Sei sulla pagina 1di 13

Article

Cite This: Ind. Eng. Chem. Res. 2018, 57, 6364−6376 pubs.acs.org/IECR

Silica Nanoparticle Assisted Polymer Flooding of Heavy Crude Oil:


Emulsification, Rheology, and Wettability Alteration Characteristics
Rahul Saha, Ramgopal V. S. Uppaluri, and Pankaj Tiwari*
Department of Chemical Engineering, Indian Institute of Technology Guwahati, Guwahati, Assam 781039, India
*
S Supporting Information

ABSTRACT: Polymer flooding based enhanced oil recovery


(EOR) has been proven to be effective and successful for
heavy crude oil systems due to its ability to achieve favorable
mobility ratios driven by viscosity alterations. Nanoparticles
have the potential to intensify favorable mobility ratios as well
as various other mechanisms responsible for EOR. This work
addresses the interfacial tension (IFT), emulsification, and
rheological and wettability alteration characteristics of silica
nanoparticle (SNP) assisted polymer flooding. The stability of
the xanthan gum−SNP system was evaluated by a particle size
analyzer and zeta potential method. The rheological behavior
of the system in addition to IFT and the emulsification
mechanism at variant temperatures were also investigated. Wettability alteration studies involved contact angle measurements for
oil drops with respect to a nanoparticle saturated rock surface. SNP assisted polymer flooding enabled higher oil recovery of
20.82% at 30 °C and 18.44% at 80 °C due to IFT reduction, higher viscosity, better emulsion stability, and wettability alteration.

1. INTRODUCTION 70 to 86% by introducing preflush slug prior to polymer


Among nonrenewable imperative energy sources, crude oil, flooding at optimal slug size of 0.2 PV. Other studies reported
natural gas, and gas condensate are prominent. Existing trends that polymer degradation occurs due to the presence of amide
in ever increasing world energy demands indicate the definitive groups in polymers which interact with ions present in the
need to enhance crude oil production from oil reservoirs. To do reservoir formation water and undergo hydrolysis to form
so, primary and secondary crude oil production techniques are carboxylic groups.17,18 Thus, aqueous system viscosity gets
not sufficient enough to recover more than one-third of the reduced and precipitation occurs. Based on the data collected
original oil in place (OOIP). The secondary recovery using from various laboratory core flooding and pilot tests,
water flooding is well-known to exhibit poor sweep efficiency Saboorian-Jooybari et al.11 chose the operating temperature
and recovery factors due to viscous fingering. To circumvent to be below 65 °C. Higher operating temperatures lead to
the problem, polymer flooding has been proposed as early as degradation or break down and viscosity loss, which affect
1950.1,2 Polymer thickening alters rheological properties of the system rheology (mobility ratio). The thermal stability of the
system to promote oil recover through minimization of water system is also dependent on the salinity conditions. Few
channelling and enhancement of sweep efficiency. Further, researchers indicated that for 80 °C and 17 000 ppm salt
polymer addition facilitates surface active transformations and concentration, xanthan gum retains 60% viscosity for 300
stable oil−water emulsions can be produced.3−6 For a polymer days.19,20 The thermal stability of oil−water emulsions also
assisted enhanced oil recovery (EOR) process, surface active controls oil recovery factor and higher temperature detrimen-
transformations refer to the potential of the polymer to form an tally influences emulsion stability due to enhancement in oil
emulsion with the crude oil. Numerous successful laboratory, droplets coalescence.21,22
pilot plant, and field applications of polymer flooding have been Recent studies targeted nanoparticles applications during
reported for mature oil fields and associated systems.7−11 chemical EOR studies and indicated that such methodologies
The performance characteristics of polymer flooding have significant potential to increase oil recovery factor.23−28
operations vary significantly due to variations in reservoir Kumar et al.27 studied the role of surfactant−polymer−
conditions and crude oil properties. Few studies showed that nanoparticle Pickering emulsion formulation with light mineral
variations in temperature and salinity (moderate to extreme) do oil and were successful to achieve 24.81% additional oil
strongly influence polymer flooding performance.12−15 Alghar-
aib et al.16 recovered 18.3% additional oil by performing Received: February 2, 2018
polymer flooding. However, upon salt introduction, the oil Revised: April 4, 2018
recovery reduced due to chemical degradation of the polymer Accepted: April 20, 2018
slug. Authors could improve the cumulative oil recovery from Published: April 20, 2018

© 2018 American Chemical Society 6364 DOI: 10.1021/acs.iecr.8b00540


Ind. Eng. Chem. Res. 2018, 57, 6364−6376
Industrial & Engineering Chemistry Research Article

recovery. The synergy of SiO2 nanoparticles, surfactant investigations.40,41 Polymer−SNP solutions were subjected to
(sodium dodecyl benzenesulfonate), and polymer (carboxy characterization studies such as dynamic light scattering (DLS),
methyl cellulose) results in interfacial tension (IFT) reduction zeta potential, rheology, and wettability alterations. The oil−
and minimizes droplets to coalesce, which enhances rheological polymer and oil−polymer−SNP emulsions were evaluated for
behavior. Similarly, Sharma et al.28 observed increase in oil average droplet size, stability period, and creaming index.
recovery from 15.97% to 24.68% by introducing SiO 2 Flooding experiments were conducted using a Berea core to
nanoparticles in a surfactant (sodium dodecyl sulfate)−polymer evaluate the effect of polymer and polymer−SNP systems on
(polyacrylamide) flooding system. With particle sizes of about residual oil recovery.
100 nm, nanoparticles possess higher surface area and thus
create larger contact with oil. Nanoparticles enhance oil−water 2. MATERIALS AND METHODS
emulsion stability by adsorbing on the oil−water interface 2.1. Materials. Crude oil and reservoir formation water
which minimizes flocculation and emulsion droplet coales- samples were obtained from Oil India Limited (OIL), Assam,
cence.29 Nanoparticles in the polymer flooding system enhance India. Berea sandstone core (Idaho Gray) was procured from
the system viscosity which improves the mobility ratio and Kocurek Industries, Hard rock division, USA. Sodium chloride
results in better displacement efficiency.25,28,30,31 Targeted (NaCl), potassium chloride (KCl), magnesium sulfate
nanoparticles for polymer flooding studies include TiO2, Al2O3, heptahydrate (MgSO4·7H2O), and calcium chloride dihydrate
and SiO2.32,33 Karimi et al.34 studied the effect of ZrO2 (CaCl2·2H2O) were purchased from Merck Specialties Pvt.
nanofluids on carbonate rock and found that the system Ltd., India. Sodium bicarbonate (NaHCO3) and xanthan gum
wettability changed from strongly oil-wet to strongly water-wet. were acquired from Sigma-Aldrich Chemicals Pvt. Ltd., India.
The reported investigations conveyed that the nanoparticle Hydrophilic silicon dioxide nanoparticles (SNP) with an
distribution in the pores and throats of the reservoir rock can average particle size of 15 nm was supplied by Sisco Research
alter the wettability of the system.35 It has been reported that Laboratories Pvt. Ltd., India.
polymer could encourage bridging induced flocculation for the 2.2. Methods. 2.2.1. Characterization of Crude Oil,
nanoparticles existing in the aqueous phase and contribute to Reservoir Formation Water, and Berea Core. The properties
enhancement in the rheology properties.36−38 Also, polymer− of Assam crude oil such as acid number, viscosity, density, and
nanoparticle interaction has been inferred to provide better salt API gravity have been evaluated to be 2.72 mg KOH/g sample,
and temperature tolerance.38 20.1 mPa·s (at 30 °C), 926.6 kg/m3, and 21.2°, respectively.
A critical insight indicates that both polymer flooding and Reservoir formation water was characterized to examine its
nanoparticle assisted polymer flooding received little attention ionic composition as Na+ = 1308 ppm, Ca2+ = 18 ppm, K+ = 65
with respect to integrated studies associated with IFT ppm, Mg2+ = 12 ppm, Cl− = 876 ppm, HCO3− = 2120 ppm,
reduction, wettability alteration, emulsion stability, rheology, and SO42− = 46 ppm with TDS (total dissolved solids) = 4445
and tertiary oil recovery. Few studies are available for ppm.40,41 X-ray diffraction (XRD) analysis (make Bruker,
emulsification and emulsion stability during polymer flooding model D8-Advance) of the crushed powder Berea core enabled
studies,3,4,6 but the role of nanoparticles in such systems has determination of mineralogical constitution of the sample.
never been evaluated. In this regard, it is important to note that Figure 1 depicts the spectra of XRD analysis and affirms the
emulsion stability studies during polymer flooding can provide
insights with respect to reduction in surfactant concentration at
a later stage during alkali surfactant polymer (ASP) flooding
and thereby reduce the cost and environmental hazard risk of
the ASP process. Also, the effect of temperature and salinity of
such processes has been studied to a limited extent.21,28,39
Sharma et al.28 concluded that the effect of temperature on
viscosity for polymer (polyacrylamide)−nanoparticle (SiO2)
solution is relatively less compared to that of polymer aqueous
solution. Nanoparticles were also effective in reducing the IFT
at various temperature until saturation of nanoparticles occurs
at the crude oil−nanoparticle interface. However, the relation
between IFT reduction and emulsification using nanoparticles
was not considered in such studies. In another study, Sharma et
al.21 investigated the thermal stability of the Pickering emulsion
prepared from surfactant−polymer−nanoparticles. They varied
the temperature and salinity of the system and observed that, at
1 wt % NaCl and 1 wt % SiO2, the emulsion stability was
Figure 1. XRD spectra of Sandstone Berea core (Idaho Gray).
improved from 73 to 77 °C. Thus, the benefits associated with
the utility of polymer as a surface active agent to promote
emulsification and the trade-offs associated with the addition of
nanoparticles need to be critically examined. It is expected that Berea core sample mineral composition as quartz (45.5%),
higher nanoparticle concentrations could clog the porous dolomite (39.7%), feldspar (6.5%), illite (6.4%), kaolinite
structure of the oil reservoirs and decrease oil recovery. (1.8%), and montmorillonite (0.1%).40 In due course of SNP
This article targets the polymer and silica nanoparticle (SNP) assisted polymer flooding, there is a tendency of dispersed
assisted polymer flooding for Assam crude oil (India). Detailed nanoparticles to form large clusters and become retained on the
characterizations of Assam crude oil and reservoir rock Berea core. Hence, nanoparticle adsorption in porous structures
considered in this study have been reported in our previous was ensured by conducting field emission scanning electron
6365 DOI: 10.1021/acs.iecr.8b00540
Ind. Eng. Chem. Res. 2018, 57, 6364−6376
Industrial & Engineering Chemistry Research Article

