Download as pdf or txt
Download as pdf or txt
You are on page 1of 8

Article

Cite This: Energy Fuels 2019, 33, 739−746 pubs.acs.org/EF

Low-Salinity Surfactant Nanofluid Formulations for Wettability


Alteration of Sandstone: Role of the SiO2 Nanoparticle
Concentration and Divalent Cation/SO42− Ratio
Nilesh Kumar Jha,†,‡ Stefan Iglauer,§ Ahmed Barifcani,‡ Mohammad Sarmadivaleh,‡
and Jitendra S. Sangwai*,†

Petroleum Engineering Program, Department of Ocean Engineering, Indian Institute of Technology Madras, Chennai, Tamil Nadu
600036, India

Western Australia School of Mines (WASM): Minerals, Energy and Chemical Engineering, Curtin University, 26 Dick Perry
Downloaded via INDIAN INST OF TECH (ISM) DHANBAD on March 19, 2023 at 10:11:38 (UTC).
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

Avenue, Kensington, Western Australia 6151, Australia


§
School of Engineering, Edith Cowan University, 270 Joondalup Drive, Joondalup, Western Australia 6027, Australia
*
S Supporting Information

ABSTRACT: Low-salinity water injection emerges to be a cost-effective and environmentally friendly enhanced oil recovery
technique. Furthermore, additives, such as the surfactant and nanoparticles in combination with low-salinity water, appear to be
promising formulations for rock wettability modification and surfactant adsorption control. The detailed interaction of these
novel formulations and the rock surface is, however, not well understood. Thus, an experimental study was conducted here, and
results show that the anionic surfactant (AOT, 11.247 mM) augmented the effect of silica nanoparticles (1000−3000 mg/L
concentration) at low-salinity conditions as effective surfactant adsorption control agents when used at appropriate divalent
cation/sulfate ion ratios. Low-salinity surfactant nanofluids may thus be applied for wettability alteration of oil-bearing
sandstone reservoirs for recovering residual oil. Here, we demonstrate that the ratio of divalent cations to sulfate ions (0 ≤ M2+/
SO42− ≤ 4.427) has a significant role in surfactant adsorption, irrespective of the divalent/monovalent cation ratio or the
presence of nanoparticles when sulfate ions are present in the solution. We further show using USBM wettability measurements
that initial water-wet Berea sandstone can be rendered more water-wet when 1000 mg/L silica nanoparticles are used in the
low-salinity formulation, although a further incremental nanoparticle concentration has no significant effect on the wettability.
Wettability controls the capillary pressure and relative permeability behavior and, thus, influences the rate of hydrocarbon
displacement and ultimate recovery.

1. INTRODUCTION now being considered for chemical flooding.19,22−24 Mecha-


Wettability alteration is one of the important factors in nistically, the effect of nanoparticles on surfactant adsorption
enhanced oil recovery (EOR).1−3 The degree of wettability has not yet been investigated at low-salinity conditions or in
alteration of a porous rock can be quantified in terms of the presence of alkali ions, despite alkali being known to
wettability indices. One such method, the USBM method, was reduce surfactant adsorption as well as produce in situ soap in
developed by Donaldson et al.4 In this method, the area under the presence of active oil,25,26 which has been widely used in
the secondary capillary pressure curve is measured to estimate alkali/surfactant/polymer (ASP) flooding.27 Thus, the moti-
the wettability. Note that the USBM wettability index is vation of the present work is how the addition of additives
measured as a value throughout the range from complete (surfactant-augmented nanoparticles stabilized at high pH) to
water-wet (+∞) to complete oil-wet (−∞) and 0 for neutral a low-salinity water can help in wettability alteration of the
wettability.5 Usually, these curves are measured in centrifuges.5 sandstone cores.
The rate of oil recovery and the residual oil saturation, which Mechanisms of oil recovery from LSWI in sandstone
are the target of EOR technology, are highly dependent upon reservoirs have also been widely debated and reported in the
the wettability.5 Low-salinity water injection (LSWI), which is literature, which mainly hover around wettability alteration
an emerging EOR method, also helps in the wettability through complex formation by multi-ion exchange (MIE) and
alteration.6−15 However, how precisely wettability is linked to double-layer expansion (DLE) as a result of salinity reduction.6
LSWI is only poorly understood.16 Furthermore, the surfactant The key process of MIE has been reported to be affected by
can reduce interfacial tension between oil and water to an divalent cations and the divalent/monovalent ion ratio.28
ultralow value,17 but the surfactant can also be adsorbed on the Divalent cations are proposed to act as bridging ions between
rock surface. Such surfactant adsorption on the reservoir rock negatively charged rock surface sites and acidic components in
surface detrimentally influences oil recovery.18,19 However,
some works in the past few years have reported that the control Received: September 27, 2018
of surfactant adsorption on the rock surface is possible using Revised: January 11, 2019
different additives.20,21 One such additive, nanoparticles, is Published: January 15, 2019

