You are on page 1of 13

Fuel 252 (2019) 622–634

Contents lists available at ScienceDirect

Fuel
journal homepage: www.elsevier.com/locate/fuel

Full Length Article

Heavy oil recovery by surface modified silica nanoparticle/HPAM T


nanofluids
⁎ ⁎
Laura M. Corredor1, Ehsan Aliabadian1, Maen Husein , Zhangxin Chen, Brij Maini ,

Uttandaraman Sundararaj
Department of Chemical and Petroleum Engineering, University of Calgary, 2500 University Dr NW, Calgary T2N 1N4, Canada

A R T I C LE I N FO A B S T R A C T

Keywords: Recent studies showed that the presence of dispersed nanoparticles (NPs) can increase the efficiency of polymer
Linear and nonlinear rheology flooding. The network structure of polymer/NP hybrid dispersion has a significant impact on oil recovery. In this
Enhanced oil recovery (EOR) work, the surface of SiO2 NPs was modified by chemical grafting of octyltriethoxysilane (SiO2-OTES), oleic acid
Flow in porous media (SiO2-OAA) and stearic acid (SiO2-SAA) in an attempt to stimulate higher degree of interaction with partially
Modified nanosilica particles
hydrolyzed polyacrylamide (HPAM). Thermogravimetric analysis (TGA), differential scanning calorimetry
Polymer flooding
Hydrolyzed polyacrylamide
(DSC), and Fourier transform infrared (FTIR) spectroscopy were employed to characterize the modified silica
NPs. In addition, ζ-potential, cryo-scanning electron microscopy (cryo-SEM), and linear and nonlinear rheology
were used to evaluate the characteristics of the hybrid dispersion of HPAM/unmodified and modified SiO2 NPs.
ζ-potential measurements showed that the addition of HPAM improved the colloidal stability of the modified and
unmodified SiO2 NPs dispersed in deionized (DI) water. Moreover, the addition of HPAM reduced the size of the
NP aggregates by effective steric repulsion. Small and large shear oscillatory deformation results showed that
SiO2-OTES NPs improved the HPAM network significantly. Nevertheless, HPAM/NPs network formed with all
the different SiO2 NPs severely decreased the intra-cycle shear-thickening behavior of the polymer. Lastly, the
addition of 0.2 wt% SiO2-OTES NPs to the HPAM solution increased the ultimate oil recovery from 71.4% to
75.7% OOIP.

1. Introduction achieved by altering the silanol groups on the surface of the SiO2 NPs
via attaching an organic functionality, either by physical association or
Two classes of polymers are primarily used in polymer flooding; chemical bonding. The principle of the physical modification method is
namely synthetic polymers and biopolymers [1]. Partially hydrolyzed the preferential adsorption of a polar group of a macromolecule on the
polyacrylamide (HPAM) is the most commonly used synthetic polymer silica surface through hydrogen-bonding or electrostatic interaction.
to date [2]. HPAM solutions often suffer from viscosity loss due to harsh SiO2 NPs have been modified by physical methods with fatty acids such
reservoir conditions; including high temperature, high pressure, and as stearic acid [15], and oleic acid [16,17], and with cationic salts such
presence of chemical substances such as divalent cations and clay [3]. as cetyltrimethylammonium bromide [18]. Modification of the surface
Xanthan gum (XG) is the most widely used biopolymer [4]. XG has been by chemical reactions has received more attention, since it typically
proposed as an alternative to HPAM due to its excellent tolerance to contributes to much stronger interaction between the modifiers and the
mechanical shearing and high salinity and temperature. Nevertheless, nanosilica particles. Chemical methods involve modification either by
XG use is relatively limited compared with HPAM due to its higher cost, grafting polymers or modifying agents on the nanosilica surface
higher susceptibility to biodegradation and lower availability [5–7]. [19,20]. The resultant nanocomposites have been extensively studied
Recent studies revealed higher polymer flooding performance upon for applications such as adhesives, biomaterials, protective coatings,
adding silica nanoparticles (NPs) to polyacrylamide solutions for microelectronics and proton exchange membranes [21]. However, their
medium and heavy oil recovery [8–12]. The best performance of these application for EOR processes has been limited.
hybrid dispersions is obtained when the NPs are uniformly dispersed in Ponnapati et al. [22] synthesized SiO2-ethylene oxide-based
the polymer solution [13,14]. Higher extent of dispersion can be polymer/NPs hybrids using a “grafting from” method. They reported


Corresponding authors.
E-mail addresses: mhusein@ucalgary.ca (M. Husein), bmaini@ucalgary.ca (B. Maini), u.sundararaj@ucalgary.ca (U. Sundararaj).
1
L.C. and E.A. made equal contribution to this work.

https://doi.org/10.1016/j.fuel.2019.04.145
Received 13 February 2019; Received in revised form 22 April 2019; Accepted 26 April 2019
Available online 06 May 2019
0016-2361/ © 2019 Elsevier Ltd. All rights reserved.
L.M. Corredor, et al. Fuel 252 (2019) 622–634

Nomenclature S Intra-cycle strain-stiffening ratio


SA Stearic acid
ASNP zwitterionic poly[2-methacryloyloxyethylphosphorylcho- SAOS Small amplitude oscillatory shear
line (MPC) – co – divinylbenzene (DVB)] shell-coated si- SEM Scanning electron microscopy
lica nanoparticles SiO2-OAA Silica modified with oleic acid-Method A
DSC Differential scanning calorimetry SiO2-OTES Silica modified with Octyltriethoxysilane
EOR Enhanced oil recovery SiO2-SAA Silica modified with stearic acid-Method A
FTIR Fourier transform infrared T Intra-cycle shear-thickening ratio
HPAM Hydrolyzed polyacrylamide WF Waterflooding
KBr Potassium Bromide G' Storage modulus
LAOS Large amplitude oscillatory shear G' Loss modulus
'
LVR Linear viscoelastic region GM Minimum-strain modulus
MW Molecular weight GL' Large-strain modulus
OA Oleic acid ω Angular frequency
OOIP Original oil in place γ Strain
OTES Octyltriethoxysilane ηM' Minimum-strain rate viscosity
PV Pore Volume ηL' Large-strain rate viscosity
PF Polymer flooding