Figure 2. Schematic diagram of the core flooding experimental setup.

microscopy (FESEM, make Zeiss, model Sizma) of core phase. The glass vessel was then placed in the measuring
samples obtained after SNP polymer flooding experiments. position of the analytical instrument and the movement of the
2.2.2. Nanoparticle Sample Preparation. Homogeneous vessel was controlled using system software (DYNALYZER).
xanthan gum (XG) polymer solutions of 1000−5000 ppm The IFT measurements were performed in the temperature
concentrations were prepared by slowly adding the polymer to range of 30−70 °C.
the aqueous phase (synthetic formation water) and continue Emulsification test was conducted by shaking the sample
stirring the sample for 12 h to avoid agglomeration. Polymer− rigorously in an orbital shaker for 30 min, maintaining a crude
SNP solutions of 0.1−0.5 wt % were prepared by slowly adding oil to aqueous phase ratio of 3:7 (v/v) in the vial.44 The
silica nanoparticles in the homogeneous polymer solution. The emulsion produced was collected in a slit and exposed to a
samples were rigorously mixed followed by 1 h sonication in a microscope (make Nikon, model Eclipse LV100N POL).
sonication bath (make Elma, model Elmasonic S 30 H).42,43 Microscopic images were scrutinized to calculate the average
2.2.3. Particle Size Analyzer and Zeta Potential. The droplet size of the emulsion using ImageJ software.41 The
particle size distributions of aqueous polymer−SNP solutions deposition of SNPs in the oil droplets and the oil−water
and their zeta potential values were measured at 25 °C by interface was examined by FESEM analysis of the oil−
dynamic light scattering (DLS) using a particle size analyzer polymer−SNP emulsion. An emulsion drop was placed in the
instrument (make Beckman Coulter, model Delsa Nano C). cover glasses with size similar to that of aluminum stub. The
The sample was transferred into the cuvette which was housed drop was then spread in the cover glasses, and after drying, the
in the chamber of the instrument and a laser beam was sample was attached to the stub with adhesive taps, coated with
illuminated. Brownian motion of SNPs based on laser beam gold, and analyzed using FESEM.21
fluctuations were eventually used to evaluate particle size 2.2.6. Creaming Index. The creaming index (CI) was
distributions and zeta potential of the sample. evaluated to examine the role of silica nanoparticles on
2.2.4. Rheological Behavior. Viscosity measurements of the emulsion stability. The emulsion formed was kept in an
aqueous phase polymer−SNP solutions and the emulsions undisturbed position for days in a sample vial and was
formed due to the interaction of crude oil, polymer, and SNPs visualized periodically. The stability was observed by measuring
were analyzed using a rheometer (make Anton Paar, model the height of lower serum and the height of total emulsion. CI
MCR 301) at 30 and 80 °C. Sample viscosity was measured by of the sample was measured using the following expression:27,45
varying the shear rate (1/s) from 1 to 1000, and for simplicity, Hserum
the values at 1 shear rate (near zero shear viscosity) are CI = × 100%
Hemulsion (1)
reported with standard deviation of four measurements for each
sample.28 Frequency sweep (oscillation) measurements were where CI is the percentage creaming index, Hserum is the height
carried out to observe the behavior of G′ (storage modulus) of the lower serum, and Hemulsion is the height of total emulsion.
and G″ (loss modulus). Angular frequency (ω) was varied from 2.2.7. Contact Angle Measurements. Contact angle
0.01 to 100 (rad/s). Storage modulus (G′) indicates gel/solid- addresses the wettability alteration mechanism in oil reservoir
like behavior, and loss modulus (G″) refers to viscous/liquid- which assists in releasing trapped residual oil from the pores of
like behavior. the reservoir rock. A drop shape analyzer (make Holmarc,
2.2.5. Interfacial Tension and Emulsification Test of Crude model HO-IAD-CAM-01B) with built-in software (version
Oil with Xanthan Gum−SNPs. IFTs of crude oil and 8.1.0.0) was used to measure the contact angle of core−
polymer−SNP solution were measured using an automatic polymer−SNP−crude oil systems. Berea core was cut into
surface tensiometer (make Kwoya, model DY300) with the slices, and the slices were polished and saturated with oil for 3
Wilhelmy plate method.28 The measurement procedure days at 80 °C. The oil saturated Berea core slice was then
involved the prior burning of the plate to become red hot to saturated with the polymer−SNPs solution for another 24 h
remove impurities present. The dense phase (aqueous media) and then washed with the deionized water. The sample was
was transferred into a glass vessel, and then, the light phase (oil then dried in an oven at 40 °C for 24 h. The rock sample was
media) was poured without disturbing the surface of the dense eventually submerged in a water container, and an oil droplet
6366 DOI: 10.1021/acs.iecr.8b00540
Ind. Eng. Chem. Res. 2018, 57, 6364−6376
Industrial & Engineering Chemistry Research Article

