Vous êtes sur la page 1sur 12

Journal of Petroleum Science and Engineering 172 (2019) 1057–1068

Contents lists available at ScienceDirect

Journal of Petroleum Science and Engineering


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

Appraising the impact of metal-oxide nanoparticles on rheological T


properties of HPAM in different electrolyte solutions for enhanced oil
recovery
Mohammed Bashir Abdullahia, Kourosh Rajaeia, Radzuan Junina,∗, Ali Esfandyari Bayatb
a
School of Chemical and Energy Engineering, Faculty of Engineering, Universiti Teknologi Malaysia, Skudai, Johor Bahru, Malaysia
b
Department of Reservoir Engineering, Faculty of Petroleum and Chemical Engineering, Science and Research Branch, Islamic Azad University, Tehran, Iran

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

Keywords: Polymer has been employed as a valuable chemical to increase mobility ratio and subsequently to improve sweep
Rheology efficiency of water flooding. However, its sensitivity to harsh reservoir conditions which are high formation
Hydrophilic nanoparticles brine salinity with high temperature has degraded its thickening characteristics and lowered its performance in
HPAM oil reservoirs. Recently, the application of nanoparticles (NPs) has shown a magnificent impact on chemical
Salt tolerance
performance and eventually boosted chemical flooding recovery under reservoir condition. It has been observed
Thermo-stability
EOR
that the role of NPs type as rheology control agents in modification of the polymer performance in reservoir
condition has not earned enough attention. The aim of this study is to develop NPs-HPAM hybrid dispersions by
employing aluminium oxide, silicon dioxide and titanium oxide NPs to partially hydrolyzed polyacrylamide
(HPAM) solutions. The rheological properties of NPs-HPAM hybrid dispersions were examined in the presence of
mono- and divalent ions at different shear rates and NPs concentrations. The results revealed that increasing the
salt concentration decreases viscosity of the polymer solutions. NPs-HPAM hybrid dispersions exhibit higher
viscosity than that of HPAM solutions at the same salinity which contributed to the NPs shield against cations
present in the salt. In addition, the Fourier transform infrared spectral data established the presence of hydrogen
bonds between NPs and carbonyl groups in HPAM solutions which confirmed improve performance of HPAM
solutions by NPs under saline condition. Moreover, NPs-HPAM hybrid dispersions showed higher viscosity than
polymer solution at low and medium shear rates, while the reversed trend was detected at higher shear rate due
to the adsorption interaction of polymer molecules on NPs surface. Furthermore, the efficiency of NPs-HPAM
hybrid dispersions as innovative tertiary recovery technique was put to the test through a water-wet sand pack
saturated with intermediate oil. The results indicated that alumina-, silica- and titanium-polymers dispersions
have ultimate oil recoveries of 15%, 10% and 6% higher than polymer flooding. Results of these researches
prove that application of NPs-HPAM hybrid dispersions is the innovative and advanced method since NPs im-
prove the HPAM solution thickening behaviour and rheological properties in the high salinity reservoir condi-
tions.

1. Introduction Ramazani et al., 2010; Giese and Powers, 2002). Polymer solutions
applications have shown some limitations and chemical degradations
Giving the current and dynamic growing demand of energy in the under high salinity and temperature in the hydrocarbon reservoirs
world, crude oil remains as one of the major primary energy source. The (Seright et al., 2010). HPAM is the most popular and widely used
implementations of enhanced oil recovery (EOR) schemes have become polymer in the oil fields for polymer flooding purposes. It has shown
global practice due to the difficulties in exploring and developing new high tolerance to mechanical forces during EOR operations, less cost
oil fields. EOR techniques have been implemented to improve both effective, and more resistance to bacterial degradation. However, for-
macroscopic sweep and microscopic displacement efficiency respec- mation brine salinity, hardness and high temperature of the oil re-
tively (Sheng, 2014). Polymer can modify mobility ratio by increasing servoirs have detrimental influences on HPAM solutions rheology
viscosity of water phase as well as its relative permeability (Ahmad (Olajire, 2014).


Corresponding author.
E-mail addresses: r-radzuan@utm.my, radzuan@petroleum.utm.my (R. Junin).

https://doi.org/10.1016/j.petrol.2018.09.013
Received 4 April 2018; Received in revised form 20 July 2018; Accepted 4 September 2018
Available online 05 September 2018
0920-4105/ © 2018 Published by Elsevier B.V.
M. Bashir Abdullahi et al. Journal of Petroleum Science and Engineering 172 (2019) 1057–1068