that the injection of the 0.5 wt% nanohybrid could mobilize residual oil wide range of flow conditions can be simulated. LAOS has been applied
in the core, yielding 7.9% of the OOIP. Subsequent injection of brine in polymer nanocomposites [56,57], polymer solutions [24,58], and
yielded an additional 11% of the OOIP. The enhanced performance of colloidal gels [59,60] to link nonlinear viscoelastic behavior to micro-
the modified NPs resulted from polymer-grafted NPs clogging some structural evolution under nonlinear conditions.
water-occupied pore throats, which increased the local pressure, and In the present work, three different fields (chemistry, rheology, and
hence mobilized the trapped oil in other pores. In another work, Choi petroleum science) have been merged to show how to increase oil re-
et al. [23] studied nanofluid-enhanced oil recovery using hydrophobic covery, which distinguishes this study from other works in the litera-
zwitterionic poly[2-methacryloyloxyethylphosphorylcholine (MPC) – ture. In phase one, bare SiO2 NPs were chemically modified with car-
co – divinylbenzene (DVB)] shell-coated silica nanoparticles (ASNP). boxylic acids (SiO2-SAA and SiO2-OAA) and silane (SiO2-OTES) to
The results showed that both viscous and storage modulus of ASNP generate three differently modified SiO2 NPs. By modifying the SiO2
nanofluid were approximately 2 orders of magnitude higher than that NPs surface, we expected to (i) increase the hydrophobicity of the NPs
of the bare silica nanofluid. The core flooding experiments showed that and, hence, enhance their dispersivity into the polymer solution; and
the oil recovery was 74.1%, which was higher than water injection (ii) tailor the interaction between the treated NPs and the polymer to
(68.9%) and unmodified NP injection (72.7%). In our recent works tailor the mobility of the polymer solution and, consequently, increase
[24,25], we showed that the slightly hydrophobic fumed SiO2 NPs the oil recovery. In phase two, the network structure of the hybrid
(AEROSIL R816) are more efficient at forming a stronger network dispersions of HPAM/SiO2 NPs was studied using rheology. The effects
structure with HPAM molecules than the completely hydrophilic fumed of adding modified SiO2 NPs on the linear and nonlinear viscoelastic
SiO2 NPs (AEROSIL 300). Moreover, we showed that the surface properties were analyzed using frequency sweep test and intra-cycle
chemistry of the NPs plays a key role in the formation of 3-D elastic shear-thickening behavior, respectively. In phase three, these different
structure in the polymeric nanofluids, independent of fumed SiO2 NP hybrid dispersions were injected into a linear sand-pack to evaluate
aggregate size. their effectiveness in tertiary recovery. In our previous work [61], we
In porous media with a wide range of pore size distribution and studied the flow curves of the SiO2-SAA, SiO2-OAA and SiO2-OTES
geometry, hybrid dispersions may undergo various flow conditions HPAM/XG hybrid dispersions and developed a model based on a mul-
[26,27]. To optimize the efficiency of the polymer flooding in EOR, tilayer perceptron (MLP) neural network to predict the viscosity of the
understanding the flow behavior of such systems in porous media is of dispersions. However, that study did not include the analysis of the
great importance. Rheometry, which deals with the deformation and linear and nonlinear viscoelastic properties of the dispersions, or their
flow behavior of materials, is a robust technique that provides a valu- performance in displacing oil.
able opportunity to mimic both small and large deformation in porous To the best of our knowledge, there is no work in the literature that
media. Rheometry can capture viscoelastic properties that are affected systematically synthesized the modified SiO2 NPs, evaluated the effect
by both inherent properties of a polymer molecule and external para- of hydrophobicity of the modified SiO2 NPs on the linear and nonlinear
meters. Molecular weight [28], degree of hydrolysis [29,30], and de- viscoelastic behavior and the network structure of an aqueous HPAM
gree of branching [31] are some examples of inherent properties of a solution, and tested the performance of the HPAM/NP hybrid disper-
polymer molecule. Temperature [25,32,33], salt content [25,34–36], sions for heavy oil recovery as reported here.
and shear/extensional deformation [3,37,38] are examples of external
parameters that can significantly impact the viscoelastic behavior of
such systems. 2. Materials and methods
Small amplitude oscillatory shear (SAOS) flow [3,39–42], flow
curve [43–45], and extensional (shear-free) tests [46,47] are common 2.1. Materials
methods used to investigate the network structure of polymer solutions.
However, the applicability of these techniques is restricted due to some Fumed silica powder (7 nm) was obtained from Sigma-Aldrich
limitations like flow instabilities [48], shear inhomogeneity [49], and (USA). The chemicals used for the surface modification of the SiO2 NPs
low viscosity of solution systems. Recently, large amplitude oscillatory were oleic acid (C18H34O2, 90%), stearic acid (C18H36O2, 95%), and
shear (LAOS) test has been employed to assess the viscoelastic prop- octyl(triethoxy)silane (OTES, ≥97.5%) which were obtained from
erties of different polymer solutions [50–55]. In LAOS, the input strain Sigma-Aldrich (USA). The solvents used in the reactions were: ethanol
amplitude and frequency can be controlled independently and, hence, a (EtOH, 99%) and hydrochloric acid (ACS reagent, 37%), both obtained
from Sigma-Aldrich (USA); and cyclohexane (99.5%) and ammonium

623
L.M. Corredor, et al. Fuel 252 (2019) 622–634

hydroxide (28–30 wt% solutions of NH3 in water), which were obtained measurement, the samples were pretreated by ultrasonication for 5 min.
from Fisher Scientific (USA). The detailed synthesis of the surface A digital pH meter (model AB 15 plus, Fisher Scientific, USA) was used
modified SiO2 NPs used in this work has been reported previously by to measure the pH of the samples at 20 °C. The uncertainty of the pH
the authors [61]. n-Propanol (purity ≥ 99.5%, VWR, Canada) was used meter was < 0.05 of the reported value.
to determine the hydrophobicity of the NPs. The displacement tests
were performed with Flopaam™ 3630 s (HPAM, hydrolysis degree 5.2. Cryo-SEM images
25–35%, MW ≈ 20 × 106 Dalton) obtained from SNF Floerger (USA).
All chemicals were used as received. An FEI Quanta FEG 250 SEM equipped with a Gatan Alto 2500 cryo-
stage and xTm version 4.1.12.2162 software was used for the cryo-SEM
3. Nanoparticle characterization analysis. A small amount of the solution was loaded and frozen in a
nitrogen slush for sample preparation. Then, the surface of the sample
A Fourier transform infrared spectrometer, FTIR (model IRaffinity- was coated with a layer of gold (around 10 nm). The images were ob-
1s, Shimadzu, Japan) was used to analyze the SiO2 NPs. The back- tained using either of secondary electron or backscattered electron
ground spectrum was obtained with dry KBr, and each spectrum was detectors at an accelerating voltage of 5 to 10 kV. For cryo-SEM, 3–5
recorded over a range of wavelength of 4000–400 cm−1. different locations were captured for each sample.
Thermogravimetric analysis/differential scanning calorimetry (TGA/
DSC) measurements were performed by heating 6 mg sample of mod- 5.3. Rheological characterization
ified/unmodified NPs from 20 to 800 °C at a heating rate of 10 °C/min
under air atmosphere (SDT Q600, TA Instruments, Inc., New Castle, An Anton Paar MCR302 rheometer with a coaxial cylinder geometry
DE). For FTIR and TGA/DSC, every sample was tested three times for (measuring bob and measuring cup had radii of 13.329 mm OD and
reproducibility. The expected coupling mechanism between the silica 14.463 mm ID, respectively) was used to measure the rheological
surface and silanes is represented in Fig. 1. properties of the hybrid dispersions. Rheoplus/32 V3.62 software was
The hydrophobicity of the modified silica NPs was determined by connected to the rheometer to record the measurements. All the rheo-
using a dispersion stability test. For this procedure, 0.1 g of each NP logical measurements were done at a fixed temperature (T = 35 °C). To
type were dispersed in different mixtures of propanol/DI water (10 g) assure all systems were in the linear viscoelastic region for frequency
and ultrasonicated for 1 h. After that, the samples were left to settle for sweep measurements, a strain amplitude equal to 1% was selected
10 min, and the amount of precipitate was determined for each pro- (based on strain sweep results). In the frequency sweep test, the angular
panol/DI water mixture. frequency varied between 0.1 and 20 rad/s. Strain amplitude sweep test
in which strain amplitude varied between 1% and 1000%, was designed
4. Hybrid dispersion preparation to evaluate the nonlinear viscoelasticity of the hybrid dispersions. The
angular frequency was set to 0.5 rad/s, to prevent any flow instability,
Four different concentrations of unmodified and modified SiO2 NPs torque overload and wall-slip effects.
(1.0, 2.0, 3.0 and 4.0 wt%) were used to prepare hybrid dispersions. At
first, the SiO2 NPs were dispersed in DI water and ultrasonicated for 1 h. 5.4. Flooding experiments and set up
Then, HPAM powder was added at 0.4 wt% to the aqueous dispersions
under gentle stirring for 48 h. Fig. 2 depicts a schematic of the flooding setup. It consisted of a
pump, tubing, two transfer vessels, pressure transducer, sand-pack
5. Hybrid dispersion characterization holder, graduated cylinders, and back pressure regulator. The sand-
pack characteristics and the injected fluid properties are listed in
5.1. ζ-potential measurements Table 1. For the flooding experiments, each sand-pack was fully satu-
rated with water by imbibition method. The absolute permeability was
ζ-potential values were measured at 20 °C with a Zetasizer Nano ZS calculated using Darcy’s law. For this procedure, water was injected at
unit (Malvern Instruments Ltd, UK) with uncertainty of ± 1 to 6% of different flow rates and the corresponding pressure drop was recorded.
the reported value. 35 specimens were tested in the ζ-potential mea- The drainage process was carried out by injecting 2 PV of oil. After that,
surements (20 samples + 15 reproducibility tests). Before every 0.5 PV of water were injected followed by 0.5 PV of HPAM/SiO2 NPs

Fig. 1. Reaction route of surface modification on silica particles with (A) SA and OA, and (B) OTES [62].

624
L.M. Corredor, et al. Fuel 252 (2019) 622–634

Fig. 2. Sand-pack flooding set up.

