is lower in 4dSW0Mg. stone surface which accounts for the charge difference of natural
Disjoining pressure calculations and zeta potential measure- limestone and synthetic calcite (Table 3). The same assumption is
ments only consider the average of surface charge and fail to pro- used in the previous work by Song et al. [36] and the readers are
vide information about particular charged species on the limestone referred to their paper for detailed justification. The charge contri-
surface. Therefore, a surface complexation model (SCM) is also bution from these two species remain constant in all the tested
employed to provide insights of low-salinity-triggered wettability brines.
alteration. The surface concentration of all the charged surface spe- A significantly different mix of species is found for the rock min-
cies can be calculated from a SCM which captures the electrostatic eral in 5 M NaCl and other brines in Fig. 7. In the high salinity initial
properties of the mineral. A previously proposed SCM for Indiana brine, 91% of major surface site >CaOH0.75 is converted to the Na-
limestone [36] is used without re-fitting any of the model param- complex >CaOH  Na+0.25, and 67% of the other major site
eters. The impact of naphthenic acid partitioning from AN1.5MO to >CO3H+0.75 is converted to the Cl-complex >CO3H  Cl0.25. Even

Fig. 5. Electrostatic component of disjoining pressure for Indiana limestone/AN1.5MO in various tested brines. Calculation is based on the zeta potential results in this work.
J. Song et al. / Journal of Colloid and Interface Science 563 (2020) 145–155 153

Fig. 6. Comparison between measurements and prediction from surface complexation model (SCM) for Ca concentration and rock/brine zeta potential in various equilibrated
brines.

Fig. 7. Comparison of surface charge density contribution from various species in different brines based on surface complexation modeling results.

though Na+ and Cl are generally considered indifferent to the cal- These previous MD simulations and AFM studies are consistent
cite surface, their surface interaction with calcite is still significant with our SCM calculation of charged species concentration, indicat-
if present in a high concentration such as 5 M [12,36]. Based on this ing that the reduction of NaCl concentration triggers wettability
observation, the calcite surface interaction with Na+ in high salinity alteration due to removal of Na+ from calcite surface. This also
brine is hypothesized to be a critical mechanism responsible for agrees well with the spontaneous imbibition results. Interestingly,
the oil-wetness. A similar finding has also been reported by Liu, there is no clear link between additional oil recovery and the other
et al. [47], who performed a comprehensive study of calcite wetta- charged surface species. Moreover, the contribution of CO2 3 to sur-
bility alteration by quantum molecular dynamics (MD) simulation. face charge is minimal in all the tested brines, even though CO23 is
The MD simulation reveals that Na+ resides closer to the calcite considered to be a potential determining ion for the calcite surface.
surface compared to the divalent ions Ca2+and Mg2+, and it modi- One reason why CO2 3 ions show almost no impact is that its bulk
fies the interfacial water structure. As a result, a high Na+ concen- concentration is too low (~0.01 mM) in all the tested brines. CO2
3 is
tration reduces the adhesion energy of calcite/water and makes only present as a result of calcite dissolution and it is mostly con-
calcite less water-wet. In another MD simulation study by Koleini verted to HCO 3 at the pH around 8.0.
et al. [48], the Na+ ion is also found to be in the Stern layer, and clo-
ser to the calcite surface than the divalent ions. A direct visualiza- 4. Conclusions
tion study confirming that Na+ indeed exists in the Stern layer is
reported by Ricci et al. [49] using atomic force microscopy The effect of salinity, Mg2+ and SO2
4 on ‘‘smart water”-induced
(AFM). Moreover, another AFM study by Guo and Kovscek [50] also wettability alteration is examined using Indiana limestone cores
demonstrates that Na+ reduces the decay length of short-range and a model oil with added naphthenic acid. Low salinity brines
non-DLVO repulsion and disturbs the innermost water structure. with or without added divalent ions are found to trigger wettabil-
154 J. Song et al. / Journal of Colloid and Interface Science 563 (2020) 145–155

ity alteration and improve oil recovery in the spontaneous imbibi- Zhang for his help in the EDX measurement for the core sample.
tion test. The SO2
4 in the low salinity brines further promotes wet- We also thank Dr. Yongchao Zeng, Dr. Pengfei Dong and Mr. Zhuq-
tability alteration, while the presence of Mg2+ is unfavorable for ing Zhang for their insightful discussions.
wettability alteration. Such observation is consistent in both
single-electrolyte brines and seawater-based brines. Moreover,
other promotive mechanisms other than ionic strength and SO2 4 Appendix A. Supplementary material
is found when comparing single-electrolyte brines and seawater-
based brines. The additional oil recovery in LowSal Na2SO4, a Supplementary data to this article can be found online at
single-electrolyte solution with 55 mM SO2 4 is lower than that in https://doi.org/10.1016/j.jcis.2019.12.040.