Nomenclature PAM Polyacrylamide


PV Pore volume
Al2O3 Aluminium oxide SB Synthetic brine
CaCl2 Calcium chloride SiO2 Silicon dioxide
DIW Deionised water Swc Connate water saturation
DLS Dynamic light scatting TEM Transmission electron microscopy
EOR Enhanced oil recovery TiO2 Titanium Oxide
FTIR Fourier transform infrared XRD X-Ray diffraction
HPAM Partially hydrolyzed polyacrylamide τo Yield Point Stress
IFT Interfacial tension τ Shear stress (Pa)
MgCl2 Magnesium chloride K Flow consistency index
NaCl Sodium chloride n Flow behaviour exponent
NPs Nanoparticles γ Shear rate (s−1)
NS Nanosilica

Recently, many research findings show that NPs are valuable and (PAM) solution can increase the pseudoplasticity behaviour of PAM
effective inorganic agents in oil recovery (Bayat et al., 2014; Nguyen solution. The ultimate oil recovery was boosted by 10% after im-
et al., 2014; ShamsiJazeyi et al., 2014; Zargartalebi et al., 2015; Zhang plementing nanodispersion flooding in comparison with PAM flooding
et al., 2014). The performance of aluminium and silicon oxides (Al2O3 attributable to interaction between NPs and polymer chain by im-
and SiO2) NPs were studied and tested as a tertiary recovery method. proving mobility ratio at low and medium shear rates (Maghzi et al.,
The total oil recovery for NPs dispersions were found to be in the range 2013). The effect of TiO2 concentration in HPAM solution from 1.9, 2.1,
of 62–81%. The main factor contributed for high recovery efficiency is 2.3, and 2.5 wt% on heavy oil recovery was investigated. The results
interfacial tension reduction and viscosity modification of the injected show increasing in the viscosity of polymer solution by increasing the
fluid (Ragab and Hannora, 2015). Maghzi et al. (2013) reported that NPs weight fraction as more NPs in nanopolymer dispersion have direct
introducing NPs can modify the viscosity of polymer solution during influence on fluid shear rate. It was also noticed that viscosity of na-
polymer flooding to improve sweep efficiency. Addition of SiO2 NPs nopolymer is greater in low shear rate than in greater shear rate. Dis-
concentration higher than 0.1 wt% can be studied to determine the placement tests results showed an increase of about 4% in oil recovery
possibility of increasing viscosity of nano dispersion. At low shear rate, for nanopolymer solution (Cheraghian, 2016).
0.1 wt% concentrations can increase polymer viscosity (Maghzi et al., The aim of this experimental research, therefore, was to solve
2013). SiO2 and Al2O3 mixture were put to the test at 0.05 wt% con- polymer chemical degradation challenges by investigating the influence
centration. The result revealed that the viscosity of crude oil emulsion of NPs types on HPAM solution viscosity under saline environment
was decreased by 25% by introducing SiO2 and Al2O3 NPs. Moreover, (30̗000 ppm) and at elevated temperature (up to 70 °C). Moreover, the
the incremental oil recovery was achieved by 9.6% after utilizing the applicability of NPs types on HPAM solution flooding at ambient tem-
mixture of SiO2 and Al2O3 NPs (Alomair et al., 2015). The performance perature was explored in order to justify the NPs influences on rheo-
of the nanosilica (NS) in the polymer flooding in the presence of salt on logical behaviour of HPAM solution under high salinity condition.
the heavy oil recovery was investigated. The oil recovery factor after
injecting 1 pore volume (PV) of polymer/NS dispersion was improved
by 10% which contributed to the viscosity enhancement of the polymer 2. Experiment
in the presence of SiO2 NPs (Yousefvand and Jafari, 2015). Ogolo et al.
(2012) reported that viscosity of crude oil after adding Al2O3 NPs dis- 2.1. Material
persion was reduced which is the main mechanism to increase oil re-
covery. Furthermore, they observed interfacial tension (IFT) and heavy Partially hydrolyzed polyacrylamide (HPAM, degree of hydrolysis
oil viscosity reduction by employing Al2O3 NPs. It was found that not 12 mol%, MW 13 × 10+6 Da, water-soluble with negative surface
only the presence of SiO2 NPs can change the rock wettability, but also charges along its chain) was used to prepare polymer solutions. Three
can reduce IFT in the presence of ethanol. Bhardwaj et al. (2008) in- available commercial NPs namely, aluminium oxide NP (Al2O3, with
dicated that there was strong interaction between silanol functional 99% purity, size of 20–30 nm and specific surface area of 80 m2/g),
group of SiO2 NPs and the polymer chain by forming hydrogen bonding silicon dioxide NP (SiO2, with 99.5% purity, size of 20 nm, specific
which can increase the thermal stability of the NPs-HPAM hybrid dis- surface area of 160 m2/g and non-porous) and titanium oxide NPs
persions. It was reported that addition of SiO2 NPs to polyacrylamide (TiO2, with 99% purity, size of 20–25 nm and specific surface area of
60 m2/g) were purchased from Sky Spring Nanomaterial, Inc., Houston,