Table 1 3400 cm−1 which corresponded to free silanol groups on the silica
The sand-pack specifications and properties of the injected fluid. surface [64].
Properties Values The bonding of the OA and SA molecules to SiO2 NPs was confirmed
by the asymmetric and symmetric (C–H) stretching vibrations of –CH3
Core holder length 30.4 cm or –CH2 groups observed at ∼2851 cm−1 and the asymmetric and
Core holder inner diameter 2.54 cm
symmetric stretching vibrations of COO– situated at ∼1536 cm−1 and
Particle size 0.0105–0.0149 cm
Sand-pack porosity 35–40%
∼1446 cm−1, respectively [65] (Fig. 4a and b).
Sand-pack permeability 5–6.5 Darcy
Connate water saturation 3.0 ± 0.2%
Pore volume 50 ± 1.5 cm3
6.1.2. TGA measurements
Oil viscosity (T = 25 °C) 940 ± 5 cP The thermogravimetric analysis (TGA) and differential scanning
Temperature T = 23–25 °C calorimetry (DSC) curves of unmodified and modified silica NPs are
shown in Fig. 5. The TGA profile of unmodified SiO2 NPs can be divided

dispersion. Finally, 2 PV of water were injected as a post-flush until the


oil cut at the production outlet was less than 1%. All the fluids were
injected at 0.1 ml/min.
Transmittance, %

6. Results and discussion

6.1. Nanoparticle characterization

6.1.1. FTIR measurements


The FTIR spectra of unmodified and modified silica NPs after nor-
malization of the peak area are shown in Figs. 3 and 4. The normal-
ization for the FTIR spectra was performed by using peaks derived from
bonds which are not involved in the modification reactions; namely
peaks between 900 and 1300 cm−1. The bonding of the OTES molecules 4000 3200 2800 2200 1600 1000 400
to SiO2 NPs was confirmed by the peaks at 2926 cm−1 and 2858 cm−1 -1
Wave number, cm
which were ascribed to very intense asymmetric and symmetric
stretching vibrations of C–H bonds of the CH2 groups in the octyl SiO2 SiO2-OTES OTES
substituent [63]. Additionally, the peak at 960 cm−1 which was as-
cribed to Si–O–Si bond, and the disappearance of the peak at Fig. 3. IR spectrum of SiO2, SiO2-OTES, and OTES.

625
L.M. Corredor, et al. Fuel 252 (2019) 622–634

Transmittance, %

Transmittance, %
(a) (b)

4000 3200 2800 2200 1600 1000 400 4000 3200 2800 2200 1600 1000 400
-1 -1
Wave number, cm Wave number, cm

SiO2 SiO2-OAA OA SiO2 SiO2-SAA SA

Fig. 4. IR spectrum of (a) SiO2-OAA, and (b) SiO2-SAA.

100 120 7. Hybrid dispersion characterization


90
80 7.1. ζ-potential measurements
80
Heat flow, mW
Weight, %

40 The ζ-potential values of the different NP dispersions are provided


70
60
in Table 2. Typically, NP dispersions with ζ-potential values lower than
0 −30 mV or larger than +30 mV are stable [69]. SiO2-OAA (−32 mV)
50 and SiO2-OTES (−29.6 mV) NPs display better stability than SiO2-SAA
-40 (−26.1 mV) NPs in water, despite the borderline zeta potentials, due to
40
Exo up the formation of a larger hydration layer. The hydration layer results
30 -80 from the hydrogen bonding between the silanol groups on the NPs with
0 200 400 600 800
o
the water molecules. This conclusion is supported by the TGA results
Temperature, C which showed that the weight loss associated to water for SiO2-SAA was
only 6.34%, whereas for SiO2-OAA and SiO2-OTES, it was 10.99% and
SiO2 TGA SiO2-OAA TGA SiO2-SAA TGA SiO2-OTES TGA 31.83%, respectively. Examining Table 2, it is concluded that all the
SiO2 DSC SiO2-OAA DSC SiO2-SAA DSC SiO2-OTES DSC hybrid dispersions prepared in this study are stabilized chiefly by steric
repulsion of the adsorbed polymeric chains. The adsorption of HPAM
Fig. 5. TG/DSC profiles of SiO2, SiO2-OTES, SiO2-SAA, and SiO2-OAA. on the surface occurs via hydrophobic interactions between the
polymer backbone (–CH2–CH2–) and the R-CH3 chain of the modifiers
into two regions, desorption of physisorbed water completed up to or by hydrogen bonding between the oxygen or nitrogen from HPAM
400 °C and dehydroxylation process of adjacent –OH groups between and the hydrogen from the SiO2 surface (SiO–H⋯N–H or SiO–-
400 °C and 800 °C [66]. The percentage of weight loss related to each H⋯O–CNH2) or between the hydrogen from the HPAM and the oxygen
region was 25.7% and 6.0%, respectively. of the SiO2 surface (SiO⋯HNH–CO–C) [70]. The effect of polymer ad-
A three-step weight loss was identified in the TGA profiles of all sorption onto dispersed particles in polymer solutions have been de-
modified NPs. The first step occurred below 200 °C which was attrib- scribed in detail by Kawaguchi [71]. In this study, the effect of pH on
uted to the evaporation of physically adsorbed water and volatile sol- the configuration of the adsorbed polymer molecules on the NP surface
vents. The weight loss in this region was 6.34%, 10.99% and 31.83% was not considered because the pH values of the hybrids did not change
for SiO2-SAA, SiO2-OAA and SiO2-OTES NPs, respectively. The second significantly (between 6.9 and 8.18).
step was attributed to the thermal decomposition of each modifier
grafted onto the surface of the NPs. For SiO2-OAA, the second step was 7.2. Cryo-SEM results
observed from 220 °C to 330 °C (Fig. 5, purple line) [67] whereas for
SiO2-SAA the second step was observed from 210 °C to 330 °C (Fig. 5, Fig. 7 depicts cryo-SEM images of aqueous unmodified and modified
blue line) [68] and for SiO2-OTES from 200 °C to 370 °C [23] (Fig. 5, SiO2 dispersions before and after adding HPAM. The size of the un-
green line). The weight loss in this region was 11.7%, 13% and 15.5% modified SiO2 NP aggregates in DI water is larger than those of the
for SiO2-OAA, SiO2-SAA and SiO2-OTES NPs, respectively. In the final modified SiO2 NPs (see Fig. 7A-D). This can be attributed to the hy-
step, the weight loss was attributed to the combustion of the degraded drogen bonding between the –OH groups on the surface of the un-
product. modified SiO2 NPs. The size of the aggregates of the unmodified NPs
Fig. 6 shows the NPs dispersed in different mixtures of DI water/ decreased after the surface was modified with OTES because the re-
propanol (W/A). It was observed that the precipitate of SiO2-SAA NPs pulsive forces between OTES molecules prevent the aggregation of the
and SiO2-OAA NPs increased as the concentration of DI water in the NPs (see Fig. 7A (1) and B (1)). Smaller aggregates are also observed for
mixture increased, while the precipitate of SiO2-OTES NPs reduced. SiO2-SAA and SiO2-OAA NPs. Their high hydrophobicity leads to the
This test and the water weight losses in the TGA analysis suggests that formation of larger aggregates which precipitated out, despite the ul-
the hydrophobicity of modified SiO2 can be ranked as follows: SiO2- trasonication, and, hence, could not be detected by the cryo-SEM
SAA > SiO2-OAA > SiO2-OTES. images (see Fig. 7C (1-2) and D (1-2)). When HPAM was added to the
aqueous dispersions, some polymer chains were adsorbed onto the SiO2
surface through hydrogen bonding or hydrophobic interaction. For the
SiO2-SAA, SiO2-OAA and SiO2-OTES hybrid dispersions, the repulsive
electrostatic forces between coated NP-coated NP and coated NP-

626
L.M. Corredor, et al. Fuel 252 (2019) 622–634

Fig. 6. Dispersion stability test of 1.0 wt% (A) SiO2-SAA, (B) SiO2-OAA, and (C) SiO2-OTES in water/propanol (W/A) mixtures.

Table 2 [72]. We believe the role of the depletion attraction forces is minimal
ζ-potential and pH values of aqueous and hybrid dispersions of the modified for the modified NPs, especially since for the modified NP hybrid dis-
and unmodified SiO2 NPs at 20 °C. persions the interaction with the HPAM chains prevented the floccu-
NP type HPAM NP concentration, ζ-Potential @ pH @ lation of the coated NPs, as evident from the different state of aggre-
concentration, ppm wt% 20 °C 20 °C gration of silica NP before and after adding HPAM molecules seen in
Fig. 7, hence minimizing the role of the depletion attractive forces. The
SiO2 4000 1.0 −85.4 7.48
adsorbed polymeric chains positively influenced the dispersivity of the
4000 2.0 −84.5 7.26
4000 3.0 −71.4 7.14 NPs, reducing the size of the aggregates by effective steric repulsion
4000 4.0 −66.3 6.9 (see Fig. 7A-D (3-4)). Some of the large aggregates formed by SiO2-SAA
0 2.0 −15.3 6 and SiO2-OAA NPs were captured in the cryo-SEM images after the
SiO2-OTES 4000 1.0 −82 8.1 addition of HPAM (see Fig. 7C (4) and D (4)).
4000 2.0 −76.7 7.86
4000 3.0 −73.3 7.73
4000 4.0 −63.2 7.69 8. Rheological characterization
0 2.0 −29.6 5.81