4dSW0Mg, a seawater-based brine with only 6 mM SO2 4 . Both bri-
nes have identical ionic strength, very similar Na+ concentration
and no Mg2+. A higher Ca2+ concentration in 4dSW0Mg is believed References
to facilitate wettability alteration, as Ca2+ has also been identified
[1] S. Strand, E.J. Høgnesen, T. Austad, Wettability alteration of carbonates—effects
as a wettability modifier in the literature [1,15]. of potential determining ions (Ca2+ and SO42) and temperature, Colloids
The role of electrostatic interaction is then investigated by zeta Surf. Physicochem. Eng. Asp. 275 (1–3) (2006) 1–10, https://doi.org/10.1016/
potential measurements, disjoining pressure calculation, and sur- j.colsurfa.2005.10.061.
[2] G. Hirasaki, D.L. Zhang, Surface chemistry of oil recovery from fractured, oil-
face complexation modeling. The intensity of rock/oil electrostatic wet, carbonate formations, SPE J. 9 (02) (2004) 151–162, https://doi.org/
repulsion aligns with the additional oil recovery in spontaneous 10.2118/88365-PA.
imbibition for the three single-electrolyte brines (Na2SO4 > MgSO4 >- [3] D. Leslie Zhang, S. Liu, M. Puerto, C.A. Miller, G.J. Hirasaki, Wettability
alteration and spontaneous imbibition in oil-wet carbonate formations, J. Pet.
NaCl). However, electrostatic repulsion does not explain the better Sci. Eng. 52 (1–4) (2006) 213–226, https://doi.org/10.1016/j.
oil recovery in seawater-based brines. Wettability alteration is petrol.2006.03.009.
observed in 4dSW in which the rock/oil have electrostatic attrac- [4] G. Sharma, K.K. Mohanty, Wettability Alteration in High Temperature and High
Salinity Carbonate Reservoirs; Society of Petroleum Engineers, 2011,
tion. The better performance of seawater-based brines is explained https://doi.org/10.2118/147306-MS.
by the higher Ca2+ concentration, which promotes the wettability [5] M. Mohammed, T. Babadagli, Wettability alteration: a comprehensive review
alteration via removal of adsorbed hydrophobic carboxylic acids. of materials/methods and testing the selected ones on heavy-oil containing
oil-wet systems, Adv. Colloid Interface Sci. 220 (2015) 54–77, https://doi.org/
According to the SCM calculation, over 90% of the major surface 10.1016/j.cis.2015.02.006.
site >CaOH0.75 converts to its Na+-binding complex [6] D.C. Standnes, T. Austad, Wettability alteration in carbonates, Colloids Surf.
>CaOH  Na+0.25 in the high salinity 5 M NaCl brine. The reduction Physicochem. Eng. Asp. 216 (1–3) (2003) 243–259, https://doi.org/10.1016/
S0927-7757(02)00580-0.
of >CaOH  Na+0.25 concentration is significant in all the low salin-
[7] A. Kumar, A. Mandal, Critical investigation of zwitterionic surfactant for
ity brines. The interaction with Na+ is believed to be responsible for enhanced oil recovery from both sandstone and carbonate reservoirs:
the oil-wetness for the limestone surface. Such hypothesis is sup- adsorption, wettability alteration and imbibition studies, Chem. Eng. Sci. 209
ported by MD simulation studies [47,48] and AFM studies (2019) 115222, https://doi.org/10.1016/j.ces.2019.115222.
[8] N. Pal, N. Saxena, K.V. Divya Laxmi, A. Mandal, Interfacial behaviour,
[49,50], which show that Na+ is in the closest proximity to calcite wettability alteration and emulsification characteristics of a novel
surface and cause the calcite/water adhesion energy to decrease. surfactant: implications for enhanced oil recovery, Chem. Eng. Sci. 187
Therefore, one of the key mechanisms of low-salinity-induced wet- (2018) 200–212, https://doi.org/10.1016/j.ces.2018.04.062.
[9] P. Pillai, A. Mandal, Wettability modification and adsorption characteristics of
tability alteration is the reduction of Na+ near the rock surface. imidazole-based ionic liquid on carbonate rock: implications for enhanced oil
Further investigation is required to illuminate how the divalent recovery, Energy Fuels 33 (2) (2019) 727–738, https://doi.org/10.1021/acs.
ions affect limestone wettability in low salinity brines. Moreover, a energyfuels.8b03376.