© 2019 American Chemical Society 739 DOI: 10.1021/acs.energyfuels.8b03406


Energy Fuels 2019, 33, 739−746
Energy & Fuels Article

the crude oil, promoting their binding. Because the clay in the Table 1. List of Chemicals Used in This Study
rock covers most of the rock surface, clay imparts an overall
purity CAS
negative charge to the surface. In one of the works, it was (mass Registry
shown that retention of dodecyl benzenesulfonate on kaolinite chemical supplier fraction) Number
clay increases with an increase in the H+ concentration, i.e., at sodium chloride Merck Specialties, ≥0.99 7647-14-5
acidic pH ranges.29 The role of ion types (divalent and (NaCl) Mumbai, India
monovalent) has also been investigated in terms of the calcium chloride Alfa Aesar, U.K. ≥0.98 10043-52-4
recovery performance and their influence on oil/brine/rock (CaCl2)
interfacial behavior at low-salinity conditions. Wei et al.30 have magnesium sulfate Rankem, Mumbai, ≥0.99 10034-99-8
heptahydrate India
highlighted the roles of HCO3−, Mg2+, and SO42− because they (MgSO4·7H2O)
constructed the most viscoelastic oil−water interfaces sodium sulfate Merck Specialties, ≥0.99 7757-82-6
compared to other ions. This viscoelasticity of the oil−water (Na2SO4) Mumbai, India
interface rendered the stability of the interface, prevented snap- sodium hydroxide Sisco Research ≥0.98 1310-73-2
(NaOH) Laboratories,
off of oil,31 improved relative permeability, and thus increased Mumbai, India
oil recovery. However, the role of these ions in wettability silicon dioxide Sisco Research ≥0.99 7631-86-9
alteration resulting in detachment of polar species from the nanopowder (SiO2) Laboratories,
sandstone rock surface could not be explained. Earlier research Mumbai, India
suggests that chromate and sulfate can be adsorbed on dioctyl sulfosuccinate Sigma-Aldrich, U.S.A. ≥0.97 577-11-7
sodium salt (AOT)
kaolinite surfaces when the pH value is in the range of 5−7. light paraffin liquid S.D. Fine Chemicals, 8012-95-1
This phenomenon was explained by a site-binding model of Mumbai, India
the kaolinite edge, where the edge is viewed as composite clay powder The Pioneer Chemical
layers of Al and Si oxide, which are typically positively Company, Delhi,
India
charged.32
We observe from the existing literature that adsorption of
prepared using ultrapure water of 18.2 MΩ cm resistivity, and pH
surfactants and/or other ionic species can have an effect on the
6.6−7.1 of the fluids was measured using a pH meter (PC 2700,
wettability of the rock surface. Hammond and Unsal, by Eutech Instruments, U.S.A.) at 25 °C. Light paraffin oil used in this
dynamic pore network modeling, showed how wettability study has an American Petroleum Institute (API) gravity of 38.98
alteration of a mixed-wet rock surface toward a more water-wet (830 kg/m3) and kinematic viscosity of 28.07 cS at 25 °C.
condition can mobilize trapped oil, which migrates toward Berea sandstone with permeability less than 200 mD is known to
larger pores and can thus reduce the trapped oil saturation in contain feldspars and muscovite in the range of 7−8%.36 Clay powder
the displacement process.49 Thus, it is the matter of interest of was heated in an oven at 120 °C for 2 h to remove water and any
the present work of how adsorption of surfactants and/or other other adsorbed materials.37 The powder was analyzed by X-ray
ionic species on the rock surface affect its wettability and how diffraction (XRD) using a Bruker D8 Discover powder diffractometer
with Cu Kα (40 mA/40 kV) radiation for phase identification by
it could be controlled by varying the proportion of ionic
determining peaks associated with various phases present in it (Figure
species and/or using nanoparticles in low-salinity conditions. 1). The instrument was operated with a step size of 0.017° 2θ angle,
Thus, in this work, we investigate the effect of the divalent ion/ while the data collection was from 10° to 90° 2θ angle. The data were
SO42− ratio on surfactant adsorption on clay at low-salinity, processed and analyzed using PANlytical X’Pert High Score Plus
high-pH aqueous nanofluids in the presence of SiO 2 software to determine the phases and their composition. Clay powder
nanoparticles (SNPs). Because the pH of the solutions mainly contained muscovite (76.2 wt %) and pyrophyllite (23.8 wt
(using NaOH) is high, we observe that the ratio of divalent/ %). Muscovite is the most common mineral in the family of mica.
monovalent cations for all of the solutions is negligible. 2.2. Method. All salt solutions were prepared gravimetrically using
Surfactant-augmented nanofluids proved to be a strong a LC GC RADWAG AS/X 220 analytical balance with ±0.000 04
mass fraction uncertainty. The aqueous phase solutions were prepared
wettability modifier for reservoir rock representative sam-
using a magnetic stirrer (IKA big squid, Germany) at 400−600 rpm at
ples1−3,23,33−35 as a result of co-adsorption of nanoaggregates 25 °C (Table 2). High-salinity brine (HSW) was prepared and later
over the surface of the rock; thus, this work is of further diluted 10 times to make low-salinity brine (LSW1). NaOH was
interest to low-salinity cases in sandstone core plugs. mixed with LSW1 to increase its pH to 12.5. Later, low-salinity brines
Therefore, in the present work, the aqueous solution [high- (LSW2, LSW3, and LSW4) with pH 12.5 were prepared in ultrapure
pH, low-salinity surfactant solutions (LSS)], which showed the water, separately. Further, high-pH, low-salinity brines containing
maximum average surfactant adsorption on the clay, was AOT (11.247 mM) were prepared and, from here onward, labeled as
chosen for the imbibition process with varying SNP high-pH, low-salinity surfactants,28 as LSS1, LSS2, LSS3, and LSS4
concentrations (wt %), i.e., LSS1 + 0% SNP, LSS1 + 0.1% (prepared using low-salinity water LSW1, LSW2, LSW3, and LSW4,
respectively). High-pH, low-salinity surfactants were used to prepare
SNP, and LSS1 + 0.2% SNP.
12 combinations of nanofluids with 1000 mg/L (0.1 wt %), 2000 mg/
L (0.2 wt %), and 3000 mg/L (0.3 wt %) SNPs by homogenizing
2. MATERIALS AND METHODS (IKA Digital-Turrax T 25, Germany) at a speed of 10 000 rpm for 15
2.1. Materials. Table 1 provides the list of chemicals used in this min for each combination before mixing it with nanoparticles and
study. Sodium chloride (NaCl, ≥99% mass fraction) and sodium then sonication.38,39 Table 3 gives the combinations of the prepared
sulfate (Na2SO4, ≥99% mass fraction) were used as monovalent aqueous solutions with and without SNPs used in this work.
background salts. Calcium chloride (CaCl2, ≥98% mass fraction) and 2.2.1. Surfactant Retention Measurements. Clay powder and
magnesium sulfate heptahydrate (MgSO4·7H2O, ≥99% mass fraction) prepared LSS + SNP solution were mixed at a 1:20 weight ratio in a
were used as divalent background salts. Dioctyl sulfosuccinate sodium test tube and then shaken for 4 h at 25 °C in an electrical shake.37 The
salt (AOT, ≥97% mass fraction) was the anionic surfactant used, and mixture was aged for 2 weeks at 25 °C and atmospheric pressure; the
light paraffin oil (saturates, >99% mass fraction) was used as an oleic tests tubes containing the aged mixtures were then centrifuged for at
phase. SNPs (average particle size of 15 nm) were used to prepare least 30 min at 6000 rpm to separate the supernatant from the clay.
nanofluids. Aqueous solutions of individual salts and surfactants were The separated supernatant was analyzed to determine the remaining