SiO2-OAA 4000 1.0 −75.3 8.18 8.1. Linear viscoelastic properties


4000 2.0 −70.8 8.11
4000 3.0 −67.3 8.10 Rheometry provides a distinctive, sophisticated tool to study the
4000 4.0 −64.8 8.0
0 2.0 −32 5.29
network structure of flocculated colloidal gels, yielding significant in-
formation on the nanofiller structural characteristics in a polymer so-
SiO2-SAA 4000 1.0 −68 7.82
lution [24,70]. The linear viscoelastic properties are affected by the
4000 2.0 −63 7.14
4000 3.0 −59.8 6.95 network structure of polymer chains in the polymer solution systems,
4000 4.0 −56.1 6.84 and this network structure can be captured in frequency sweep test
0 2.0 −26.1 4.99 [38,40,73]. The storage modulus (G') and the loss modulus (G'') are the
two main parameters that represent the linear viscoelastic properties. G'
provides information on the rigidity of the network structure, while G″
polymer (both negatively charged) contribute to disperse the NPs into quantifies how readily an applied stress to the structure is relaxed or
the polymer solution and to keep the polymer in an extended config- dissipated [74,75]. Fig. 8 shows G' and G″ of hybrid dispersions of
uration. However, the electrostatic repulsive forces do not play a major different unmodified and modified silica NPs and HPAM under small-
role in the adsorption of the polymer onto the coated NP surface. The amplitude oscillatory shear (strain amplitude γ0 = 0.1%) for an angular
van der Waals forces (hydrogen bonding) between NP-NP and NP- frequency range from 0.1 to 10 rad/s at 35 °C. It should be noted that in
polymer are reduced following modifying the NP surface with OAA, log-log scale, the errors are within the data point markers.
SAA and OTES because the number of –OH groups available on the NP As can be seen in Fig. 8b and c, increasing the content of SiO2-SAA
surface is reduced. In the meantime, the role of the –OH groups in and SiO2-OAA NPs damages the network structure of HPAM molecules
polymer adsorption is not significant. Otherwise, these forces are sig- and decreases the storage modulus of hybrid dispersions, particularly in
nificant in the absence of steric effects, i.e. for unmodified silica NPs the low-frequency region. In other words, SAA-based and OAA-based

627
L.M. Corredor, et al. Fuel 252 (2019) 622–634

Fig. 7. Cryo-SEM images of A) unmodified SiO2, B) SiO2-OTES, C) SiO2-OAA, and D) SiO2-SAA NPs dispersed in DI water or 4000 ppm of HPAM solution at low (1–3)
and high magnification (2–4).

hybrid dispersions have a weaker network structure compared to the hydrophobic interaction between the backbone of HPAM and the hy-
HPAM solution. Conversely, incorporating unmodified and SiO2-OTES drophobic functional group on SiO2-OTES NPs is the main driving
NPs improve the storage and loss moduli of HPAM solution, particularly mechanism to bring HPAM molecules to the vicinity of SiO2-OTES NPs.
at low-frequencies (see Fig. 8a and d). Due to large occupied space of silane on the surface SiO2-OTES NPs and
The reason for such differences in the network strength of hybrid the strength of hydrophobic affinity and electrostatic repulsion between
dispersions originates from the surface chemistry of unmodified and HPAM molecules and SiO2-OTES NPs, the contact of HPAM molecules
modified silica NPs. Such different behavior of unmodified and mod- and SiO2-OTES NPs occurs with a minimal number of segments at-
ified silica NPs can be related to their dispersion state in HPAM solution tached to the surface of SiO2-OTES NPs. This phenomenon increases the
and the configuration of adsorbed HPAM molecule on the surface of probability of creating configuration (A) in Fig. 9 for OTES-hybrid
silica NPs. Firstly, the dispersion of unmodified and modified silica NPs dispersions. In such configuration, polymer chains are moving and
with and without HPAM molecules was explained in the cryo-SEM protruding freely into the aqueous phase with high probability for
image discussion. Secondly, different configurations of adsorbed bridging between SiO2-OTES NPs and creating a 3D network formation.
polymer molecules onto the surface of carboxylic acid modified surfaces Frequency sweep results also confirm that by incorporating OTES NPs,
(SAA and OAA) and silane modified surface (OTES) are expected to the G' of hybrid dispersion increases and a stronger network is formed.
have a significant effect in addition to the dispersion state [76]. Possible Similarly, in SAA-based and OAA-based hybrid dispersions, hydro-
configurations of adsorbed HPAM molecule on the surface of modified phobic interaction between the backbone of HPAM and the hydro-
NPs are depicted in Fig. 9. phobic functional group of modified silica NPs control the adsorption of
We showed that the modified silica NPs can be ranked based on HPAM molecules. However, more hydrophobic affinity of SiO2-SAA and
their hydrophobicity as follows: SiO2-SAA > SiO2-OAA > SiO2-OTES. SiO2-OAA NPs compared to SiO2-OTES NPs may increase the number of
For SiO2-OTES NPs, due to the hydrophobicity of the modified chain, HPAM segments attached to the SAA and OAA silica NPs. As a result,

628
L.M. Corredor, et al. Fuel 252 (2019) 622–634

Fig. 8. Storage (G' , closed symbols) and loss (G'' , open symbols) moduli of the hybrid dispersions at 0.0, 1.0, 2.0, and 4.0 wt% of (a) unmodified and (b) SiO2-SAA
NPs, (c) SiO2-OAA NPs, and (d) SiO2-OTES modified silica NPs. HPAM = 4000 ppm.

configuration (B) in Fig. 9 is predicted to be more likely in SAA-based


and OAA-based hybrid dispersions. Any standing configuration for SAA-
based and OAA-based hybrid dispersions is not consistent with the
frequency sweep results. If the HPAM molecule can stand on the surface
of SiO2-SAA or SiO2-OAA NPs, by increasing the content of SiO2-SAA or
SiO2-OAA NPs, the probability of polymer bridging increases, and
larger storage modulus should be expected, which is not the case for
SAA-based and OAA-based hybrid dispersions (see storage modulus in
Fig. 9. Possible configuration of HPAM molecules adsorbed at NP/water in- Fig. 8b and 8c). Moreover, due to the unsaturated nature of the SiO2-
terface (A) random coil (high molecular weight polymers), (B) flat multiple site
OAA NPs, the hydrophobic affinity of SiO2-OAA NPs is less than SiO2-
attachments (strong adsorption).
SAA NPs and as a result, the severity of disruption of HPAM network by
the addition of OAA modified silica NPs is less than the addition of

629
L.M. Corredor, et al. Fuel 252 (2019) 622–634

SiO2-SAA NPs. solution. In contrast, for OTES-based hybrid dispersions, by increasing


the concentration of SiO2-OTES NPs, the plateau storage modulus in-
creases fivefold to 40 Pa for HPAM/4.0 wt% of SiO2-OTES NP solution
8.2. Nonlinear viscoelastic properties
(see Fig. 10d). Overall, the strain sweep results confirm that SiO2-OTES
NPs improve the viscoelastic properties of HPAM solution significantly
Due to converging and diverging pores in porous media, hybrid
in the linear region.
dispersions may experience large deformation, which can be exten-
sional or shear. Thus, from a practical point of view, the material re-
sponse will likely fall into a nonlinear regime of deformation. In this 8.3. Intra-cycle shear-thickening behavior
work, large shear deformation is studied to characterize the flow be-
havior of the hybrid dispersions. Fig. 10 depicts the G' of various hybrid It is desirable to find a clear physical meaning for the measured
dispersions at a constant frequency (=0.5 rad/s) and various strain storage and loss moduli in the nonlinear region. In the linear viscoe-
amplitudes (1–1000%) (The loss modulus (G″) plot in the strain sweep lastic region, the output stress signals are simple sinusoidal functions.
test is included in the Supplemental information, Fig. S1). The errors in However, by increasing the strain amplitude and shifting into the
modulus for the strain sweep test are within the data point markers. The nonlinear region, shear stress cannot be described simply by sinusoidal
plateau storage modulus (modulus in linear viscoelastic (LVR) region) waves. The mathematical details of shear stress (τ) response in the
and the critical strain amplitude (the strain amplitude at which G' starts linear and the nonlinear regions and the stress decomposition method
decreasing and reaches 90% of its original value) are the two main can be found in our previous works [24,25,77].
parameters which can be extracted from strain sweep test. In LVR, the When we are in one single cycle of oscillations at a constant strain
applied strain amplitude is not large enough to disrupt the at-rest amplitude, elastic and viscous local responses of a material at minimum
structures (i.e. storage modulus is independent of applied strain am- and maximum instantaneous strains and strain-rates, which are denoted
'
plitude). However, by further increase in strain amplitude, the non-LVR byGM , GL' , ηM' , ηL' , respectively, can be very helpful to characterize the
is approached (storage modulus changes with applied strain amplitude) nonlinear viscoelastic behavior and link the network structure to the
(see Fig. 10). rheological data. It should be noted here that all local nonlinear
As expected from the frequency sweep test, by increasing the con- properties are compiled directly by the rheometer software (Anton-Paar
tent of SiO2-OAA NPs, the plateau G' did not change significantly Rheoplus/32 V3.62). To interpret local nonlinear properties quantita-
compared with HPAM solution (Fig. 10c). In addition, the critical strain tively, shear-thickening and strain-stiffening ratios were defined as
amplitude for the OAA-based hybrid dispersion did not shift sig- follows [78].
nificantly relative to HPAM solution. However, by increasing the con- GL' − GM
'
tent of SiO2-SAA NPs (see Fig. 10b), the plateau for G' decreases to S= ,
GL' (1)
0.8 Pa for HPAM/4.0 wt% of SiO2-SAA NP solution from the plateau G'
for HPAM solution, which is 8 Pa. Moreover, the critical strain ampli- η' L − η' M
tude reaches approximately 8% for HPAM/4.0 wt% of SiO2-SAA NP T= .
η' L (2)
solution (the critical strain amplitude for HPAM solution is about 60%).
SiO2-SAA NPs aggregates can adsorb HPAM chains and extract them Positive S and T represent intra-cycle strain-stiffening and intra-
from the aqueous media and thus prevent them from creating a rigid cycle shear-thickening behaviors, respectively, and negative S and T
network, resulting in lower plateau modulus compared with HPAM indicate intra-cycle strain-softening and intra-cycle shear-thinning