was placed on the rock surface using a needle. The variation in 3. RESULTS AND DISCUSSION
contact angle was then measured at various time intervals.39,46 3.1. Stability Evaluation of SNPs by Particle Size and
2.2.8. Core Flooding Experiments. Figure 2 illustrates a Zeta Potential Measurements. The stability of the XG−
schematic diagram of the core flooding experimental setup that SNP solution and formation water−SNP systems were
refers to an assembled arrangement of a horizontal core reactor examined by visualizing the time-dependent transparency of
(hosting a core holder of dimensions 3.9 cm diameter and 11 sample in the vial. An XG concentration of 5000 ppm was
cm length). The temperature of the system was maintained by a selected based on the emulsification studies (see section 3.4)
heat blower. Pressure transducers were placed at the inlet and and SNPs of 0.3 wt % considering emulsification and IFT
outlet of the core reactor to measure the pressure difference behavior (see section 3.3). Figure 3a and b depict images of
across the core. Three hydraulic pumps were deployed to
supply crude oil, water/brine, and chemicals to the core reactor,
and an additional pump was used to create overburden
pressure. A back pressure regulator at the outlet of the reactor
was used to govern the system pressure.41,47
Sandstone Berea core (Idaho Gray) was used to perform the
flooding experiments. Reservoir formation water was first
injected in the core reactor for brine saturation and absolute
brine permeability was then calculated using Darcy’s law with
time-dependent measurements of brine effluent volume and Figure 3. Images of nanoparticles in the aqueous phase at 30 °C (a)
pressure difference across the core reactor. 5000 ppm xanthan gum and 0.3 wt % SNPs and (b) reservoir
formation water and 0.3 wt % SNPs.
Q = −kAΔp /(μL) (2)
samples prepared for XG−SNP aqueous system and formation
where Q = total discharge (cm3/s), A = cross-sectional flow water−SNP, respectively (kept for 1 and 10 days). Figure 3a
area (cm2), Δp = pressure drop between the inlet and outlet of indicates the better stability of XG−SNP systems due to the
the core reactor (atm), μ = viscosity of brine (cP), and L = absence of sedimentation phenomena. Figure 3b shows
length of core (cm). contrasting stability behavior for formation water−SNP system,
Oil saturation was then facilitated by injecting oil in the core with rigorous sedimentation of 70% nanoparticles in first 24 h
until brine cut reaches to less than 1%, and the system was left and settling of almost all nanoparticles after 10 days indicating a
for 24 h to attain uniformity. After achieving reservoir transparent and clear aqueous phase. Similar inferences have
conditions, the initial oil in place (IOIP), irreducible brine been deduced for formation water−SNP systems prepared with
saturation/initial oil saturation, and porosity of the core were 0.1 and 0.5 wt % SNPs. The results suggest that the SNPs can
measured through material balance expressions be helpful in polymer flooding not in water flooding.
To further validate the negligible effect of sedimentation
Φ = [Vp/Vt] × 100 (3) phenomena of the XG−SNP solution system, particle size
distribution and zeta potential were evaluated using dynamic
Vp = Virb + Vo (4) light scattering (DLS). Figure 4 presents the average nano-

Virb = (Vib) − (Vrb) (5)

Swi = Virb/Vp (6)

Soi = Vo/Vp (7)

Soi + Swi = 1 (8)

Where Φ = porosity (%), Vp = pore volume of Berea core


(mL), Vt = total volume of Bera core (mL), Virb = irreducible
brine saturation (mL), Vo = oil volume occupied in Berea core
during oil saturation (mL), Vib = volume of brine inside Berea
core after brine saturation (mL), Vrb = volume of brine
removed from Berea core during oil saturation (mL), Swi =
initial water/brine saturation (%), Soi = initial oil saturation Figure 4. Particle size distributions for XG−SNP solution systems
(%). (0.1, 0.3, and 0.5 wt % SNPs in 5000 ppm XG) at 30 °C.
Water was then injected to measure the percentage of oil
recovered through water flooding until the oil cut in the particle cluster sizes with variation in the concentration of SNPs
effluent reaches to a value lower than 1%. A chemical (0.1, 0.3, and 0.5 wt %). The nanoparticles underwent
(Polymer/polymer−SNP) slug of 0.5 PV (pore volume) clustering due to agglomeration, because of which the cluster
followed by chase water (1.5 PV) flooding was accomplished size increased to about ∼3700−4700 nm. Agglomeration rate
to evaluate the oil recovery. The optimal slug size of 0.5 PV was and cluster size can be enhanced with an increase in the
chosen based on an earlier screening study that was conducted polymer solution concentration.28 However, Figure 4 suggests
for the optimality of an alkali flooding process.41 that the agglomeration has no impact on the system stability.
6367 DOI: 10.1021/acs.iecr.8b00540
Ind. Eng. Chem. Res. 2018, 57, 6364−6376
Industrial & Engineering Chemistry Research Article

After 10 days, the average particle size of the agglomerated role of divalent and monovalent ions in altering the zeta
nanoparticles marginally varied from 3.95 to 3.73 μm for 0.1 wt potential values at different pH.48
% SNPs and 4.64 to 4.03 μm for 0.5 wt % SNPs. The reduction 3.2. Rheological (Viscosity) Evaluation of Xanthan
in average particle size with time is due to the fact that large Gum−SNP Systems. The variation in the viscosity values with
agglomerated nanoparticles in the solution settle at the bottom shear rate for the XG−SNP system is depicted in Figure 5a (30
and smaller cluster agglomerated nanoparticles are retained in °C) and b (80 °C). The viscosity of the system varied from
the supernatant system. Hence, a marginal reduction in the 2460−2970 mPa s to almost 50 mPa s when the shear rate
average particle size is indicative of a stable aqueous polymer− changed from 1 to 200 s−1. The viscosity was found higher at
SNP system. 28 This confirms negligible sedimentation low shear rate of 1 s−1 and reduced drastically to a smaller
phenomena and better stability of nanoparticles in polymer constant value at higher shear rate (∼200 s−1). A similar trend
phase. Similar phenomena were reported by Sharma et al.28 was also observed by Son et al.49 using SiO2 nanoparticles.
where the average nanoparticle cluster size increased from 3.54 Figure 5c describes the variation in viscosity of the system with
to 4.54 μm with an enhancement in SiO2 nanoparticle temperature for XG and XG−SNP systems at shear rate of 1
concentration from 0.5 to 2 wt % in 1000 ppm polyacrylamide s−1. The initial viscosity observed with XG at 30 °C was 2460
concentration. mPa s (0 wt % SNPs) which enhanced to 2970 mPa s when 0.5
Table 1 summarizes the zeta potential values of the XG− wt % SNPs was introduced. Similar behavior was observed at
SNP solutions. For 5000 ppm XG solution with an SNP elevated temperature (80 °C); the viscosity increased from 593
(0 wt % SNPs) to 771 mPa s (0.5 wt % SNPs). The viscosity of
Table 1. Zeta Potential Values for Variant SNP the system with SNPs was improved even at higher
Concentrations in 5000 ppm Xanthan Gum Sample at 30 °C temperature and reservoir formation water conditions which
silica nanoparticles zeta potential (mV) zeta potential (mV) is probably due to the formation of complex three-dimensional
(wt %) day 1 day 10 macromolecular structures.38 The viscosity enhancement with
0.1 −21.87 −21.35 SNPs in the polymer solution is due to adsorption of polymer
0.3 −25.77 −24.36 on the nanosilica particle surface that is driven by a hydrogen
0.5 −26.52 −25.64 bonding based interaction. At higher temperature, the
adsorption of polymer on silica surface reduces and thus
weakens the hydrogen bonding between polymer and silica and
concentration of 0.1, 0.3, and 0.5 wt %, the zeta potential values deteriorates the three-dimensional network structure.38,50
were evaluated to be −21.87, −25.77, and −26.52 mV Figure 6 presents the storage modulus (G′) and loss modulus
respectively with a standard deviation of <±2%. Negligible (G″) behavior of the samples. The obtained values revealed
variation in the values of zeta potential (∼2−4%) was observed that at a lower initial frequency of 0.01 Hz, G′ is much higher
after 10 days, and hence the XG−SNP solution system can be than G″ for all the concentrations of SNPs. As frequency
considered as partially stable as no rapid coagulation or increased from 0.01 to 0.1 Hz, the storage and loss modulus
flocculation was observed. The reason for not achieving a value reduced drastically and the difference between G′ and G″
of ±30 mV (stable system) could be due to the effect of becomes constant. However, as frequency increased further, the
synthetic formation water/salinity/different ion composition in rate of enhancement of storage modulus was faster than that of
the formation water in addition to the pH of the system. Study the loss modulus and thereby confirms viscoelastic properties of
carried out by Farooq et al. clearly explained the characteristic the system.27 No crossover point existed between G′ and G′

Figure 5. Viscosity and rheological behavior of the XG (5000 ppm) SNP system (a) at 30 and (b) 80 °C and (c) at shear rate of 1 s−1.