[10] H. Ding, S. Rahman, Experimental and theoretical study of wettability
different optimal low salinity or ionic strength may exist for differ- alteration during low salinity water flooding-an state of the art review,
ent systems. The ionic strength of low salinity brines is kept con- Colloids Surf. Physicochem. Eng. Asp. 520 (2017) 622–639, https://doi.org/
stant as 0.164 M in this work but the impact of varying ionic 10.1016/j.colsurfa.2017.02.006.
[11] M.A. Sohal, G. Thyne, E.G. Søgaard, Review of recovery mechanisms of ionically
strength should also be studied in the future. modified waterflood in carbonate reservoirs, Energy Fuels 30 (3) (2016) 1904–
1914, https://doi.org/10.1021/acs.energyfuels.5b02749.
CRediT authorship contribution statement [12] J. Song, Y. Zeng, L. Wang, X. Duan, M. Puerto, W.G. Chapman, S.L. Biswal, G.J.
Hirasaki, Surface complexation modeling of calcite zeta potential
measurements in brines with mixed potential determining ions (Ca2+, CO2- 3 ,
Jin Song: Methodology, Software, Formal analysis, Investigation, Mg2+, SO2- 4 ) for characterizing carbonate wettability, J. Colloid Interface Sci.

Writing - original draft. Qing Wang: Investigation. Imran Shaik: 506 (2017) 169–179, https://doi.org/10.1016/j.jcis.2017.06.096.
[13] T. Austad, S.F. Shariatpanahi, S. Strand, C.J.J. Black, K.J. Webb, Conditions for a
Methodology, Validation. Maura Puerto: Methodology, Resources, low-salinity enhanced oil recovery (EOR) effect in carbonate oil reservoirs,
Writing - review & editing. Prem Bikkina: Writing - review & edit- Energy Fuels 26 (1) (2012) 569–575, https://doi.org/10.1021/ef201435g.
ing. Clint Aichele: Methodology, Validation. Sibani L. Biswal: Con- [14] T. Austad, S.F. Shariatpanahi, S. Strand, H. Aksulu, T. Puntervold, Low salinity
EOR effects in limestone reservoir cores containing anhydrite: a discussion of
ceptualization, Methodology, Resources, Writing - review & editing,
the chemical mechanism, Energy Fuels 29 (11) (2015) 6903–6911, https://doi.
Supervision, Project administration, Funding acquisition. George J. org/10.1021/acs.energyfuels.5b01099.
Hirasaki: Conceptualization, Methodology, Resources, Writing - [15] P. Zhang, M.T. Tweheyo, T. Austad, Wettability alteration and improved oil
review & editing, Supervision, Funding acquisition. recovery by spontaneous imbibition of seawater into chalk: impact of the
potential determining ions Ca2+, Mg2+, and SO42, Colloids Surf.
Physicochem. Eng. Asp. 301 (1–3) (2007) 199–208, https://doi.org/10.1016/
Declaration of Competing Interest j.colsurfa.2006.12.058.
[16] S.J. Fathi, T. Austad, S. Strand, Effect of water-extractable carboxylic acids in
The authors declare that they have no known competing finan- crude oil on wettability in carbonates, Energy Fuels 25 (6) (2011) 2587–2592,
https://doi.org/10.1021/ef200302d.
cial interests or personal relationships that could have appeared [17] S.J. Fathi, T. Austad, S. Strand, Water-based Enhanced Oil Recovery (EOR) by
to influence the work reported in this paper. ‘‘smart water”: optimal ionic composition for EOR in carbonates, Energy Fuels
25 (11) (2011) 5173–5179, https://doi.org/10.1021/ef201019k.
[18] S.F. Shariatpanahi, P. Hopkins, H. Aksulu, S. Strand, T. Puntervold, T. Austad,
Acknowledgement Water based EOR by wettability alteration in dolomite, Energy Fuels 30 (1)
(2016) 180–187, https://doi.org/10.1021/acs.energyfuels.5b02239.
We acknowledge the financial support from Abu Dhabi National [19] A.A. Yousef, S.H. Al-Saleh, A. Al-Kaabi, M.S. Al-Jawfi, Laboratory investigation
of the impact of injection-water salinity and ionic content on oil recovery from
Oil Company (ADNOC) and the Rice University Consortium for Pro- carbonate reservoirs, SPE Reserv. Eval. Eng. 14 (05) (2011) 578–593, https://
cesses in Porous Media (Houston, TX, USA). We thank Mr. Leilei doi.org/10.2118/137634-PA.