740 DOI: 10.1021/acs.energyfuels.8b03406


Energy Fuels 2019, 33, 739−746
Energy & Fuels Article

Figure 1. XRD pattern of muscovite and associated clay mineral.

Table 2. Compositions of the Brines


brine Na+ (mg/L) Ca2+ (mg/L) Mg2+ (mg/L) Cl− (mg/L) SO42− (mg/L) M2+/SO42−
HSW 11494.9 601.2 36.5 18790.1 144.1 4.424
LSW1 1149.5 60.1 3.6 1879.0 14.4 4.424
LSW2 1207.7 37.1 0 1928.1 14.4 2.576
LSW3 1252.8 11 0 1951.4 14.4 0.764
LSW4 1271.6 0 0 1961.0 14.4 0.000

Table 3. Compositions of the Nanofluids helium gas expansion porosimeter and permeameter (Coreval 30,
Vinci-Technologies, France) with a pressure transducer accuracy of
SNP SNP SNP SNP 0.1% of the full scale. Dried cores were left to saturate with HSW for
LSS (0 mg/L) (1000 mg/L) (2000 mg/L) (3000 mg/L)
another 24 h in a high-pressure saturator (Vinci-Technologies,
LSS1 LSS1 + 0% LSS1 + 0.1% LSS1 + 0.2% LSS1 + 0.3% France) at a pressure of 13.79 MPa and 25 °C. Subsequently, HSW-
LSS2 LSS2 + 0% LSS2 + 0.1% LSS2 + 0.2% LSS2 + 0.3% saturated core plugs were fixed into the drainage cell and filled with
LSS3 LSS3 + 0% LSS3 + 0.1% LSS3 + 0.2% LSS3 + 0.3% light paraffin oil. Centrifugal force was then applied stepwise to the
LSS4 LSS4 + 0% LSS4 + 0.1% LSS4 + 0.2% LSS4 + 0.3% plugs by rotation (1000−10 000 rpm) to obtain the capillary pressure
curves against average water saturation. A mounted automated video
camera (EO Edmund Optics, U.S.A.), precalibrated for the initial oil/
surfactant concentration; the change in solution surfactant concen-
tration was assumed to be equal to the amount of surfactant retained aqueous phase interface, was used to measure the change in the fluid
on clay. Specifically, an aliquot of the supernatant solution was level by change in pixels with a step-up of the rotational speed. This
analyzed by ultraviolet (UV) absorption using a spectrophotometer pixel change with known cell dimensions was used to calculate the
(JASCO V-630, U.S.A.) to obtain the absorbance values for various cumulative fluid production and, thus, water saturation automatically
wavelengths until the equilibrium concentration was reached to for each angular velocity. Centrifugation was stopped at a point when
determine the surfactant concentration based on known standards. further water production stopped, and plugs were partially saturated
The difference between the initial and remaining surfactant by oil. This process gave the primary drainage capillary pressure data
concentration provided the total surfactant retention by the clay against average water saturation as well as the connate water
powder. AOT adsorption on the clay surface were calculated in terms saturation for each plug. The plugs were then fixed into imbibition
of milligrams of AOT adsorbed per gram of clay powder. Each cells and filled with high-pH, low-salinity surfactant solutions with and
adsorption data presented in this work is the average value of three without SNPs. Then, imbibition capillary pressure curves were
measurements. generated against average water saturation by centrifugation. The
2.2.2. Wettability Index Measurements. Berea core plugs obtained drainage processes were repeated for core plugs to obtain secondary
from Kocurek Industries, Inc., U.S.A., were cut into appropriate sizes drainage capillary pressure curves against average water saturation.
suitable for the drainage/imbibition cell of the automated centrifuge The data obtained for primary imbibition and secondary drainage
system (ACES-300, Corelab, U.S.A.) (Table 4). Core plugs were then were processed numerically with a curve fitting tool application of
cleaned from free sand particles by dipping them in methanol and MATLAB (version R2017a) software to obtain functions (R2 > 0.98)
sonicated (90 W and 35 kHz) for 30 min in two sessions. Cleaned
to capture the curve trend in the given domain of saturations and
core plugs were left to dry in a hot air oven at 100 °C for 24 h.
range of capillary pressures. These equations were further integrated
Porosity and permeability of the cores were measured at 25 °C by a
for limits of end-point saturations (Table S1 of the Supporting
Information) to obtain the areas under the curves. The logarithm of
Table 4. Berea Sandstone Core Plug Properties the ratio of the areas under the secondary drainage curve (A1) and
imbibition curve (A2) were taken to give the USBM wettability index
plug ID length (cm) diameter (cm) porosity (%) permeability (mD) (I).
B2 2.395 2.555 20.926 203.04
B3 2.367 2.558 20.526 193.08 A1
I = log
B4 2.481 2.558 20.939 202.18 A2