6368 DOI: 10.1021/acs.iecr.8b00540


Ind. Eng. Chem. Res. 2018, 57, 6364−6376
Industrial & Engineering Chemistry Research Article

temperature cases. Hence, 0.3 wt % SNPs was regarded to be


optimum for core flooding experiments. The inability to further
lower IFT value with higher SNPs concentration (>0.3 wt %)
could be due to the saturation of SNPs at oil−polymer solution
interface at their optimal concentration.28,55 The reduction in
IFT was better at an elevated temperature for Assam crude oil,
and this is in agreement with the trends presented in the
literature.28 The reduction in IFT value with implementation of
nanoparticles causes easy flow of trapped residual oil, as the
work required to deform the oil reduces which moves the oil
droplet in the pore throat and improves oil recovery.56,57
3.4. Effect of SNPs on Stability of Emulsions.
Emulsification stability studies have been conducted for the
crude oil−XG polymer system in the concentration range of
Figure 6. Storage modulus (G′) and loss modulus (G″) behavior of
1000−5000 ppm. Figure 8a−d shows that while emulsification
the samples at different SNP concentrations. existed for the entire concentration range, its stability varied. At
a lower polymer concentration, due to poor coverage of
which indicates solution behavior as a gel in a bulk phase.51,52 polymer on the oil droplet, the oil droplets undergo extremely
Similar observations and behavior have been reported by fast coalescence to form bigger droplets and thereby emulsion
Kennedy et al.53 The authors indicated viscosity enhancement stability gets reduced.58,59 With increasing polymer concen-
with viscoelastic moduli. The storage and loss modulus trends tration, emulsion stability enhances due to greater adsorption of
infer that SNPs are favorable for EOR at higher temperatures. polymer on the oil droplet, bridging between two oil droplets
3.3. Effect of SNPs on Interfacial Tension of Crude occurs and facilitates flocculation.60 The emulsion formed at
Oil−Aqueous Polymer Systems. The interfacial tension 1000 ppm XG disappeared very quickly (within 30 min).
between crude oil and XG (5000 ppm) SNP solutions was However, the emulsion stability for this case enhanced to 5−6 h
measured at lower (30 ± 2 °C) and higher (70 ± 2 °C) due to the addition of SNPs (Figure 8b). Similar trends were
temperature (Figure 7). In the absence of SNPs, the IFT of found at 3000 and 5000 ppm XG where emulsion stability
enhanced to 18 and more than 25 days respectively (Figure 8c
and d). Therefore, as 5000 ppm XG showed higher
emulsification and better stability period (Figure 8a), the effect
of SNPs (0.1−0.5 wt %) at 5000 ppm polymer concentration
was evaluated. For more than 0.3 wt % SNPs concentration, the
emulsion stability reached a stagnant point and could not be
improved further (Figure 8d). SNPs get adsorbed at the oil−
water interface and facilitate IFT reduction and contribute
toward emulsion stability.27 At 0.3 wt %, SNPs adsorption at
the oil−water interface attained a saturation phase due to which
the stability could not be increased further. Therefore,
considering the cost of polymer and SNPs and the emulsion
stability of the polymer−SNP solution, 5000 ppm XG, and 0.3
wt % SNPs can be inferred to be the optimal values.
The emulsion quality can also be ensured by examining the
microscopic images of the emulsions in terms of average size
and droplet size distributions at various XG and SNPs
concentrations. Figure 9 demonstrates that with increase in
concentration of SNPs, the oil droplet size distributions
Figure 7. Effect of SNP concentration and temperature on the IFT of reduced. The droplet diameter varied from 0 to 22 μm (large
the crude oil−polymer−SNP solution system at 5000 ppm xanthan droplets) in the absence of SNPs and reduced to 0 to 13 μm
gum. (small-moderate droplet) for 0.5 wt % SNPs concentration.
The average diameter values of the droplets for 0, 0.1, 0.3, and
crude oil−XG solution system was found to be 17.8 mN/m at 0.5 wt % SNPs solutions were found to be 7.52, 6.78, 5.97, and
30 °C. After incorporation of SNPs, the IFT of the system 5.12 μm, respectively. Kumar et al.27 also observed an average
reduced to 10.75, 8.67, and 8.54 mN/m as SNP concentration droplet size of 4.8 μm for carboxy methyl cellulose polymer
increased to 0.1, 0.3, and 0.5 wt % respectively. Similar trends (5000 ppm) with SiO2 (0.3 wt %). The reduction in average
exist at higher temperature (70 °C), where the IFT of the droplet diameter or increase in the formation of small droplets
system reduced from 14.64 mN/m (without SNPs) to 8.54, prevents the aggregation of the droplets in which cohesive
6.46, and 6.86 mN/m for 0.1, 0.3, and 0.5 wt % of SNPs forces between droplets reduce due to the presence of silica
concentration, respectively. Several factors such as large surface nanoparticles.21,27 Binks and Whitby61 also detected similar
to volume ratio, higher adsorption, suspension stability, and behavior where the emulsion droplet decreased from 50 to ≤10
hydrophilic−hydrophobic interaction at the oil−water interface μm with increase in nanoparticle concentration. They reported
have contributed toward IFT reduction at higher SNPs that initially the phase breaks into drop fragments of similar
concentrations.54,55 Also, the figure confirms that IFT remains extent irrespective of nanoparticles concentration. However,
constant after 0.3 wt % of SNPs for both lower and higher the retention of reduced droplet size depends upon the rate of
6369 DOI: 10.1021/acs.iecr.8b00540
Ind. Eng. Chem. Res. 2018, 57, 6364−6376
Industrial & Engineering Chemistry Research Article

Figure 8. Emulsification images of crude oil−xanthan gum−SNP systems at 30 °C: (a) 1000−5000 ppm xanthan gum in the absence of SNPs, (b)
1000 ppm xanthan gum and 0.1−0.5 wt % SNPs, (c) 3000 ppm xanthan gum and 0.1−0.5 wt % SNPs, and (d) 5000 ppm xanthan gum and 0.1−0.5
wt % SNPs.

Figure 9. Effect of SNP concentration on emulsion quality and droplet size distributions at 30 °C.

coalesce of the oil droplets which is greatly controlled by the of the emulsion showed that SNPs adsorbed on the surface of
adsorption of nanoparticles on the drop surface. To observe the the oil droplet. The detail structure of SNPs both in clusters
presence of SNPs on the oil−polymer−SNP emulsion, and segregated state were found in FESEM analysis.
microscopic (Figure 10a) and FESEM (Figure 10b) analyses 3.5. Effect of SNPs on Crude Oil−Xanthan Gum
were performed at optimum concentrations. Microscopic image Emulsion (Creaming Index). Creaming index (CI) analysis
6370 DOI: 10.1021/acs.iecr.8b00540
Ind. Eng. Chem. Res. 2018, 57, 6364−6376
Industrial & Engineering Chemistry Research Article

Figure 10. Images of the oil−polymer−SNP emulsion at optimum concentration (5000 ppm XG−0.3 SNPs) denoting deposition of SNPs on the oil
droplets: (a) microscope analysis and (b) FESEM analysis.