J. Song et al. / Journal of Colloid and Interface Science 563 (2020) 145–155 155

[20] A. Al Harrasi, R.S. Al-maamari, S.K. Masalmeh, Laboratory Investigation of Low carboxylic acids and inorganic silica impurities on the surface charge of
Salinity Waterflooding for Carbonate Reservoirs; Society of Petroleum natural carbonates using an extended surface complexation model, Energy
Engineers, 2012. https://doi.org/10.2118/161468-MS. Fuels (2019), https://doi.org/10.1021/acs.energyfuels.8b03896.
[21] K.K. Mohanty, S. Chandrasekhar, Wettability Alteration with Brine [37] T.E. Havre, J. Sjöblom, J.E. Vindstad, Oil/water-partitioning and interfacial
Composition in High Temperature Carbonate Reservoirs; Society of behavior of naphthenic acids, J. Dispers. Sci. Technol. 24 (6) (2003) 789–801,
Petroleum Engineers, 2013, https://doi.org/10.2118/166280-MS. https://doi.org/10.1081/DIS-120025547.
[22] A. Gandomkar, M.R. Rahimpour, Investigation of low-salinity waterflooding in [38] R. Huang, P. Chelme-Ayala, Y. Zhang, M. Changalov, M. Gamal El-Din,
secondary and tertiary enhanced oil recovery in limestone reservoirs, Energy Investigation of dissociation constants for individual and total naphthenic
Fuels 29 (12) (2015) 7781–7792, https://doi.org/10.1021/acs. acids species using ultra performance liquid chromatography ion mobility
energyfuels.5b01236. time-of-flight mass spectrometry analysis, Chemosphere 184 (2017) 738–746,
[23] H. Mahani, A.L. Keya, S. Berg, R. Nasralla, Electrokinetics of carbonate/brine https://doi.org/10.1016/j.chemosphere.2017.06.067.
interface in low-salinity waterflooding: effect of brine salinity, composition, [39] J.S. Buckley, K. Takamura, N.R. Morrow, Influence of electrical surface charges
rock type, and PH on f-potential and a surface-complexation model, SPE J. 22 on the wetting properties of crude oils, SPE Reserv. Eng. 4 (03) (1989)
(01) (2017) 053–068, https://doi.org/10.2118/181745-PA. 332–340, https://doi.org/10.2118/16964-PA.
[24] H. Sharma, K.K. Mohanty, An experimental and modeling study to investigate [40] P.V. Brady, N.R. Morrow, A. Fogden, V. Deniz, N. Loahardjo, Winoto,
brine-rock interactions during low salinity water flooding in carbonates, J. Pet. Electrostatics and the low salinity effect in sandstone reservoirs, Energy
Sci. Eng. (2017), https://doi.org/10.1016/j.petrol.2017.11.052. Fuels 29 (2) (2015) 666–677, https://doi.org/10.1021/ef502474a.
[25] A.A. Eftekhari, K. Thomsen, E.H. Stenby, H.M. Nick, Thermodynamic analysis of [41] I.K. Shaik, J. Song, S.L. Biswal, G.J. Hirasaki, P.K. Bikkina, C.P. Aichele, Effect of
chalk–brine–oil interactions, Energy Fuels 31 (11) (2017) 11773–11782, brine type and ionic strength on the wettability alteration of naphthenic-acid-
https://doi.org/10.1021/acs.energyfuels.7b02019. adsorbed calcite surfaces, https://doi.org/10.1016/j.petrol.2019.106567.
[26] G.J. Hirasaki, Wettability: fundamentals and surface forces, SPE Form. Eval. 6 [42] Y.-L. Chen, L. Zhang, J. Song, G. Jian, G. Hirasaki, K. Johnston, S.L. Biswal, Two-
(02) (1991) 217–226, https://doi.org/10.2118/17367-PA. step adsorption of a switchable tertiary amine surfactant measured using a
[27] Q. Xie, A. Saeedi, E. Pooryousefy, Y. Liu, Extended DLVO-based estimates of quartz crystal microbalance with dissipation, Langmuir 35 (3) (2019) 695–
surface force in low salinity water flooding, J. Mol. Liq. 221 (2016) 658–665, 701, https://doi.org/10.1021/acs.langmuir.8b03150.
https://doi.org/10.1016/j.molliq.2016.06.004. [43] H. Mahani, R. Menezes, S. Berg, A. Fadili, R. Nasralla, D. Voskov, V. Joekar-
[28] M.J. Alshakhs, A.R. Kovscek, Understanding the role of brine ionic composition Niasar, Insights into the impact of temperature on the wettability alteration by
on oil recovery by assessment of wettability from colloidal forces, Adv. Colloid low salinity in carbonate rocks, Energy Fuels 31 (8) (2017) 7839–7853, https://
Interface Sci. 233 (2016) 126–138, https://doi.org/10.1016/j.cis.2015.08.004. doi.org/10.1021/acs.energyfuels.7b00776.