10 2 10 2
(a) (b)

10 1 10 1
G'>Pa]

G'>Pa]

10 0 10 0

G'
HPAM 3630 G'
10 -1 HPAM 3630/1.0 wt%-Unmod-SiO2 10 -1 HPAM 3630
HPAM 3630/1.0 wt%-SiO2-SAA
HPAM 3630/2.0 wt%-Unmod-SiO2
HPAM 3630/2.0 wt%-SiO2-SAA
HPAM 3630/4.0 wt%-Unmod-SiO2
HPAM 3630/4.0 wt%-SiO2-SAA
10 -2 0 1 2 3 10 -2 0 1 2 3
10 10 10 10 10 10 10 10
Ȗ[%] Ȗ[%]
10 2 10 2
(c) (d)

10 1 10 1
G'>Pa]

G'>Pa]

10 0 10 0

G' G'
HPAM 3630 HPAM 3630
10 -1 10 -1 HPAM 3630/1.0 wt%-SiO2-OTES
HPAM 3630/1.0 wt%-SiO2-OAA
HPAM 3630/2.0 wt%-SiO2-OAA HPAM 3630/2.0 wt%-SiO2-OTES
HPAM 3630/4.0 wt%-SiO2-OAA HPAM 3630/4.0 wt%-SiO2-OTES
10 -2 0 1 2 3 10 -2 0 1 2 3
10 10 10 10 10 10 10 10
Ȗ[%] Ȗ[%]

Fig. 10. Oscillatory strain amplitude sweep response of the hybrid dispersions of HPAM and the different silica NPs at a fixed angular frequency (0.5 rad/s).
4000 ppm of HPAM solution with (a) unmodified silica NPs, (b) SiO2-SAA NPs, (c) SiO2-OAA NPs, and (d) SiO2-OTES NPs.

630
L.M. Corredor, et al. Fuel 252 (2019) 622–634

responses, respectively. In this study, only intra-cycle shear-thickening unmodified silica NPs leads to a similar effect on the behavior of the
ratio of the hybrid dispersions are presented in Fig. 11 (the strain-stif- hybrid dispersions at higher strain amplitudes, well above the yield
fening ratios are included in the Supplemental information, Fig. S2). As point of all systems (where the strain amplitude is much larger than the
can be seen in Fig. 11, for HPAM solution without silica NPs, T value is critical strain amplitude). As an example, for HPAM 3630/unmodified
approximately zero up to γ ≈ 60%, then T reaches a positive maximum SiO2, the shear-thinning (T < 0) behavior starts at γ0 > 200% (see
value at γ ≈ 250% followed by a decreasing trend back to approxi- Fig. 11a, blue line). By increasing the weight fraction of the different
mately zero (see black line in Fig. 11a). The positive T reveals the oc- types of modified and unmodified silica NPs, the value of the strain
currence of intra-cycle shear-thickening response. Generally speaking, amplitude at which T changes sign to negative values, decreases. All
in the low-frequency region, the nonlinear viscoelastic behavior is hybrid dispersions exhibit an intra-cycle shear-thinning viscoelastic
mainly described using a mechanism controlled by strain-rate-induced behavior (T < 0) at higher strain amplitudes. Such behavior could be
effects [79]. T > 0 could be related to the rate of flow-enhanced for- explained in conjunction with the wide-spread bond breakup in the
mation of network bonds and the rate of bond breakage for HPAM fumed silica NP flocculated structures, the orientation of the rigid
solution when deformation is applied in a single cycle. When the former polymer chains parallel to the flow direction and slippage at the NP-
is greater than the latter in one cycle, intra-cycle shear-thickening be- polymer interface [85–89]. Overall, introducing the different SiO2 NPs
havior occurs. Dupuis et al. [80] also studied the rheological properties decreases the shear-thickening behavior of the hybrid dispersions, de-
of solutions of high molecular weight partially hydrolyzed poly- creases the critical strain amplitude, and enhances the shear-thinning
acrylamide. They suggested that the shear-thickening behavior of such behavior above the yield point of all systems.
polymers originates from the formation and destruction of aggregates,
which are related to intra- and inter-molecular interactions. More de-
tails on the shear-thickening behavior can also be found in Hu and 8.4. Core flooding results
Jaimeson [81], and Bharadwaj et al. [79].
However, by increasing the concentration of SiO2-SAA NPs, the In the last part, the performance of all hybrid dispersions of HPAM/
maximum positive value decreases and shear-thinning (T < 0) beha- NPs for displacement of heavy oil was evaluated. Figs. 12 and 13 show
vior arises (see Fig. 11b). For HPAM/4.0 wt% of SiO2-SAA NP, the the pressure difference of sand-pack and the cumulative oil recovery
transition to nonlinear region starts at lower strain amplitudes versus pore volume of the injected fluid at 25 °C, respectively. The
(γ = 10%), and the maximum of T disappears. The same intra-cycle pressure difference drops during secondary water flooding (see WF#2
shear thickening behavior is observed for the addition of other silica in Fig. 12) and remains low and constant around 2.5 psi. During the
NPs (see Fig. 11a, c, and d). The shear-thickening ratio (T ) decreased as injection of the HPAM solution and the HPAM/NP dispersions (see PF
unmodified and modified silica NP concentration increased. This phe- in Fig. 12), the pressure difference increased. The pressure differences
nomenon could be attributed to immobilization of polymer chains at observed for 0.5 PV of HPAM/1.0 wt% SiO2-SAA and HPAM/1.0 wt%
the interface which makes a portion of the chains unavailable for de- SiO2-OAA dispersions were lower than those of the HPAM solution (0.5
formation-induced structuring in the polymer phase [82–84]. Such PV) due to their lower viscoelastic properties and higher mobility ratio.
behavior is observed for the whole range of concentration of all hybrid The details of viscoelastic property reduction of SiO2-SAA and SiO2-
dispersions. OAA NPs was explained earlier. In contrast, the higher-pressure dif-
Surprisingly, introducing the different types of modified and ferences of the HPAM/1.0 wt% SiO2-OTES dispersions can be related to
higher viscoelastic properties, lower mobility ratio, and pore blockage

80
60 (a) (b)
40
20
T [%]

0
-20 HPAM 3630 HPAM 3630
-40 HPAM 3630/1.0 wt%-Unmod-SiO2 HPAM 3630/1.0 wt%-SiO2-SAA
HPAM 3630/2.0 wt%-Unmod-SiO2 HPAM 3630/2.0 wt%-SiO2-SAA
-60
HPAM 3630/4.0 wt%-Unmod-SiO2 HPAM 3630/4.0 wt%-SiO2-SAA
-80
0 1 2 3 0 1 2 3
10 10 10 10 10 10 10 10
Ȗ[%] Ȗ[%]
80
60 (c) (d)
40
20
T [%]

0
-20 HPAM 3630 HPAM 3630
-40 HPAM 3630/1.0 wt%-SiO2-OAA HPAM 3630/1.0 wt%-SiO2-OTES
HPAM 3630/2.0 wt%-SiO2-OAA HPAM 3630/2.0 wt%-SiO2-OTES
-60
HPAM 3630/4.0 wt%-SiO2-OAA HPAM 3630/4.0 wt%-SiO2-OTES
-80 0 1 2 3 0 1 2 3
10 10 10 10 10 10 10 10
Ȗ[%] Ȗ[%]
Fig. 11. Intra-cycle shear-thickening ratio of the hybrid dispersion of HPAM and all silica NPs. 4000 ppm of HPAM solution with (a) unmodified silica NPs, (b) SAA
NPs, (c) OAA NPs, and (d) OTES NPs.