Figure 11. Creaming indices of crude oil−xanthan gum−SNP system at (a) 30 and (b) 80 °C.

was carried out to have a better understanding of the emulsion remained constant with further increasing SNPs concentration
stability process. CI values were calculated for 0.1−0.5 wt % to 0.5 wt %. The data obtained suggest that emulsions formed
SNP concentrations (Figure 11). The CI studies involved initial are unstable at higher temperature and low SNPs concentration
shaking of the crude oil−XG−SNP system followed by (≤0.1 wt %) due to high creaming rate. However, a higher
undisturbed conditions for a said time period to evaluate amount of SNPs (0.3−0.5 wt %) at elevated temperature
creaming (separation) out of the aqueous and oil layers. exhibited better performance with enhanced stability of the
Initially (day 0), the XG−SNP solutions were found effective in emulsion to some extent. The creaming rate decreases with
emulsifying the crude oil. The rate of creaming was observed at increase in nanoparticle concentration which reduces the
30 (Figure 11a) and 80 °C (Figure 11b). At lower temperature average droplet diameter of the emulsion. The stability of
(30 °C), the rate of creaming was found slow (day 1) with CI such emulsion depends on the droplet stability which can be
value as 88.6% and 92.9% for 0 and 0.1 wt % SNPs respectively enhanced by continuous dense packing of the adsorbed
(Figure 11a). After a duration of 10 days for the same samples, nanoparticles at the oil−water interface.61
CI reduced to 78.6 for 0% SNPs and 85.7% for 0.1 wt % SNPs. 3.6. Effect of Temperature on Emulsion Viscosity.
However, no signs of creaming were observed at higher SNPs Viscosity enhancement improves sweep efficiency or favorable
concentrations (0.3 and 0.5 wt %) at 30 °C and duration of 10 mobility which facilitates injected aqueous phase to reach
days. At 80 °C, the creaming rate increased enormously on day unswept regions of the reservoir rock to enhance capillary
1 with CI as 7.1% for 0% SNPs and 35.7% for 0.1 wt % SNPs number and oil recovery.62,63 The viscosity is a function of
(Figure 11b). After 10 days, the CI of the sample reduced temperature of the system and such parameter can severely
drastically to 0% and 14.2% for 0 and 0.1 wt % SNPs, affect the oil recovery. Hence the behavior of viscosity at
respectively. For 0.3 wt % SNPs sample (at 80 °C), the CI elevated temperature was observed (Table 2). The viscosity of
reached to 57.1% on day 1 and 14.2% on day 10. The CI values polymer−oil emulsion improved with the addition of nano-
6371 DOI: 10.1021/acs.iecr.8b00540
Ind. Eng. Chem. Res. 2018, 57, 6364−6376
Industrial & Engineering Chemistry Research Article

Table 2. Viscosity of the Emulsions Formed for Crude Oil− which replaces the carboxylate group from oil-wet Berea
Xanthan Gum−SNP Systems at a Shear Rate of 1 s−1 core.39,66,67 Maghzi et al.67 investigated the mechanism
responsible for wettability alteration using silica nanoparticles
polymer concentration (ppm) 5000 5000 5000 5000
in a five spot model. The analysis exposed a formation of strong
SNPs (wt %) 0 0.1 0.3 0.5 hydrogen bonding between silica nanoparticles and water which
viscosity at 30 °C (mPa s) 3256 4473 4770 4919 increases the surface free energy and changes the wettability
viscosity at 80 °C (mPa s) 1894 1965 2177 2524 from oil-wet to water-wet.
3.8. Oil Recovery by Core Flooding Experiments. Lab-
scale flooding experiments were performed using Berea core
particles. The data at different shear rates (supplementary
sample at various nanoparticle concentrations. The data
Figure S1) showed similar behavior as for polymer−SNPs−
obtained from the flooding experiments are summarized in
aqueous solution (Figure 5a and b). The viscosity of XG−
Figure 13 and Table 3. Initially, a recovery factor of 14.5% oil in
SNP−oil emulsion reduced at higher temperature; however, the
values were found good enough to improve sweep efficiency place (OIP) was obtained with only XG (5000 ppm) at 30 °C.
(see section 3.8 and Table 3). The viscosity of emulsion is The oil recovery increased from 16.3% to 20.8% (OIP) with
always greater than that of crude oil which is helpful to spread increasing concentration of SNPs from 0.1 to 0.3 wt %. At 0.5
the chemical into the unswept region.64 To validate the role of wt % of SNPs, the oil recovery factor reduced to 18.51% of
stability (creaming index), droplet size, and viscous nature of OIP. Oil recovery improves due to reduction in IFT, better
the emulsion, flooding experiments were conducted at two emulsion stability, higher viscosity, and wettability alteration.
different temperatures (30 and 80 °C) and residual oil recovery However, SNPs adversely affect the oil recovery factor due to
was calculated (Table 3). However, more detailed inves- the reduction in porosity and permeability.26,28,68 The oil
tigations are required to optimize these parameters which can recovery trend can be further validated by the pressure drop
widely vary with respect to temperature and salinity of the oil curves as shown in Figure 13b. For polymer flooding, lower
field. pressure drop (24.5 psi) was measured which increases with
3.7. Effect of SNPs on Wettability Alteration. Contact increasing SNP concentration due to higher viscosity effect.28
angle measurements were conducted to visualize the effect of The pressure drop was maximum (46.6 psi) at 0.3 wt % SNPs
nanoparticles on wettability alteration for the chosen crude oil- (optimum concentration) and decreased with further increase
polymer (5000 ppm) system (Figure 12). The initial contact in SNPs concentration (0.5 wt %). Greater pressure drop
angle values decrease with increase in SNPs concentration: signifies large formation of oil bank which eventually results in
86.2°, 72.7°, 66.5°, and 50.4° for 0%, 0.1%, 0.3%, and 0.5% higher oil recovery.9,69,70
SNPs, respectively (Figure 12b).65 The variation in the contact At optimal concentration of SNPs (0.3 wt %), the oil
angle with time (until 10 min) for polymer solution (0 wt % recovery reduced marginally with increase in temperature:
SNPs) was not significant (86.2−51.2°) to alter the wettability 20.8% (at 30 °C) to 18.4% (at 80 °C). The pressure drop curve
(remains intermediate oil wet). However, after inducing 0.1 wt also revealed that at 80 °C, the oil bank formation is sufficient
% SNPs, the contact angle decreased from 72.7° to 22.8°. The enough to increase recovery factor. Chen et al.69 inferred that a
rate of change of contact angle was observed to be dominant direct correlation exists with respect to pressure drop and oil
with increasing the SNP concentration, and the lowest value of recovery, i.e., higher oil recovery can be obtained at higher
contact angle (18.84°) was achieved at 0.5 wt % SNPs. At pressure drop. The emulsion so formed under this condition
higher concentration, the electrostatic repulsion force within improves sweep efficiency by reducing viscous fingering/water
the nanoparticles is large which causes the particles to spread channelling. The marginal reduction in oil recovery at higher
along the solid surface decreasing contact angle.65 It was also temperature is probably due to higher creaming rate
observed that the time required to reach a stagnant value (confirmed through Figure 11). FESEM images of Berea core
decreased as the concentration of SNPs increases (approx- before and after flooding experiments (Figure 14) clearly depict
imately 6−7 min, 5−6 min and 4−5 min for 0.1%, 0.3% and that SNPs do get adsorbed on Berea rock surface to form
0.5% SNPs respectively). Thus, the inclusion of SNPs in the clusters. This is also in agreement with DLS characterization
system alter the system wettability from intermediate oil wet to (Figure 4).
strongly water wet. This enables release of the trapped residual Table 4 presents a summary of the results obtained in this
oil from the pores of the reservoir rock by increasing the work along with the literature findings for nanoparticles assisted
capillary number and hence promotes oil recovery. The chemical EOR systems. It can be observed that nanoparticles
wettability alteration mechanism in the present study is very are effective in enhancing residual oil recovery. Sharma et al.28
likely due to the adsorption of SNPs on the oil-wet surface could increase the residual oil recovery using SNPs and the

Table 3. Cumulative Oil Recovery Data Obtained for Core Flooding Experiments Conducted with Xanthan Gum−SNP
Solutions

polymer
Brine xanthan gum nanoparticles slug initial oil water flooding flooding cumulative total
sr. porosity permeability concentration concentration size saturation temperature recovery recovery recovery
no. (%) (mD) (ppm) (wt %) (PV) (%) (°C) (% OIP) (% OIP) (% OIP)
1 26.0 927 5000 0 0.5 78.3 30 31.84 14.47 46.32
2 25.8 834 5000 0.1 0.5 79.6 30 32.12 16.29 48.41
3 26.5 1002 5000 0.3 0.5 80.1 30 33.31 20.82 54.13
4 25.1 746 5000 0.5 0.5 80.5 30 31.17 18.51 49.68
5 25.3 772 5000 0.3 0.5 80.2 80 32.50 18.44 50.94

6372 DOI: 10.1021/acs.iecr.8b00540


Ind. Eng. Chem. Res. 2018, 57, 6364−6376
Industrial & Engineering Chemistry Research Article

Figure 12. (a) Variation of contact angle with time for XG (5000 ppm) SNP system. (b) Images at initial time = 0.