741 DOI: 10.1021/acs.energyfuels.8b03406


Energy Fuels 2019, 33, 739−746
Energy & Fuels Article

The method, however, cannot be compared to other methods, such as unique in terms of LSS2, LSS3, and LSS4 not containing Mg2+,
contact-angle measurement40,48,51 and Amott method, when the whereas Mg2+ is present in LSS1.
wettability of the native or restored-state core is measured.41 Suitable In a molecular dynamics simulation study, Kobayashi et al.43
for plug-size cores, the USBM test contrasts the work necessary for showed that cation bridging can take place between the
displacement of one fluid by the other.5,41 The work required for the
displacement of the non-wetting phase by the wetting phase is less muscovite surface and the negatively charged sulfonate group
due to the favorable free-energy change when compared to the work of the AOT molecules; while Ca2+ formed cation bridges, Mg2+
required for the opposite displacement.5,41 This phenomenon is caused water bridging, and thus, Ca2+ plays a larger role in
attributed to the proportionality of required work to the area under wettability alteration. Furthermore, as for LSS2/LSS3/LSS4, it
the capillary pressure curve.4,5,41 Besides, as per our information, the was observed that the change in surfactant adsorption depends
method has never been applied for determination of wettability more upon the change in the ratio of divalent cations/sulfate
alteration in sandstone core plugs as a result of nanofluid injection. ions than on a change in the SNP concentration for higher
SNP concentrations (>1000 mg/L SNP). The isoelectric point
3. RESULTS AND DISCUSSION (IEP), which is the pH value where the ζ potential is 0, of
3.1. Effect of the SNP Concentration on Surfactant muscovite is reported at pH 3.5.44 Muscovite, a non-swelling
Adsorption. The average surfactant adsorption (49.75 mg/g) clay, has a negative overall (i.e., for basal planes and edges in
was the highest for LSS1 when compared against LSS2 (44.50 which edges account for 5−10% of the overall surface charge)
mg/g), LSS3 (38.57 mg/g), and LSS4 (42.87 mg/g) (Figure surface potential at higher pH (ζ potential > −100 mV at pH >
2). The largest decline (i.e., from highest to lowest) (55.59%) 10).44 Adsorption of surfactants and/or other ionic species are
greatly affected by the surface charges of minerals. Moreover,
pyrophyllite often accompanied by muscovite also shows an
overall negative potential at higher pH (ζ potential > −60 mV
at pH > 10) with IEP in the acidic range (pH in the range of
2.35−3.07).45 Thus, the charge density of the cations will
facilitate the adsorption behavior of the surfactant by cation
bridging. Further, Allen et al.46 investigated, using the neutron
reflection approach, that monovalent cations can also act as a
bridge between negatively charged mica and negatively charged
surfactant depending upon the hydration effect. It was
explained that cations, which are smaller and highly charged,
are more hydrated and reluctant to bind to the mineral surface
or the surfactant headgroup.46 However, hydration effect was
Figure 2. Experimental values for AOT adsorption on clay against the not fully understood but proposed to be a consideration when
SNP concentration: red circle, LSS1; yellow circle, LSS2; green circle, the valency of two ions (such as Ca2+ and Mg2+) is the same.46
LSS3; and blue circle, LSS4. Our hypothesis implies that preferential adsorption of
negatively charged SNPs by a negatively and densely charged
(structural) clay surface, facilitated by divalent cation bridging,
in surfactant adsorption with an increase in the SNP decreases the surfactant adsorption. Silica surface charge (IEP
concentration was measured for LSS3, which otherwise at pH 2 approximately) is pH-dependent and needs counterion
showed the highest adsorption (55.56 mg/g at 0 mg/L binding as a result of its negative surface charge at higher pH.47
SNP) among all solutions. LSS3 and LSS4 (27.28% decrease This further implies that AOT adsorption on the SNP surface,
from highest to lowest) showed a continuously decreasing however, lesser than that adsorbed by the clay surface, also
trend in surfactant adsorption with an increasing SNP occurs by divalent cation bridging because a monovalent
concentration (from 0 to 3000 mg/L). This can be attributed cation, such as Na+, does not act as a bridge as a result of the
to the decrease of the adsorption area as a result of the strong hydration effect. In this process, possibly, precipitation
adsorbance of silica nanoparticles on clay platelets in the of CaSO4 can also take place.6 The next section will thus focus
presence of salt. Baird and Walz42 observed that, in the on the dependency of surfactant adsorption upon clay surfaces
presence of salt and SNP, a microscopic ordering begins to and how this is related to the divalent cation/sulfate ion ratio.
develop, which otherwise is randomly arranged. This micro- 3.2. Effect of the Divalent Cation/Sulfate Ion Ratio on
scopic ordering took place where a small “sponge-like” Surfactant Adsorption. Figure 3 shows the surfactant
structure was visible. Whereas for LSS1, surfactant adsorption adsorption on the clay surface as a function of the divalent
increased (54.23 mg/g at 1000 mg/L SNP) and then showed a cation/sulfate ion ratio (M 2+ /SO 42−) at varying SNP
maximum decrease (13.95%) with an increase in the SNP concentrations. Average surfactant adsorptions were 50.20
concentration. The reported value of AOT adsorption on the mg/g (at 0 mg/L SNP), 47.02 mg/g (at 1000 mg/L SNP),
kaolinite surface is 51.3 mg/g at a neutral pH condition,37 39.64 mg/g (at 2000 mg/L SNP), and 38.84 mg/g (at 3000
whereas kaolinite shows no adsorption of AOT at a high pH mg/L SNP), respectively. This shows that the average
condition (pH 11).50 However, LSS2 showed a declining trend surfactant adsorption decreased with an increasing SNP
in surfactant adsorption initially and then increased with an concentration in the aqueous solutions. Moreover, there is a
increasing SNP concentration. Thus, it can be clearly observed reversal in trend of the surfactant adsorption against the M2+/
that there is a reversal in the surfactant adsorption trend SO42− curve between 0 and 3000 mg/L SNP. Interestingly,
between LSS1 and LSS2 with an increase in the SNP both the highest (55.56 mg/g at 0 mg/L SNP) and lowest
concentration (Figure 2). Interestingly, surfactant adsorption (24.68 mg/g at 3000 mg/L SNP) obtained surfactant
tended to converge at 1000 mg/L SNP concentration for adsorption values were measured at 0.763 M2+/SO42−.
LSS2, LSS3, and LSS4. Moreover, this phenomenon was However, it was observed that the change in surfactant
742 DOI: 10.1021/acs.energyfuels.8b03406
Energy Fuels 2019, 33, 739−746
Energy & Fuels Article