[29] H. Sakuma, M.P. Andersson, K. Bechgaard, S.L.S. Stipp, Surface tension [44] J. Gregory, Interaction of unequal double layers at constant charge, J. Colloid
alteration on calcite, induced by ion substitution, J. Phys. Chem. C 118 (6) Interface Sci. 51 (1) (1975) 44–51, https://doi.org/10.1016/0021-9797(75)
(2014) 3078–3087, https://doi.org/10.1021/jp411151u. 90081-8.
[30] Sci. Rep. 6 (1) (2016), https://doi.org/10.1038/srep28854. [45] F. Heberling, T.P. Trainor, J. Lützenkirchen, P. Eng, M.A. Denecke, D. Bosbach,
[31] L. Chen, G. Zhang, L. Wang, W. Wu, J. Ge, Zeta potential of limestone in a large Structure and reactivity of the calcite-water interface, J. Colloid Interface Sci.
range of salinity, Colloids Surf. Physicochem. Eng. Asp. 450 (2014) 1–8, https:// 354 (2) (2011) 843–857, https://doi.org/10.1016/j.jcis.2010.10.047.
doi.org/10.1016/j.colsurfa.2014.03.006. [46] J. Israelachvili, Intermolecular and Surface Forces, third ed., Elsevier (2011).
[32] H. Mahani, A.L. Keya, S. Berg, W.-B. Bartels, R. Nasralla, W.R. Rossen, Insights [47] J. Liu, O.B. Wani, S.M. Alhassan, S.T. Pantelides, Wettability alteration and
into the mechanism of wettability alteration by low-salinity flooding (LSF) in enhanced oil recovery induced by proximal adsorption of Na +, Cl , Ca 2 +, Mg
carbonates, Energy Fuels 29 (3) (2015) 1352–1367, https://doi.org/10.1021/ 2 +, and SO 4 2  ions on calcite, Phys. Rev. Appl. 10 (3) (2018), https://doi.org/
ef5023847. 10.1103/PhysRevApplied.10.034064.
[33] S. Rashid, M.S. Mousapour, S. Ayatollahi, M. Vossoughi, A.H. Beigy, Wettability [48] M.M. Koleini, M.F. Mehraban, S. Ayatollahi, Effects of low salinity water on
alteration in carbonates during ‘‘smart waterflood”: underlying mechanisms calcite/brine interface: a molecular dynamics simulation study, Colloids Surf.
and the effect of individual ions, Colloids Surf. Physicochem. Eng. Asp. 487 Physicochem. Eng. Asp. 537 (2018) 61–68, https://doi.org/10.1016/
(2015) 142–153, https://doi.org/10.1016/j.colsurfa.2015.09.067. j.colsurfa.2017.10.024.
[34] M. Lashkarbolooki, S. Ayatollahi, Investigating injection of low salinity brine in [49] M. Ricci, P. Spijker, F. Stellacci, J.-F. Molinari, K. Voïtchovsky, Direct
carbonate rock with the assist of works of cohesion and adhesion and visualization of single ions in the stern layer of calcite, Langmuir 29 (7)
spreading coefficient calculations, J. Pet. Sci. Eng. 161 (2018) 381–389, https:// (2013) 2207–2216, https://doi.org/10.1021/la3044736.
doi.org/10.1016/j.petrol.2017.12.010. [50] H. Guo, A.R. Kovscek, Investigation of the effects of ions on short-range non-
[35] T. Fan, J.S. Buckley, Acid number measurements revisited, SPE J. 12 (04) (2007) DLVO forces at the calcite/brine interface and implications for low salinity oil-
496–500, https://doi.org/10.2118/99884-PA. recovery processes, J. Colloid Interface Sci. 552 (2019) 295–311, https://doi.
[36] J. Song, S. Rezaee, L. Zhang, Z. Zhang, M. Puerto, O.B. Wani, F. Vargas, S. org/10.1016/j.jcis.2019.05.049.
Alhassan, S.L. Biswal, G.J. Hirasaki, Characterizing the influence of organic

You might also like