631
L.M. Corredor, et al. Fuel 252 (2019) 622–634

40 30 80
WF #2 PF WF #3
WF #2 PF 70
25
Pressure (psi)

30

Oil Recovery (%)


60

Pressure (psi)
20 20
50

15 40
10
30
0 10
0 0.5 1 1.5 2 2.5 3 20
Pore Volume 5
WF #3 10

Fig. 12. Pressure drop as a function of pore volume of the injected fluids at 0 0
25 °C (black line – HPAM, blue line – HPAM/ 1.0 wt% SiO2-SAA, purple line – 0 0.5 1 1.5 2 2.5 3
HPAM/ 1.0 wt% SiO2-OAA, red line – HPAM/ 1.0 wt% SiO2, green line – Pore Volume
HPAM/ 1.0 wt% SiO2-OTES). (For interpretation of the references to colour in
this figure legend, the reader is referred to the web version of this article.) Fig. 14. Cumulative oil recovery and pressure drop as a function of pore vo-
lume of the injected fluids at 25 °C (green line – HPAM/1.0 wt% SiO2-OTES,
light purple line – HPAM/0.5 wt% SiO2-OTES, grey line – HPAM/0.2 wt% SiO2-
80 71.6%
WF #2 PF WF #3 OTES). (For interpretation of the references to colour in this figure legend, the
71.4%
70.9% reader is referred to the web version of this article.)
62.6%
60
59.3%
Oil Recovery (%)

pores. By adding NPs, if they can enter the smaller pores, they can block
these pores and force water to enter the larger pores and displace the
43.1% remaining oil and increase the ultimate oil recovery [91]. Regarding the
40
pressure difference and the cumulative oil recovery of HPAM/1.0 wt%
SiO2-OTES, it can be concluded that the aggregate size of 1.0 wt% SiO2-
20 OTES NPs is not sufficiently small to go to the smaller water-wet pores,
block them, and improve oil recovery compared to HPAM solution. To
test this hypothesis, two lower concentrations of OTES NPs (0.2 and
0 0.5 wt%), which should have smaller aggregates, were injected into the
0 0.5 1 1.5 2 2.5 3
sand-pack (Fig. 14). The cumulative oil recovery of the HPAM solution
Pore Volume
increased to 75.7% and 73.8% of OOIP upon adding 0.2 and 0.5 wt% of
Fig. 13. Cumulative oil recovery as a function of pore volume of the injected SiO2-OTES NPs, respectively. Higher recoveries at lower OTES con-
fluids at 25 °C (light blue line – water, black line – HPAM, blue line – HPAM/ centrations suggest better propagation of the NPs inside the sand-pack
1.0 wt% SiO2-SAA, purple line – HPAM/ 1.0 wt% SiO2-OAA, red line – HPAM/ and blockage of water-wet pores.
1.0 wt% SiO2, green line – HPAM/ 1.0 wt% SiO2-OTES). (For interpretation of In general, it can be concluded that the performance of the HPAM/
the references to colour in this figure legend, the reader is referred to the web NPs for EOR processes is mainly related to the size of the NP aggregates
version of this article.)
and the interaction of the polymer and the NP in the aqueous solution,
i.e. the network structure of the HPAM/NP dispersion. The NPs (SiO2-
of the SiO2-OTES NPs in the sand-pack (to have a better understanding OAA and SiO2-SAA NPs) that negatively affected the viscoelastic
of viscosity change at different shear rates, the flow curves of all hybrid properties and the network structure of the polymer solution yielded
dispersions are included in the Supplemental information, Fig. S3). The lower oil recovery than that of the HPAM solution and the NPs with
pore blockage mechanism of NPs in porous media and its effect on ul- positive effect (SiO2-OTES NPs) on the network structure of HPAM
timate oil recovery will be discussed later. solution. The SiO2-OTES NPs system exhibited better dispersion, and
At the end of the extended water flooding (see WF #3 in Fig. 12), was stable in HPAM solution, thus yielding higher oil recovery. It
the pressure differences reached their initial value of ∼2.5 psi for the should be mentioned here that the effects of salinity and temperature
HPAM solution and the HPAM/1.0 wt% unmodified SiO2, HPAM/1.0 wt on the viscoelastic properties and heavy oil recovery of the hybrid
% SiO2-OAA, and HPAM/1.0 wt% SiO2-OAA dispersions. The higher dispersions were not assessed in this study.
pressure drop observed for HPAM/1.0 wt% SiO2-OTES dispersions, at
the end of the extended water flooding, suggest a reduction in the 9. Conclusions
permeability of the porous medium by virtue of the higher adsorption/
retention of the NPs in the sand-pack [90]. By altering the silanols of the surface of SiO2 NPs, three differently
The cumulative oil recovery after water flooding and (0.5 PV) modified SiO2 NPs were synthesized and characterized using TGA, DSC,
HPAM flooding was 43.1% and 71.4% OOIP, respectively (Fig. 13). It is and FTIR. The colloidal stability, dispersion state of the NPs, and the
observed that the addition of 1.0 wt% of unmodified SiO2 and SiO2- flow behavior of the aqueous solution of HPAM/different SiO2 NPs
OTES NPs to the HPAM solution did not change the performance of the were investigated using ζ-potential, cryo-SEM, and linear and nonlinear
polymer flooding, while the addition of 1.0 wt% of SiO2-OAA and SiO2- rheology, respectively. Finally, the performance of HPAM/all SiO2 NPs
SAA NPs reduced the cumulative oil recovery from 71.4% OOIP (for 0.5 while displacing heavy oil were evaluated in a linear sand-pack set up.
PV HPAM solution) to 59.3% and 62.6% OOIP, respectively. In the linear viscoelastic region, it was shown that bare SiO2 NPs che-
The main reason for lower cumulative oil recovery for water mically modified with silane (SiO2-OTES NPs) improved the viscoelastic
flooding is viscous fingering. Adding polymer molecules increases the properties of HPAM solution compared to carboxylic acids modified
viscosity of the injected fluid and decreases the mobility ratio and NPs (SiO2-SAA and SiO2-OAA). Such improvement originated from
prevents viscous fingering to a certain extent. It is well documented in better dispersion state of SiO2-OTES NPs and standing configuration of
the literature that in water-wet systems, water (wetting phase) occupies HPAM on the surface of SiO2-OTES NPs, which increased the prob-
the smaller pores and oil (non-wetting phase) mainly resides in the ability of polymer bridging. However, in the nonlinear region, all silica
larger pores [3]. As a result, during water flooding, water tends to enter NPs decreased the intra-cycle shear-thickening behavior and led to se-
smaller pores and bypasses large amount of oil residing in the larger vere intra-cycle shear-thinning after yielding of HPAM/ SiO2 NPs