Figure 3. Experimental values for AOT adsorption on clay against the


divalent cation/sulfate ion ratio (M2+/SO42−): red square, 0 mg/L
SNP; yellow square, 1000 mg/L SNP; blue square, 2000 mg/L SNP;
and green square, 3000 mg/L SNP.

adsorption depends more upon the change in the SNP


concentration than on the change in the divalent cation/sulfate
ion ratio for lower M2+/SO42−values (<2.575). Furthermore,
there was a dramatic increase in surfactant adsorption (89.29%
for 3000 mg/L SNP at 4.427 M2+/SO42− and 66.50% for 2000
mg/L SNP at 4.427 M2+/SO42−) with an increase in M2+/
SO42− values of >0.763. At 1000 mg/L SNP concentration,
surfactant adsorption remained constant for 0 ≤ M2+/SO42− ≤
2.575 but increased thereafter.
These trends in the surfactant adsorption curve can be
attributed to the interplay of availability of cations (depending
upon their concentrations) for cation bridging between the
negatively charged muscovite surface and the headgroup of the
anionic surfactant, between the negatively charged muscovite
surface and the negatively charged SNP surface, between the
negatively charged SNP surface and the headgroup of the
anionic surfactant, and for precipitation by SO42− ions, at a
constant SO42− ion concentration (14.4 mg/L). When no SNP
is present (red squares in Figure 3), the introduction of cations
(at 0.763 M2+/SO42−) increases surfactant adsorption, while a
further increase in the cation concentration (37.1 and 60.1
mg/L) decreases the surfactant adsorption as a result of
precipitation of cations by SO42− ions, which prevents further Figure 4. Hysteresis loop of capillary pressure curves for (a) B2, high-
adsorption of the surfactant on the muscovite surface. At 0.2 pH, low-salinity surfactant solutions of LSS1 + 0% SNP: blue circle,
and 0.3 wt % SNP, the introduction of cations (at 0.763 M2+/ primary drainage; gray circle, secondary drainage; and orange circle,
SO42−) decreases the adsorption of surfactants on the imbibition; (b) B3, high-pH, low-salinity surfactant solutions of LSS1
muscovite surface because they preferentially facilitate bridging + 0.1% SNP: blue circle, primary drainage; gray circle, secondary
between SNPs and the muscovite surface, while a further drainage; and orange circle, imbibition; and (c) B4, high-pH, low-
increase in the cation concentration also facilitates surfactant salinity surfactant solutions of LSS1 + 0.2% SNP: blue circle, primary
drainage; gray circle, secondary drainage; and orange circle,
adsorption directly by the muscovite surface and/or already imbibition.
adsorbed SNP. For 0.1 wt % SNP, surfactant adsorption is
nearly constant until the cation concentration has increased to Table 5. Wettability Index of Berea Sandstone Core Plugs
37.1 mg/L. Furthermore, on the basis of the evidence from the
work of Baird and Walz,42 we explain that preferential plug ID imbibition fluids wettability index
adsorption of SNP facilitated by cation bridging at 0.763 B2 LSS1 + 0% SNP −0.07
M2+/SO42 is increased as a result of the increase in the SNP B3 LSS1 + 0.1% SNP 0.35
concentration from 0.1 to 0.3 wt %, thereby decreasing the B4 LSS1 + 0.2% SNP 0.16
surfactant adsorption.
3.3. Effect of SNPs on the Wettability of Berea
Sandstone Core Plugs. Panels a−c of Figure 4 show the Clearly, when exposed to anionic surfactant only, Berea core
capillary pressure hysteresis loops for three core plugs. The was mixed-wet. However, the wettability of the Berea core
loops were obtained by displacing HSW by light paraffin oil; changed from mixed-wet to water-wet48 when SNP were used
these fluids were displaced by imbibition fluids; and all fluids (Table 5). However, despite the fact that surfactant adsorption
were again displaced by light paraffin oil. Table 5 gives the for LSS1 on clay surfaces were similar for surfactant adsorption
wettability index. (49.17 mg/g adsorption for 0 mg/L SNP concentration and
743 DOI: 10.1021/acs.energyfuels.8b03406
Energy Fuels 2019, 33, 739−746
Energy & Fuels Article

Figure 5. Schematic of surfactant adsorption on the sandstone surface.

Figure 6. Schematic of nanoaggregate co-adsorption on the sandstone surface.