Figure 13. Core flooding studies conducted with optimal polymer (5000 ppm xanthan gum) concentration and 0.1−0.5 wt % SNP concentration:
(a) cumulative oil recovery and (b) pressure drop profile.

Figure 14. FESEM images of Berea core (Idaho gray) before and after core flooding experiments (chemical composition 5000 ppm xanthan gum
and 0.3 wt % SNPs).

recovery was not affected with temperature (until 90 °C). They found that TiO2 nanoparticles could enhance the oil recovery
also confirmed the combined role of IFT reduction and up to 43.3% IOIP. Cheraghian72 and Maghzi et al.25 also
wettability alteration to recover residual oil. However, detailed inferred similar behavior of residual oil recovery enhancement
investigation of the emulsion formed, its stability, and flow with SNPs. However, additional investigations with respect to
properties have not been considered. Yousefvand and Jafari71 contact angle, IFT studies, emulsion formation, and its stability
6373 DOI: 10.1021/acs.iecr.8b00540
Ind. Eng. Chem. Res. 2018, 57, 6364−6376
Industrial & Engineering Chemistry Research Article

Table 4. Data Summary of Best Results Obtained in This Work and Those Reported in the Literature for Nanoparticle Induced
Polymer Enhanced Oil Recovery
tertiary
concentration concentration recovery cumulative wettability IFT
reference rock chemicals (ppm) nanoparticles (wt %) (%) recovery (%) alteration reduction
Sharma et al.28 sandpack polyacrylamide 1000 silica dioxide 1.5 17.53 63.12 yes yes
(PAM)
Yousefvand sandstone hydrolyzed 3150 titanium 2.3 43.3 52.8
and Jafari71 polyacrylamide dioxide
(HPAM)
Cheraghian72 micromodel HPAM 800 silica dioxide 0.5 18.37 35.0
Maghzi et al.25 micromodel PAM 3000 silica dioxide 0.1 68.0
Ehsan et al.73 sandstone SDS 2150 19.74 72.55 yes yes
core
Ehsan et al.73 sandstone SDS 2150 silica dioxide 0.1 26.95 79.76 yes yes
core
present work Berea core xanthan gum 5000 silica dioxide 0.3 20.82 54.13 yes yes

were not considered in the context of their role in polymer Viscosity vs shear rate curves of emulsions produced at
based EOR. Ehsan et al.73 achieved 26.95% OIP with different nanoparticle concentration and temperature
surfactant−nanoparticles compared to sodium dodecyl sulfate (PDF)


surfactant alone (19.74%). Nanoparticles (SiO2) stimulate IFT
reduction and lessen contact angle to achieve a strong water-
wet reservoir. Detailed investigation considering other mech-
AUTHOR INFORMATION
anism such as emulsion stability, creaming rate, rheology, etc., Corresponding Author
have not been reported. Considering these, this work attempts *Tel.: +91 361 2582263. Fax: +91 361 2582291. E-mail
to provide a greater insight into the role of these complex address: pankaj.tiwari@iitg.ernet.in (P.T.).
interactions and phenomena for enhanced oil recovery of crude ORCID
oil−polymer−SNP systems. Pankaj Tiwari: 0000-0003-2578-3462
Notes
4. CONCLUSIONS
The authors declare no competing financial interest.


The present work provides several additional insights with
respect to nanoparticle assisted polymer flooding of heavy ACKNOWLEDGMENTS
crude oil systems. An optimal combination of polymer (5000
The authors sincerely acknowledge the financial grant provided
ppm) and SNPs (0.3 wt %) was achieved based on SNP
by Department of Science and Technology, Government of
stability, rheology behavior, IFT reduction, emulsification,
India (Grant No: SB/S3/CE/057/2013), for the research
creaming rate, wettability alteration, and core flooding
carried out. Authors also acknowledge the Centre of Excellence
experiments. Stability analysis by DLS revealed the stable
on Sustainable Polymer and Central Instruments Facility at
nature of SNPs in polymer solution but not in aqueous phase
Indian Institute of Technology Guwahati for providing
(formation water). The viscosity of the polymer solution
analytical facilities.


improved when SNPs were introduced in the system and such
behavior retained even at elevated temperature. Similar
favorable behavior was observed for IFT reduction phenomena REFERENCES
too. The reduction in IFT promoted stable oil−water (1) Pye, D. J. Improved secondary recovery by control of water
emulsions by reducing the droplet sizes and size distributions mobility. JPT, J. Pet. Technol. 1964, 16 (8), 911−916.
(7.52−5.12 μm). Creaming index values exposed the role of (2) Sandiford, B. B. Laboratory and field studies of water floods using
polymer solutions to increase oil recoveries. JPT, J. Pet. Technol. 1964,
SNPs in effectively improving the emulsion stability. SNPs also
16 (8), 917−922.
altered the wettability of the chosen system from an (3) Akiyama, E.; Kashimoto, A.; Fukuda, K.; Hotta, H.; Suzuki, T.;
intermediate oil-wet to completely water-wet state. Pressure Kitsuki, T. Thickening properties and emulsification mechanisms of
drop data indicated the formation of oil bank which resulted new derivatives of polysaccharides in aqueous solution. J. Colloid
additional oil recovery of 20.82% at 30 °C and 18.44% at 80 °C Interface Sci. 2005, 282, 448−457.
during SNP assisted polymer flooding. Overall, the SNP− (4) Akiyama, E.; Yamamoto, T.; Yago, Y.; Hotta, H.; Ihara, t.; Kitsuki,
polymer solution was found highly effective for EOR T. Thickening properties and emulsification mechanisms of new
applications in moderate to higher temperature reservoirs. derivatives of polysaccharide in aqueous solution 2. The effect of the
Further insights with respect to SNP induced polymer EOR can substitution ratio of hydrophobic/hydrophilic moieties. J. Colloid
be obtained by investigating variations in oil to chemical ratio, Interface Sci. 2007, 311, 438−446.
(5) Seright, R. S.; Fan, T. G.; Wavrik, K.; Wan, H.; Gaillard, N.;
detailed rheological studies, salinity/pH effect on zeta potential,
Favero, C. Rheology of a new sulfonic associative polymer in porous
techno-economic analysis, and cost analysis. media. SPE Reservoir Eval. Eng. 2011, 14, 726−734.


*
ASSOCIATED CONTENT
S Supporting Information
(6) Lu, Y.; Kang, W.; Jiang, J.; Chen, J.; Xu, D.; Zhang, P.; Zhang, L.;
Feng, H.; Wu, H. Study on the stabilization mechanism of crude oil
emulsion with an amphiphilic polymer using the b-cyclodextrin
inclusion method. RSC Adv. 2017, 7, 8156−8166.
The Supporting Information is available free of charge on the (7) Jewett, R. L.; Schurz, G. F. Polymer flooding-a current appraisal.
ACS Publications website at DOI: 10.1021/acs.iecr.8b00540. JPT, J. Pet. Technol. 1970, 22 (6), 675−684.