632
L.M. Corredor, et al. Fuel 252 (2019) 622–634

network. The displacement tests showed that adding either SiO2-SAA [22] Ponnapati R, Karazincir O, Dao E, Ng R, Mohanty KK, Krishnamoorti R. Polymer-
and SiO2-OAA NPs to HPAM solution, both of which had a negative functionalized nanoparticles for improving waterflood sweep efficiency: char-
acterization and transport properties. Ind Eng Chem Res 2011;50(23):13030–6.
effect on the HPAM network structure, reduced the oil recovery of the [23] Choi SK, Son HA, Kim HT, Kim JW. Nanofluid enhanced oil recovery using hy-
HPAM solution from 71.4% OOIP to 59.3% and 62.6% OOIP, respec- drophobically associative zwitterionic polymer-coated silica nanoparticles. Energy
tively. In contrast, incorporation of 0.2 wt% SiO2-OTES NPs to HPAM Fuels 2017;31(8):7777–82.
[24] Aliabadian E, Sadeghi S, Kamkar M, Chen Z, Sundararaj U. Rheology of fumed silica
solution increased the oil recovery to 75.7% OOIP. nanoparticles/partially hydrolyzed polyacrylamide aqueous solutions under small
and large amplitude oscillatory shear deformations. J Rheol 2018;62(5):1197–216.
Acknowledgement [25] Aliabadian E, Kamkar M, Chen Z, Sundararaj U. Prevention of network destruction
of partially hydrolyzed polyacrylamide (Hpam): effects of salt, temperature, and
fumed silica nanoparticles. Phys Fluids 2019;31(1):013104.
The authors would like to thank Amitabha Majumdar and Ola Jabar [26] Delshad M, Kim DH, Magbagbeola OA, Huh C, Pope GA, Tarahhom F. presented at
for their technical assistance. Laura M. Corredor gratefully acknowl- the SPE Symposium on Improved Oil Recovery, 2008 (unpublished).
[27] Clarke A, Howe AM, Mitchell J, Staniland J, Hawkes LA. How viscoelastic-polymer
edges Ecopetrol S.A for the scholarship to pursue her graduate studies.
flooding enhances displacement efficiency. SPE J 2016;21(03):675–87.
This research was partially funded by NSERC Discovery Grants [28] Taylor KC, Nasr-El-Din HA. Water-soluble hydrophobically associating polymers for
#05503-2015 (US) and 03769-2017 (ZC), NSERC/AIEES/Energi improved oil recovery: a literature review. J Petrol Sci Eng 1998;19(3):265–80.
Simulation IRC in Reservoir Simulation, AITF (iCore) Chair in Reservoir [29] Smets G, Hesbain A. Hydrolysis of polyacrylamide and acrylic acid-acrylamide
copolymers. J Polym Sci, Part A: Polym Chem 1959;40(136):217–26.
Modelling, and the Frank and Sarah Meyer Foundation CMG [30] Ryles R. Chemical stability limits of water-soluble polymers used in oil recovery
Collaboration Centre. processes. SPE Reservoir Eng 1988;3(01):23–34.
[31] Wever D, Picchioni F, Broekhuis A. Polymers for enhanced oil recovery: a paradigm
for structure–property relationship in aqueous solution. Prog. Polymer Sci.
Appendix A. Supplementary data 2011;36(11):1558–628.
[32] Muller G, Fenyo J, Selegny E. High molecular weight hydrolyzed polyacrylamides.
Supplementary data to this article can be found online at https:// Iii. Effect of temperature on chemical stability. J Appl Polym Sci
1980;25(4):627–33.
doi.org/10.1016/j.fuel.2019.04.145. [33] Levitt D, Pope GA. Presented at the SPE Symposium on Improved Oil Recovery,
2008 (unpublished).
References [34] Samanta A, Bera A, Ojha K, Mandal A. Effects of alkali, salts, and surfactant on
rheological behavior of partially hydrolyzed polyacrylamide solutions. J Chem Eng
Data 2010;55(10):4315–22.
[1] Sheng JJ, Leonhardt B, Azri N. Status of polymer-flooding technology, SPE-174541- [35] Zaitoun A, Potie B. Presented at the SPE Oilfield and Geothermal Chemistry
PA 54 (02), 116-126 (2015). Symposium, 1983 (unpublished).
[2] Wever DAZ, Picchioni F, Broekhuis AA. Polymers for enhanced oil recovery: a [36] Sharma T, Sangwai JS. Effects of electrolytes on the stability and dynamic rheo-
paradigm for structure-property relationship in aqueous solution. Prog Polym Sci logical properties of an oil-in-water pickering emulsion stabilized by a nanoparti-
2011;36(11):1558–628. cle–surfactant–polymer system. Ind Eng Chem Res 2015;54(21):5842–52.
[3] Sorbie KS. Polymer-Improved Oil Recovery. (Springer Science & Business. Media [37] Martin F. Mechanical degradation of polyacrylamide solutions in core plugs from
2013. several carbonate reservoirs. SPE Form Eval 1986;1(02):139–50.
[4] Abidin AZ, Puspasari T, Nugroho WA. Polymers for enhanced oil recovery tech- [38] Xiang-Guo L, Wen-Hua Y, He-Long S, Zhen-Huan G. An experimental study on shear
nology. Procedia Chem 2012;4:11–6. degradation of Hpam solutions flowing through porous pedia. Oilfield Chem
[5] Wei B. Flow characteristics of three enhanced oil recovery polymers in porous 1995;4.
media. J Appl Polym Sci 2015;132(10). [39] Colby RH. Structure and linear viscoelasticity of flexible polymer solutions: com-
[6] Hou CT, Barnabe N, Greaney K. Biodegradation of xanthan by salt-tolerant aerobic parison of polyelectrolyte and neutral polymer solutions. Rheol Acta
microorganisms. J Ind Microbiol Biotechnol 1986;1(1):31–7. 2010;49(5):425–42.
[7] Katzbauer B. Properties and applications of xanthan gum. Polym Degrad Stab [40] Ferry JD. Viscoelastic properties of polymers. John Wiley & Sons; 1980.
1998;59(1–3):81–4. [41] Raman Ujjwal R, Sharma T, Sangwai JS, Ojha U. Rheological investigation of a
[8] Maurya NK, Mandal A. Studies on behavior of suspension of silica nanoparticle in random copolymer of polyacrylamide and polyacryloyl hydrazide (Pam-Ran-Pah)
aqueous polyacrylamide solution for application in enhanced oil recovery. Pet Sci for oil recovery applications. J Appl Polym Sci 2017;134(13).
Technol 2016;34(5):429–36. [42] Sharma T, Kumar GS, Sangwai JS. Viscoelastic properties of oil-in-water (O/W)
[9] Maghzi A, Mohebbi A, Kharrat R, Ghazanfari M. An experimental investigation of pickering emulsion stabilized by surfactant-polymer and nanoparticle–-
silica nanoparticles effect on the rheological behavior of polyacrylamide solution to surfactant–polymer systems. Ind Eng Chem Res 2015;54(5):1576–84.
enhance heavy oil recovery. Pet Sci Technol 2013;31(5):500–8. [43] Ghannam MT, Esmail MN. Rheological properties of aqueous polyacrylamide so-
[10] Maghzi A, Kharrat R, Mohebbi A, Ghazanfari MH. The impact of silica nanoparticles lutions. J Appl Polym Sci 1998;69(8):1587–97.
on the performance of polymer solution in presence of salts in polymer flooding for [44] Sharma T, Kumar GS, Chon BH, Sangwai JS. Viscosity of the oil-in-water pickering
heavy oil recovery. Fuel 2014;123:123–32. emulsion stabilized by surfactant-polymer and nanoparticle-surfactant-polymer
[11] Sharma T, Iglauer S, Sangwai JS. Silica nanofluids in an oilfield polymer poly- system. Korea-Australia Rheol J 2014;26(4):377–87.
acrylamide: interfacial properties, wettability alteration, and applications for che- [45] Sharma T, Sangwai JS. Silica nanofluids in polyacrylamide with and without sur-
mical enhanced oil recovery. Ind Eng Chem Res 2016;55(48):12387–97. factant: viscosity, surface tension, and interfacial tension with liquid paraffin. J
[12] Sharma T, Kumar GS, Sangwai JS. Comparative effectiveness of production per- Petrol Sci Eng 2017;152:575–85.
formance of pickering emulsion stabilized by nanoparticle–surfactant–polymer over [46] Costanzo S, Huang Q, Ianniruberto G, Marrucci G, Hassager O, Vlassopoulos D.
surfactant-polymer (Sp) flooding for enhanced oil recovery for brownfield reservoir. Shear and extensional rheology of polystyrene melts and solutions with the same
J Petrol Sci Eng 2015;129:221–32. number of entanglements. Macromolecules 2016;49(10):3925–35.
[13] El-Hoshoudy A, Desouky S, Betiha M, Alsabagh A. Use of 1-vinyl imidazole based [47] Dinic J, Zhang Y, Jimenez LN, Sharma V. Extensional relaxation times of dilute,
surfmers for preparation of polyacrylamide–SiO2 nanocomposite through aza- aqueous polymer solutions. ACS Macro Lett 2015;4(7):804–8.
Michael addition copolymerization reaction for rock wettability alteration. Fuel [48] Fardin MA, Perge C, Casanellas L, Hollis T, Taberlet N, Ortín J, et al. Flow in-
2016;170:161–75. stabilities in large amplitude oscillatory shear: a cautionary tale. Rheol Acta
[14] Althues H, Henle J, Kaskel S. Functional inorganic nanofillers for transparent 2014;53(12):885–98.
polymers. Chem Soc Rev 2007;36(9):1454–65. [49] Tanner R. Some methods for estimating the normal stress functions in viscometric
[15] Ahn SH, Kim SH, Lee SG. Surface-modified silica nanoparticle-reinforced poly flows. Trans Soc Rheol 1970;14(4):483–507.
(ethylene 2, 6-naphthalate). J Appl Polym Sci 2004;94(2):812–8. [50] Giacomin A, Dealy J, editors. Rheological measurement. Springer; 1998. p. 327–56.
[16] Mahdavian AR, Ashjari M, Makoo AB. Preparation of poly (styrene–methyl me- [51] Giacomin AJ, Dealy JM, editors. Techniques in rheological measurement. Springer;
thacrylate)/SiO2 composite nanoparticles via emulsion polymerization. An in- 1993. p. 99–121.
vestigation into the compatiblization. Eur Polymer J 2007;43(2):336–44. [52] Hyun K, Wilhelm M, Klein CO, Cho KS, Nam JG, Ahn KH, et al. A review of non-
[17] Ding X, Zhao J, Liu Y, Zhang H, Wang Z. Silica nanoparticles encapsulated by linear oscillatory shear tests: analysis and application of large amplitude oscillatory
polystyrene via surface grafting and in situ emulsion polymerization. Mater Lett shear (laos). Prog Polym Sci 2011;36(12):1697–753.
2004;58(25):3126–30. [53] Szopinski D, Luinstra GA. Viscoelastic properties of aqueous guar gum derivative
[18] Wu TM, Chu MS. Preparation and characterization of thermoplastic vulcanizate/ solutions under large amplitude oscillatory shear (laos). Carbohydr Polym
silica nanocomposites. J Appl Polym Sci 2005;98(5):2058–63. 2016;153:312–9.
[19] Ogoshi T, Chujo Y. Synthesis of organic–inorganic polymer hybrids by means of [54] Giacomin AJ, Gilbert PH, Merger D, Wilhelm M. Large-amplitude oscillatory shear:
host–guest interaction utilizing cyclodextrin. Macromolecules 2003;36(3):654–60. comparing parallel-disk with cone-plate flow. Rheol Acta 2015;54(4):263–85.
[20] ShamsiJazeyi H, Miller CA, Wong MS, Tour JM, Verduzco R. Polymer-coated na- [55] Jeyaseelan RS, Giacomin AJ. Network theory for polymer solutions in large am-
noparticles for enhanced oil recovery. J Appl Polym Sci 2014;131(15). plitude oscillatory shear. J Nonnewton Fluid Mech 2008;148(1–3):24–32.
[21] Rozenberg B, Tenne R. Polymer-assisted fabrication of nanoparticles and nano- [56] Kamkar M, Aliabadian E, Shayesteh Zeraati A, Sundararaj U. Application of non-
composites. Prog Polym Sci 2008;33(1):40–112. linear rheology to assess the effect of secondary nanofiller on network structure of