48.92 mg/g adsorption for 2000 mg/L SNP concentration), divalent cation/sulfate ion ratio with respect to surfactant
increasing the concentration of SNP from 1000 to 2000 mg/L adsorption was pronounced for low ratios (M2+/SO42− values
did not significantly increase water wettability. This implies of <2.575). High-pH, low-salinity surfactant solutions of LSS1
that the wettability alteration in the Berea sandstone may not (0−2000 mg/L SNP) were used for USBM wettability index
be due to surfactant adsorption itself (Figure 5) but the measurements in light-paraffin-oil-saturated Berea sandstone
adsorption of SNP in the presence of the surfactant at high pH core plugs. The 1000 mg/L SNP with LSS1 turned the rock
and low salinity. Furthermore, facilitated by high pH and low surface water-wet and can thus be used as a strong wettability
salinity, the formation of nanoaggregates may occur as a result modifier. Further increasing the SNP concentration had no or
of the availability of more surfactant molecules in the aqueous minimal effect on the sandstone wettability. This wettability
phase; subsequently, co-adsorption of the formed nano- effect can be attributed to the co-adsorption of nanoaggregate
aggregates on the sandstone surface takes place, which might surfactant complexes and surfactant molecules on the sand-
be responsible for wettability alteration toward the water-wet stone surface, which can be prominent at an optimal SNP
condition (Figure 6). On the contrary, increasing the SNP concentration. Besides, our hypothesis is that there is an
concentration can cause repulsion between the initially co- interplay of availability of divalent cations for bridging between
adsorbed (with surfactants adsorbed directly to the sandstone the clay/sandstone surface and SNP, between the SNP and
surface) and additionally formed SNP−surfactant nano- anionic surfactant, or between the clay surface and anionic
aggregate complexes, preventing them from further co- surfactant. With the facilitation by cation bridging, co-
adsorption at the negatively charged sandstone surface. adsorption of nanoaggregates and surfactants takes place over
the clay-containing sandstone surface and can alter its
4. CONCLUSION wettability, possibly favorably toward oil recovery by low-
salinity surfactant nanofluid injection.


The effect of the divalent cation/sulfate ion ratio on anionic
surfactant adsorption on clay in the presence of nanoparticles
at high-pH and low-salinity conditions was investigated for its
ASSOCIATED CONTENT
potential use in high-pH, low-salinity surfactant nanofluid *
S Supporting Information

EOR. The formulations, which gave the maximum average The Supporting Information is available free of charge on the
adsorption of surfactant on clay, were used for imbibition ACS Publications website at DOI: 10.1021/acs.energy-
experiments in Berea sandstone to measure the wettability fuels.8b03406.
alteration effect as a result of the high-pH, low-salinity Capillary pressure data and equation (PDF)


surfactant nanofluids. Clearly, the divalent cation/sulfate ion
ratio played a prominent role in surfactant adsorption on the
clay when used in combination with silica nanoparticles. The AUTHOR INFORMATION
effect of the silica nanoparticle concentration on anionic Corresponding Author
surfactant adsorption was pronounced when used in higher *Telephone: +91-44-2257-4825. Fax: +91-44-2257-4802. E-
concentrations (>1000 mg/L), whereas the effect of the mail: jitendrasangwai@iitm.ac.in.
744 DOI: 10.1021/acs.energyfuels.8b03406
Energy Fuels 2019, 33, 739−746
Energy & Fuels Article

ORCID (16) Hamon, G. Low Salinity Waterflooding: Facts, Inconsistencies


Jitendra S. Sangwai: 0000-0001-8931-0483 and the Way Forward. Petrophysics 2016, 57, 41−50.
(17) Iglauer, S.; Wu, Y.; Shuler, P.; Tang, Y.; Goddard, W. A. Alkyl
Notes Polyglycoside/1-Naphthol Formulations: A Case Study of Surfactant
The authors declare no competing financial interest. Enhanced Oil Recovery. Tenside, Surfactants, Deterg. 2011, 48, 121−


126.
ACKNOWLEDGMENTS (18) Ahmadall, T.; Gonzalez, M. V.; Harwell, J. H.; Scamehorn, J. F.
Reducing Surfactant Adsorption in Carbonate Reservoirs. SPE
The authors are grateful for the partial financial support Reservoir Eng. 1993, 8, 117−122.
provided by the Science and Engineering Research Board (19) Wu, Y.; Chen, W.; Dai, C.; Huang, Y.; Li, H.; Zhao, M.; He, L.;
(SERB), Department of Science & Technology, Government Jiao, B. Reducing Surfactant Adsorption on Rock by Silica
of India (Project EMR/2016/005360) and Industrial Con- Nanoparticles for Enhanced Oil Recovery. J. Pet. Sci. Eng. 2017,
sultancy and Sponsored Research (IC&SR), Indian Institute of 153, 283−287.
Technology (IIT) Madras (Project ICS/16-17/831/RFIE/ (20) Ahmadi, M. A.; Shadizadeh, S. R. Adsorption of Novel
MAHS) and the Institute of Reservoir Studies, Oil and Natural Nonionic Surfactant and Particles Mixture in Carbonates: Enhanced
Oil Recovery Implication. Energy Fuels 2012, 26, 4655−4663.
Gas Corporation (ONGC) Limited, Ahmedabad, India, for (21) Shamsijazeyi, H.; Hirasaki, G. J.; Verduzco, R. Sacrificial Agent
providing a facility to carry out a part of this work. for Reducing Adsorption of Anionic Surfactants. Proceedings of the