6374 DOI: 10.1021/acs.iecr.8b00540


Ind. Eng. Chem. Res. 2018, 57, 6364−6376
Industrial & Engineering Chemistry Research Article

(8) Chang, H. L. Polymer flooding technology yesterday, today, and food-grade particles to form stable o/w and w/o pickering emulsions.
tomorrow. JPT, J. Pet. Technol. 1978, 30 (08), 1113−1128. J. Colloid Interface Sci. 2016, 473 (1), 9−21.
(9) Pei, H.; Zhang, G.; Ge, J.; Zhang, L.; Wang, H. Effect of polymer (30) Guo, Z.; Dong, M.; Chen, Z.; Yao, J. A fast and effective method
on the interaction of alkali with heavy oil and its use in improving oil to evaluate the polymer flooding potential for heavy oil reservoirs in
recovery. Colloids Surf., A 2014, 446, 57−64. western canada. J. Pet. Sci. Eng. 2013, 112, 335−340.
(10) Demin, W.; Zhang, Z.; Chun, L.; Cheng, J.; Du, X.; Li, Q. A (31) Wang, J.; Dong, M. Optimum effective viscosity of polymer
pilot for polymer flooding of saertu formation s ii 10−16 in the north solution for improving heavy oil recovery. J. Pet. Sci. Eng. 2009, 67,
of daqing oil field. SPE Asia Pacific Oil and Gas Conference, Adelaide, 155−158.
Australia, 28−31 October; Society of Petroleum Engineers, 1996; pp (32) Esfandyari Bayat, A. E.; Junin, R.; Samsuri, A.; Piroozian, A.;
431−441. Hokmabadi, M. Impact of metal oxide nanoparticles on enhanced oil
(11) Saboorian-Jooybari, H.; Dejam, M.; Chen, Z. Heavy oil polymer recovery from limestone media at several temperatures. Energy Fuels
flooding from laboratory core floods to pilot tests and field 2014, 28, 6255−6266.
applications: Half-century studies. J. Pet. Sci. Eng. 2016, 142, 85−100. (33) Zallaghi, M.; Kharrat, R.; Hashemi, A. Improving the
(12) Leung, W. M.; Axelson, D. E.; Van Dyke, J. D. Thermal microscopic sweep efficiency of water flooding using silica nano-
degradation of polyacrylamide and poly (acrylamide-coacrylate). J. particles. J. Pet. Explor. Prod. Technol. 2018, 8, 259−269.
Polym. Sci., Part A: Polym. Chem. 1987, 25, 1825−1846. (34) Karimi, A.; Fakhroueian, Z.; Bahramian, A.; Pour Khiabani, N.
(13) Yang, M. H. Rheological behavior of polyacrylamide solution. J. P.; Darabad, J. B.; Azin, R.; Arya, S. Wettability alteration in carbonates
Polym. Eng. 1999, 19 (5), 371−381. using zirconium oxide nanofluids: Eor implications. Energy Fuels 2012,
(14) Kheradmand, H.; Francois, J.; Plazanet, V. Hydrolysis of 26, 1028−1036.
polyacrylamide and acrylic acid-acrylamide copolymers at neutral ph (35) Maghzi, A.; Mohebbi, A.; Kharrat, R.; Ghazanfari, M. H. Pore-
and high temperature. Polymer 1988, 29 (5), 860−870. scale monitoring of wettability alteration by silica nanoparticles during
(15) Chen, Q.; Wang, Y.; Lu, Z.; Feng, Y. Thermoviscosifying polymer flooding to heavy oil in a five-spot glass micromodel. Transp.
polymer used for enhanced oil recovery: Rheological behaviors and Porous Media 2011, 87, 653−664.
core flooding test. Polym. Bull. 2013, 70, 391−401. (36) Otsubo, Y.; Umeya, K. Rheological properties of silica
(16) Algharaib, M.; Alajmi, A.; Gharbi, R. Improving polymer flood suspensions in polyacrylamide solutions. J. Rheol. 1984, 28, 95−108.
performance in high salinity reservoirs. J. Pet. Sci. Eng. 2014, 115, 17− (37) Kamibayashi, M.; Ogura, H.; Otsubo, Y. Viscosity behavior of
23. silica suspensions flocculated by associating polymers. J. Colloid
(17) Muller, G. Thermal stability of high molecular weight Interface Sci. 2005, 290, 592−597.
polyacrylamide aqueous solution. Polym. Bull. 1981, 5, 31−37. (38) Maurya, N. K.; Mandal, A. Studies on behavior of suspension of
(18) Abu-Sharkh, B. F.; Yahaya, G. O.; Ali, S. A.; Hamad, E. Z.; Abu- silica nanoparticle in aqueous polyacrylamide solution for application
Reesh, I. M. Viscosity behavior and surface and interfacial activities of in enhanced oil recovery. Pet. Sci. Technol. 2016, 34 (5), 429−436.
(39) Dehghan Monfared, A.; Ghazanfari, M. H.; Jamialahmadi, M.;
hydrophobically modified water-soluble acrylamide/n-phenyl acryl-
Helalizadeh, A. Potential application of silica nanoparticles for
amide block copolymers. J. Appl. Polym. Sci. 2003, 89, 2290−2300.
wettability alteration of oil−wet calcite: A mechanistic study. Energy
(19) Han, D.-K.; Yang, C.-Z.; Zhang, Z.-Q.; Lou, Z.-H.; Chang, Y.-I.
Fuels 2016, 30, 3947−3961.
Recent development of enhanced oil recovery in china. J. Pet. Sci. Eng.
(40) Saha, R.; Uppaluri, V. S. R.; Tiwari, P. Effect of mineralogy on
1999, 22 (1), 181−188.
the adsorption characteristics of surfactant − reservoir rock system.
(20) Kamal, M. S.; Sultan, A. S.; Al-Mubaiyedh, U. A.; Hussein, I. A.
Colloids Surf., A 2017, 531, 121−132.
Review on polymer flooding: Rheology, adsorption, stability, and field
(41) Saha, R.; Uppaluri, R. V. S.; Tiwari, P. Influence of
applications of various polymer systems. Polym. Rev. 2015, 55, 491− emulsification, interfacial tension, wettability alteration and saponifi-
530. cation on residual oil recovery by alkali flooding. J. Ind. Eng. Chem.
(21) Sharma, T.; Kumar, G. S.; Chon, B. H.; Sangwai, J. S. Thermal 2018, 59, 286−296.
stability of oil-in-water pickering emulsion in the presence of (42) Yu, W.; Xie, H. A review on nanofluids: Preparation, stability
nanoparticle, surfactant, and polymer. J. Ind. Eng. Chem. 2015, 22, mechanisms, and applications. J. Nanomater. 2012, 2012, 1−17.
324−334. (43) Al-Anssari, S.; Barifcani, A.; Wang, S.; Maxim, L.; Iglauer, S.
(22) Binks, B. P.; Rocher, A. Effects of temperature on water-in-oil Wettability alteration of oil-wet carbonate by silica nanofluid. J. Colloid
emulsions stabilised solely by wax microparticles. J. Colloid Interface Sci. Interface Sci. 2016, 461, 435−442.
2009, 335, 94−104. (44) Yuan, C. D.; Pu, W. F.; Wang, X. C.; Sun, L.; Zhang, Y. C.;
(23) Hendraningrat, L.; Li, S.; Torsaeter, O. A coreflood investigation Cheng, S. Effects of interfacial tension, emulsification, and surfactant
of nanofluid enhanced oil recovery. J. Pet. Sci. Eng. 2013, 111, 128− concentration on oil recovery in surfactant flooding process for high
138. temperature and high salinity reservoirs. Energy Fuels 2015, 29, 6165−
(24) Yousefvand, H.; Jafari, A. Enhanced oil recovery using polymer/ 6176.
nanosilica. Procedia Mater. Sci. 2015, 11, 565−570. (45) Keowmaneechai, E.; McClements, D. J. Influence of edta and
(25) Maghzi, A.; Kharrat, R.; Mohebbi, A.; Ghazanfari, M. H. The citrate on physicochemical properties of whey protein-stabilized oil-in-
impact of silica nanoparticles on the performance of polymer solution water emulsions containing cacl2. J. Agric. Food Chem. 2002, 50, 7145−
in presence of salts in polymer flooding for heavy oil recovery. Fuel 7153.
2014, 123, 123−132. (46) Ershadi, M.; Alaei, M.; Rashidi, A. R. A; Khosravani, S.;
(26) Youssif, M. I.; El-Maghraby, R. M.; Saleh, S. M.; Elgibaly, A. Ramazani, A. Carbonate and sandstone reservoirs wettability improve-
Silica nanofluid flooding for enhanced oil recovery in sandstone rocks. ment without using surfactants for chemical enhanced oil recovery (c-
Egypt. J. Pet. 2018, 27, 105−110. eor). Fuel 2015, 153, 408−415.
(27) Kumar, N.; Gaur, T.; Mandal, A. Characterization of spn (47) Saha, R.; Sharma, A.; Uppaluri, R. V. S.; Tiwari, P. Interfacial
pickering emulsions for application in enhanced oil recovery. J. Ind. interaction and emulsification of crude oil to enhance oil recovery. Int.
Eng. Chem. 2017, 54, 304−315. J. Oil, Gas Coal Technol. 2018, 1, 1.
(28) Sharma, T.; Iglauer, S.; Sangwai, J. S. Silica nanofluids in an (48) Farooq, U.; Tweheyo, M. T.; Sjöblom, J.; Øye, G. Surface
oilfield polymer polyacrylamide: Interfacial properties, wettability characterization of model, outcrop, and reservoir samples in low
alteration, and applications for chemical enhanced oil recovery. Ind. salinity aqueous solutions. J. Dispersion Sci. Technol. 2011, 32 (4),
Eng. Chem. Res. 2016, 55, 12387−12397. 519−531.
(29) Duffus, L. J.; Norton, J. E.; Smith, P.; Norton, I. T.; (49) Son, H. A.; Yoon, K. Y.; Lee, G. J.; Cho, J. W.; Choi, S. K.; Kim,
Spyropoulos, F. A comparative study on the capacity of a range of J. W.; Im, K. C.; Kim, H. T.; Lee, K. S.; Sung, W. M. The potential