633
L.M. Corredor, et al. Fuel 252 (2019) 622–634

hybrid polymer nanocomposites. Phys Fluids 2018;30(2):023102. 1988.


[57] Lim HT, Ahn KH, Hong JS, Hyun K. Nonlinear viscoelasticity of polymer nano- [74] Shah S, Chen Y-L, Schweizer K, Zukoski C. Viscoelasticity and rheology of depletion
composites under large amplitude oscillatory shear flow. J Rheol flocculated gels and fluids. J Chem Phys 2003;119(16):8747–60.
2013;57(3):767–89. [75] Xu L, Gong H, Dong M, Li Y. Rheological properties and thickening mechanism of
[58] Carmona J, Ramirez P, Calero N, Munoz J. Large amplitude oscillatory shear of aqueous diutan gum solution: effects of temperature and salts. Carbohydr Polym
xanthan gum solutions. Effect of sodium chloride (Nacl) concentration. J Food Eng 2015;132:620–9.
2014;126:165–72. [76] Alagha L, Wang S, Yan L, Xu Z, Masliyah J. Probing adsorption of polyacrylamide-
[59] Renou F, Stellbrink J, Petekidis G. Yielding processes in a colloidal glass of soft star- based polymers on anisotropic basal planes of kaolinite using quartz crystal mi-
like micelles under large amplitude oscillatory shear (laos). J Rheol crobalance. Langmuir 2013;29(12):3989–98.
2010;54(6):1219–42. [77] Ewoldt RH, Hosoi AE, McKinley GH. An ontology for large amplitude oscillatory
[60] Jokinen M, Györvary E, Rosenholm JB. Viscoelastic characterization of three dif- shear flow. AIP Conf Proc 2008;1027(1):1135–7.
ferent sol-gel derived silica gels. Colloids Surf, A 1998;141(2):205–16. [78] Ewoldt RH, Hosoi A, McKinley GH. New measures for characterizing nonlinear
[61] Corredor-Rojas LM, Hemmati-Sarapardeh A, Husein MM, Dong M, Maini BB. viscoelasticity in large amplitude oscillatory shear. J Rheol 2008;52(6):1427–58.
Rheological behavior of surface modified silica nanoparticles dispersed in partially [79] Bharadwaj NA, Schweizer KS, Ewoldt RH. A strain stiffening theory for transient
hydrolyzed polyacrylamide and xanthan gum solutions: experimental measure- polymer networks under asymptotically nonlinear oscillatory shear. J Rheol
ments, mechanistic understanding, and model development. Energy & Fuels 2017;61(4):643–65.
2018;32(10):10628–38. [80] Dupuis D, Lewandowski F, Steiert P, Wolff C. Shear thickening and time-dependent
[62] Corredor LM, Husein MM, Maini BB. Effect of hydrophobic and hydrophilic metal phenomena: the case of polyacrylamide solutions. J Nonnewton Fluid Mech
oxide nanoparticles on the performance of xanthan gum solutions for heavy oil 1994;54:11–32.
recovery. Nanomaterials 2019;9(1):94. [81] Hu Y, Wang S, Jamieson A. Rheological and rheooptical studies of shear-thickening
[63] Parale V, Mahadik D, Mahadik S, Kavale M, Wagh P, Gupta SC, et al. OTES modified polyacrylamide solutions. Macromolecules 1995;28(6):1847–53.
transparent dip coated silica coatings. Ceram Int 2013;39(1):835–40. [82] Kabanemi KK, Hétu J-F. A reptation-based model to the dynamics and rheology of
[64] Al-Oweini R, El-Rassy H. Synthesis and characterization by FTIR spectroscopy of linear entangled polymers reinforced with nanoscale rigid particles. J Nonnewton
silica aerogels prepared using several Si(or)4 and R′′Si(or′)3 precursors. J Mol Fluid Mech 2010;165(15):866–78.
Struct 2009;919(1):140–5. [83] Sprakel J, Lindström SB, Kodger TE, Weitz DA. Stress enhancement in the delayed
[65] Oomens J, Steill JD. Free carboxylate stretching modes. J Phys Chem A yielding of colloidal gels. Phys Rev Lett 2011;106(24):248303.
2008;112(15):3281–3. [84] Song Y, Zheng Q. Concepts and conflicts in nanoparticles reinforcement to polymers
[66] Zhuravlev L. The surface chemistry of amorphous silica. Zhuravlev Model. Colloids beyond hydrodynamics. Prog Mater Sci 2016;84:1–58.
Surfaces A: Physicochem Eng Aspects 2000;173(1–3):1–38. [85] Chae DW, Kim BC. Effects of interface affinity on the rheological properties of zinc
[67] Mahdavi M, Ahmad MB, Haron MJ, Namvar F, Nadi B, Rahman MZA, et al. oxide nanoparticle-suspended polymer solutions. Macromol Res 2010;18(8):772–6.
Synthesis, surface modification and characterisation of biocompatible magnetic [86] Heinrich G, Klüppel M. Recent advances in the theory of filler networking in
iron oxide nanoparticles for biomedical applications. Molecules elastomers. Filled Elastomers Drug Delivery Systems 2002. p. 1–44.
2013;18(7):7533–48. [87] Cai L-H, Panyukov S, Rubinstein M. Hopping diffusion of nanoparticles in polymer
[68] Awais M, Jalil M, Zulfiqar U, Husain S. Presented at the IOP Conference Series: matrices. Macromolecules 2015;48(3):847–62.
Materials Science and Engineering, 2016 (unpublished). [88] Lin C-C, Parrish E, Composto RJ. Macromolecule and particle dynamics in confined
[69] Standard, A., D4187-82, 1985, Zeta potential of colloids in water and waste water. media. Macromolecules 2016;49(16):5755–72.
(ASTM Standard D4187-82). [89] Poling-Skutvik R, Krishnamoorti R, Conrad JC. Size-dependent dynamics of nano-
[70] Hu Z, Haruna M, Gao H, Nourafkan E, Wen D. Rheological properties of partially particles in unentangled polyelectrolyte solutions. ACS Macro Lett
hydrolyzed polyacrylamide seeded by nanoparticles. Ind Eng Chem Res 2015;4(10):1169–73.
2017;56(12):3456–63. [90] Ehtesabi H, Ahadian MM, Taghikhani V. Enhanced heavy oil recovery using TiO2
[71] Kawaguchi M. Effects of polymer adsorption and desorption on rheological prop- nanoparticles: investigation of deposition during transport in core plug. Energy
erties of silica suspensions. Colloids Surf, A 1994;86:283–9. Fuels 2014;29(1):1–8.
[72] Ottewill RH. Structure and dynamics of polymer and colloidal systems. Dordrecht: [91] Ponnapati R, Karazincir O, Dao E, Ng R, Mohanty K, Krishnamoorti R. Polymer-
Springer Netherlands; 2002. p. 1–20. functionalized nanoparticles for improving waterflood sweep efficiency: char-
[73] Doi M, Edwards SF. The theory of polymer dynamics. Oxford University Press; acterization and transport properties. Ind Eng Chem Res 2011;50(23):13030–6.

634

You might also like