■ REFERENCES
(1) Al-Anssari, S.; Barifcani, A.; Wang, S.; Maxim, L.; Iglauer, S.
SPE International Symposium on Oilfield Chemistry; The Woodlands,
TX, April 8−10, 2013; Paper SPE-164061-MS, DOI: 10.2118/
164061-MS.
Wettability Alteration of Oil-Wet Carbonate by Silica Nanofluid. J. (22) Hendraningrat, L.; Li, S.; Torsæter, O. A Coreflood
Colloid Interface Sci. 2016, 461, 435−442. Investigation of Nanofluid Enhanced Oil Recovery. J. Pet. Sci. Eng.
(2) Al-Anssari, S.; Arif, M.; Wang, S.; Barifcani, A.; Lebedev, M.; 2013, 111, 128−138.
Iglauer, S. Wettability of Nano-Treated Calcite/CO2/Brine Systems: (23) Kuang, W.; Saraji, S.; Piri, M. A Systematic Experimental
Implication for Enhanced CO2 Storage Potential. Int. J. Greenhouse Investigation on the Synergistic Effects of Aqueous Nanofluids on
Gas Control 2017, 66, 97−105. Interfacial Properties and Their Implications for Enhanced Oil
(3) Nwidee, L. N.; Lebedev, M.; Barifcani, A.; Sarmadivaleh, M.; Recovery. Fuel 2018, 220, 849−870.
Iglauer, S. Wettability Alteration of Oil-Wet Limestone Using (24) Mohajeri, M.; Hemmati, M.; Shekarabi, A. S. An Experimental
Surfactant-Nanoparticle Formulation. J. Colloid Interface Sci. 2017, Study on Using a Nanosurfactant in an EOR Process of Heavy oil in a
504, 334−345. Fractured Micromodel. J. Pet. Sci. Eng. 2015, 126, 162−173.
(4) Donaldson, E. C.; Thomas, R. D.; Lorenz, P. B. Wettability (25) Olajire, A. A. Review of ASP EOR (Alkaline Surfactant Polymer
Determination and Its Effect on Recovery Efficiency. SPEJ, Soc. Pet. Enhanced Oil Recovery) Technology in the Petroleum Industry:
Eng. J. 1969, 9, 13−20. Prospects and Challenges. Energy 2014, 77, 963−982.
(5) Tiab, D.; Donaldson, E. C. Wettability. Petrophysics 4th Edition. (26) Sharma, H.; Dufour, S.; Arachchilage, G. W. P. P.;
2012, 371−418. Weerasooriya, U.; Pope, G. A.; Mohanty, K. Alternative Alkalis for
(6) Al-Shalabi, E. W.; Sepehrnoori, K. A Comprehensive Review of ASP Flooding in Anhydrite Containing Oil Reservoirs. Fuel 2015,
Low Salinity/Engineered Water Injections and Their Applications in 140, 407−420.
Sandstone and Carbonate Rocks. J. Pet. Sci. Eng. 2016, 139, 137−161. (27) Guo, H.; Li, Y.; Wang, F.; Gu, Y. Comparison of Strong-Alkali
(7) Berg, S.; Cense, A. W.; Jansen, E.; Bakker, K. Direct and Weak-Alkali ASP-Flooding Field Tests in Daqing Oilfield. SPE
Experimental Evidence of Wettability Modification by Low Salinity. Prod. Oper. 2018, 33, 353−362.
Petrophysics 2010, 51, 314−322. (28) Hosseinzade Khanamiri, H.; Torsaeter, O.; Stensen, J. A. Effect
(8) Hadia, N. J.; Ashraf, A.; Tweheyo, M. T.; Torsæter, O. of Calcium in Pore Scale Oil Trapping by Low-Salinity Water and
Laboratory Investigation on Effects of Initial Wettabilities on Surfactant Enhanced Oil Recovery at Strongly Water-Wet Conditions:
Performance of Low Salinity Waterflooding. J. Pet. Sci. Eng. 2013, In Situ Imaging by X-ray Microtomography. Energy Fuels 2016, 30,
105, 18−25. 8114−8124.
(9) Hosseinzade Khanamiri, H.; Nourani, M.; Tichelkamp, T.; (29) Hanna, H. S.; Somasundaran, P. Equilibration of Kaolinite in
Stensen, J. Å.; Øye, G.; Torsæter, O. Low-Salinity-Surfactant Aqueous Inorganic and Surfactant Solutions. J. Colloid Interface Sci.
Enhanced Oil Recovery (EOR) with a New Surfactant Blend: Effect 1979, 70, 181−191.
of Calcium Cations. Energy Fuels 2016, 30, 984−991. (30) Wei, B.; Wu, R.; Lu, L.; Ning, X.; Xu, X.; Wood, C.; Yang, Y.
(10) Jha, N. K.; Iglauer, S.; Sangwai, J. S. Effect of Monovalent and Influence of Individual Ions on Oil/Brine/Rock Interfacial Inter-
Divalent Salts on the Interfacial Tension of n-Heptane against actions and Oil-Water Flow Behaviors in Porous Media. Energy Fuels
Aqueous Anionic Surfactant Solutions. J. Chem. Eng. Data 2018, 63, 2017, 31, 12035−12045.
2341−2350. (31) Iglauer, S.; Favretto, S.; Spinelli, G.; Schena, G.; Blunt, M. J. X-
(11) Kakati, A.; Sangwai, J. S. Effect of Monovalent and Divalent ray Tomography Measurements of Power-Law Cluster Size
Salts on the Interfacial Tension of Pure Hydrocarbon-Brine Systems Distributions for the Nonwetting Phase in Sandstones. Phys. Rev. E
Relevant for Low Salinity Water Flooding. J. Pet. Sci. Eng. 2017, 157, 2010, 82, 1−3.
1106−1114. (32) Zachara, J. M.; Cowan, C. E.; Schmidt, R. L.; Ainsworth, C. C.
(12) Mahani, H.; Berg, S.; Ilic, D.; Bartels, W. B.; Joekar-Niasar, V. Chromate Adsorption by Kaolinite. Clays Clay Miner. 1988, 36, 317−
Kinetics of Low-Salinity-Flooding Effect. SPE J. 2015, 20, 008−020. 326.
(13) Mahani, H.; Keya, A. L.; Berg, S.; Bartels, W. B.; Nasralla, R.; (33) Al-Anssari, S.; Arif, M.; Wang, S.; Barifcani, A.; Lebedev, M.;
Rossen, W. R. Insights Into the Mechanism of Wettability Alteration Iglauer, S. Wettability of Nanofluid-Modified Oil-Wet Calcite at
by Low-Salinity Flooding (LSF) in Carbonates. Energy Fuels 2015, 29, Reservoir Conditions. Fuel 2018, 211, 405−414.
1352−1367. (34) Nwidee, L. N.; Al-Anssari, S.; Barifcani, A.; Sarmadivaleh, M.;
(14) Morrow, N.; Buckley, J. Improved Oil Recovery by Low- Lebedev, M.; Iglauer, S. Nanoparticles Influence on Wetting
Salinity Waterflooding. JPT, J. Pet. Technol. 2011, 63, 106−112. Behaviour of Fractured Limestone Formation. J. Pet. Sci. Eng. 2017,
(15) Tang, G. Q.; Morrow, N. R. Influence of Brine Composition 149, 782−788.
and Fines Migration on Crude Oil/Brine/Rock Interactions and Oil (35) Nwidee, L. N.; Barifcani, A.; Sarmadivaleh, M.; Iglauer, S.
Recovery. J. Pet. Sci. Eng. 1999, 24, 99−111. Nanofluids as Novel Alternative Smart Fluids for Reservoir