6375 DOI: 10.1021/acs.iecr.8b00540


Ind. Eng. Chem. Res. 2018, 57, 6364−6376
Industrial & Engineering Chemistry Research Article

applications in oil recovery with silica nanoparticle and polyvinyl (69) Chen, L.; Zhang, G.; Ge, J.; Jiang, P.; Tang, J.; Liu, Y. Research
alcohol stabilized emulsion. J. Pet. Sci. Eng. 2015, 126, 152−161. of the heavy oil displacement mechanism by using alkaline/surfactant
(50) Wisniewska, M. The temperature effect on the adsorption flooding system. Colloids Surf., A 2013, 434, 63−71.
mechanism of polyacrylamide on the silica surface and its stability. (70) Pei, H.; Zhang, G.; Ge, J.; Jin, L.; Ma, C. Potential of alkaline
Appl. Surf. Sci. 2012, 258 (7), 3094−3101. flooding to enhance heavy oil recovery through water-in-oil
(51) Almdal, K.; Dyre, J.; Hvidt, S.; Kramer, O. Towards a emulsification. Fuel 2013, 104, 284−293.
phenomenological definition of the term ‘gel’. Polym. Gels Networks (71) Yousefvand, H.; Jafari, A. Enhanced oil recovery using polymer/
1993, 1 (1), 5−17. nanosilica. Procedia Mater. Sci. 2015, 11, 565−570.
(52) Kang, W.; Xu, B.; Wang, Y.; Li, Y.; Shan, X.; An, F.; Liu, J. (72) Cheraghian, G. Effect of nano titanium dioxide on heavy oil
Stability mechanism of w/o crude oil emulsion stabilized by polymer recovery during polymer flooding. Pet. Sci. Technol. 2016, 34 (7), 633−
and surfactant. Colloids Surf., A 2011, 384, 555−560. 641.
(53) Kennedy, J. R. M.; Kent, K. E.; Brown, J. R. Rheology of (73) Eshraghi S, E.; Y, K.; M, Q.; A, K. Investigating effect of sio2
dispersions of xanthan gum, locust bean gum and mixed biopolymer nanoparticle and sodium-dodecyl-sulfate surfactant on surface proper-
ties: Wettability alteration and ift reduction. J. Pet. Environ. Biotechnol.
gel with silicon dioxide nanoparticles. Mater. Sci. Eng., C 2015, 48,
2017, 8 (6), 1−5.
347−353.
(54) Kazemzadeh, Y.; Eshraghi, E. S.; Kazemi, K.; Sourani, S.;
Mehrabi, M.; Ahmadi, Y. Behavior of asphaltene adsorption onto the
metal oxide nanoparticle surface and its effect on heavy oil recovery.
Ind. Eng. Chem. Res. 2015, 54, 233−239.
(55) Roustaei, A.; Saffarzadeh, S.; Mohammadi, M. An evaluation of
modified silica nanoparticles’ efficiency in enhancing oil recovery of
light and intermediate oil reservoirs. Egypt. J. Pet. 2013, 22 (3), 427−
433.
(56) Donaldson, E. C.; Chilingarian, G. V.; Yen, T. F. Enhanced oil
recovery ii, processes and operations; Elsevier Science Publ. Co.: New
York, 1989.
(57) Samanta, A.; Ojha, K.; Sarkar, A.; Mandal, A. Surfactant and
surfactant-polymer flooding for enhanced oil recovery. Adv. Pet. Explor.
Dev. 2011, 2, 13−18.
(58) Cao, Y. H.; Dickinson, E.; Wedlock, D. J. Creaming and
flocculation in emulsions containing polysaccharide. Food Hydro-
colloids 1990, 4, 185−195.
(59) Dickinson, E.; Ma, J. G.; Povey, M. J. W. Creaming of
concentrated oil in-water emulsions containing xanthan. Food
Hydrocolloids 1994, 8 (55), 481−497.
(60) Ghosh, A. K.; Bandyopadhyay, P. Polysaccharide-protein
interactions and their relevance in food colloids. In The complex
world of polysaccharide; Intech open science open minds, 2012; pp
395−408.
(61) Binks, B. P.; Whitby, C. p. Silica particle-stabilized emulsions of
silicone oil and water: Aspects of emulsification. Langmuir 2004, 20,
1130−1137.
(62) Lake, L. W. Enhanced oil recovery; Society of Petroleum
Engineers, 2010.
(63) Iglauer, S.; Sarmadivaleh, M.; Geng, C.; Lebedev, M. In-situ
residual oil saturation and cluster size distribution in sandstones after
surfactant and polymer flooding imaged with x-ray micro-computed
tomography. In International Petroleum Technology Conference, Doha,
Qatar, January 19−22, 2014; IPTC-17312-MS.
(64) Arhuoma, M.; Dong, M.; Yang, D.; Idem, R. Determination of
water-in-oil emulsion viscosity in porous media. Ind. Eng. Chem. Res.
2009, 48, 7092−7102.
(65) Li, S.; Hendraningrat, L.; Torsaeter, O. Improved oil recovery
by hydrophilic silica nanoparticles suspension: 2 phase flow
experimental studies. In International Petroleum Technology Conference,
2013; pp 1−15, IPTC-16707.
(66) Dehghan Monfared, A. D.; Ghazanfari, M. H.; Jamialahmadi,
M.; Helalizadeh, A. Adsorption of silica nanoparticles onto calcite:
Equilibrium, kinetic, thermodynamic and dlvo analysis. Chem. Eng. J.
2015, 281 (1), 334−344.
(67) Maghzi, A.; Mohammadi, S.; Ghazanfari, M. H.; Kharrat, R.;
Masihi, M. Monitoring wettability alteration by silica nanoparticles
during water flooding to heavy oils in five-spot systems: A pore-level
investigation. Exp. Therm. Fluid Sci. 2012, 40, 168−176.
(68) Ehtesabi, H.; Ahadian, M. M.; Taghikhani, V.; Ghazanfari, M. H.
Enhanced heavy oil recovery in sandstone cores using tio2 nanofluids.
Energy Fuels 2014, 28, 423−430.

6376 DOI: 10.1021/acs.iecr.8b00540


Ind. Eng. Chem. Res. 2018, 57, 6364−6376

Potrebbero piacerti anche