745 DOI: 10.1021/acs.energyfuels.8b03406


Energy Fuels 2019, 33, 739−746
Energy & Fuels Article

Wettability Alteration. In Novel NanomaterialsSynthesis Applica-


tions; Kyzas, G. Z., Mitropoulos, A. C., Eds.; IntechOpen Limited:
London, U.K., 2018; pp 327−357, DOI: 10.5772/intechopen.72267.
(36) Kareem, R.; Cubillas, P.; Gluyas, J.; Bowen, L.; Hillier, S.;
Greenwell, H. C. Multi-technique Approach to the Petrophysical
Characterization of Berea Sandstone Core Plugs (Cleveland Quarries,
USA). J. Pet. Sci. Eng. 2017, 149, 436−455.
(37) Wu, Y.; Iglauer, S.; Shuler, P.; Tang, Y.; Goddard, W. A.
Experimental Study of Surfactant Retention on Kaolinite Clay.
Tenside, Surfactants, Deterg. 2011, 48, 346−358.
(38) Al-Anssari, S.; Wang, S.; Barifcani, A.; Iglauer, S. Oil-Water
Interfacial Tensions of Silica Nanoparticle-Surfactant Formulations.
Tenside, Surfactants, Deterg. 2017, 54, 334−341.
(39) Roustaei, A.; Bagherzadeh, H. Experimental Investigation of
SiO2 Nanoparticles on Enhanced Oil Recovery of Carbonate
Reservoirs. J. Pet. Explor. Prod. Technol. 2015, 5, 27−33.
(40) Kakati, A.; Sangwai, J. S. Wettability Alteration of Mineral
Surface During Low-Salinity Water Flooding: Role of Salt Type, Pure
Alkanes, and Model Oils Containing Polar Components. Energy Fuels
2018, 32, 3127−3137.
(41) Anderson, W. G. Wettability Literature SurveyPart 2:
Wettability Measurement. JPT, J. Pet. Technol. 1986, 38, 1246−1262.
(42) Baird, J. C.; Walz, J. Y. The Effects of Added Nanoparticles on
Aqueous Kaolinite Suspensions: I. Structural effects. J. Colloid
Interface Sci. 2006, 297, 161−169.
(43) Kobayashi, K.; Liang, Y.; Murata, S.; Matsuoka, T.; Takahashi,
S.; Amano, K. I.; Nishi, N.; Sakka, T. Stability Evaluation of Cation
Bridging on Muscovite Surface for Improved Description of Ion-
Specific Wettability Alteration. J. Phys. Chem. C 2017, 121, 9273−
9281.
(44) Marion, C.; Jordens, A.; McCarthy, S.; Grammatikopoulos, T.;
Waters, K. E. An Investigation Into the Flotation of Muscovite With
an Amine Collector and Calcium Lignin Sulfonate Depressant. Sep.
Purif. Technol. 2015, 149, 216−227.
(45) Liu, X.; Bai, M. Effect of Chemical Composition on the Surface
Charge Property and Flotation Behavior of Pyrophyllite Particles. Adv.
Powder Technol. 2017, 28, 836−841.
(46) Allen, F. J.; Griffin, L. R.; Alloway, R. M.; Gutfreund, P.; Lee, S.
Y.; Truscott, C. L.; Welbourn, R. J. L.; Wood, M. H.; Clarke, S. M. An
Anionic Surfactant on an Anionic Substrate: Monovalent Cation
Binding. Langmuir 2017, 33, 7881−7888.
(47) Griffin, L. R.; Browning, K. L.; Truscott, C. L.; Clifton, L. A.;
Webster, J.; Clarke, S. M. A Comparison of Didodecyldimethylam-
monium Bromide Adsorbed at Mica/Water and Silica/Water
Interfaces Using Neutron Reflection. J. Colloid Interface Sci. 2016,
478, 365−373.
(48) Iglauer, S.; Pentland, C. H.; Busch, A. CO2 Wettability of Seal
and Reservoir Rocks and the Implications for Carbon Geo-
sequestration. Water Resour. Res. 2015, 51, 729−774.
(49) Hammond, P. S.; Unsal, E. A Dynamic Pore Network Model
for Oil Displacement by Wettability-Altering Surfactant Solution.
Transp. Porous Media 2012, 92, 789−817.
(50) Suzzoni, A.; Barre, L.; Kohler, E.; Levitz, P.; Michot, L. J.;
M’Hamdi, J. Interactions Between Kaolinite Clay and AOT. Colloids
Surf., A 2018, 556, 309−315.
(51) Iglauer, S. CO2−Water−Rock Wettability: Variability, Influenc-
ing Factors, and Implications for CO2 Geostorage. Acc. Chem. Res.
2017, 50, 1134−1142.

746 DOI: 10.1021/acs.energyfuels.8b03406


Energy Fuels 2019, 33, 739−746

You might also like