Table 1
Properties of the materials used in the experiments.
Material Molecular weight (g/mol) Density (gr/cm3) Degree of hydrolysis Viscosity (cp) Specific surface area (m2/g) Particle size (nm)

HPAM 13 × 106 1.02 12% – – –


Synthetic Brine 18 1.04 – 2.5 – –
SiO2 60.08 – – – 160 20
TiO2 79.90 – – – 60 20–25
Al2O3 101.96 – – – 80 20–30
NaCl 58.44 2.16 – – – –
CaCl2 110.98 1.28 – – – –
MgCl2 95.211 2.32 – – – –
DIW 18.01 1.00 – – – –
Toluene 92.14 0.87 – – – –
Medium Oil – 0.89205 (27oAPI) – 217 – –

1058
M. Bashir Abdullahi et al. Journal of Petroleum Science and Engineering 172 (2019) 1057–1068

TX, USA. Sodium chloride (NaCl), calcium chloride (CaCl2) and mag- dispersions under reservoir salinity. To measure the spectra of HPAM
nesium chloride (MgCl2) were received from QRëC Chemical Co., Ltd. and NPs-HPAM hybrid dispersions, the Perkin-Elmer FTIR spectrometer
to make the electrolyte solutions. All the NPs and polymer were utilized 1710 with a resolution of 4 cm−1 and scanning range between 400 and
as received without further modification and purification. The details of 4000 cm−1 was used.
all material used in this study are shown in Table 1.
2.6. Viscosity measurement
2.2. Brine solution preparation
The rheology of polymer solution was analysed using the rheometer
In order to simulate reservoir salinity and study the effect of mono- and texture analyser (350 RST BROOKFIELD rheometer) equipped with
and divalent ions, two different synthetic brine with several con- temperature controller. The rheometer is equipped with a quick con-
centrations of 10̗000, 20̗000, 30̗000, 40̗000 and 60̗000 ppm were pre- nect coupling for easy spindle attachment, and a chamber is designed to
pared in deionised water (DIW). First synthetic brine (SB1) is composed accommodate the coaxial cylinder spindle and samples for rheological
of 100% NaCl, while the second synthetic brine (SB2) is made up of 95% evaluation. The chamber has roughened surface to prevent slippage
NaCl, 4% CaCl2 and 1% MgCl2 as presented in Table 2. The Synthetic effect. Viscosities of polymer solutions and NPs-HPAM hybrid disper-
Brines were used as dispersing fluids which were ultrasonicated for sions under different shear rates (1, 10, 100, 400, and 1̗000 s−1) were
15 min prior to each experiment to prevent aggregation of the particles. measured at 26 °C, 50 °C, and 70 °C. In order to calibrate the rheometer
setup, standard oil and pure water with known viscosities at different
2.3. Preparation of polymer solution temperatures were used.

HPAM was gradually and uniformly added to the upper shoulder of 2.7. Displacement test
vortex of the DIW to form a polymer solution with various concentra-
tions from 1̗000 to 5̗000 ppm. The polymer solutions were homo- The improvement of polymer flooding performance by applying
genized and stirred slowly for 48 h at 60 RPM of the spinning speed for three different NPs on oil recovery was studied by Sandpack flooding
constant vortex by magnetic stirrer. The polymer solutions were being experiments (Fig. 1). Glass bead with high purity and sand size in the
closed and sealed to avoid contact with oxygen and also chemical de- range of 50–120 μm was prepared in designed quantity to represent
gradation during preparation process. Solutions were later allowed to homogeneous porous medium. Four visible cylindrical pipes with a
stand overnight for complete hydration. Each sample was stirred for 3.4 cm ID and 44 cm length were used as a glass bead holder for dis-
20 min before any experiment. Moreover, to study the effect of salinity placement tests. Porosity and permeability of each sandpack were
on HPAM properties, 2̗000 ppm of HPAM was prepared in different measured before each displacement test and reported in Table 3. In
synthetic brine solutions from 10̗000 to 60̗000 ppm as explained ear- order to prepare sandpacks for displacement test the glass bead holders
lier. was evacuated to ensure there is no air was trapped in porous medium.
Subsequently, the holders were completely saturated with synthetic
formation water (30̗000 ppm) and crude oil was injected until connate
2.4. Preparation of NPs-HPAM hybrid dispersion
water saturation (Swc) reached. The intermediate crude oil (0.89205 gr/
cc and 217 ± 0.2 cp) at ambient temperature was utilized in this ex-
Different concentrations of SiO2 NPs (0.005, 0.075, 0.1, 0.125, 0.15,
perimental study. To evaluate the performance of waterflooding 4 PVs
0.2 and 0.25 wt%) were dispersed in DIW and ultrasonicated for 20 min
of the brine (30̗000 ppm) was injected with the 1 ml/min constant flow
via ultrasonic bath (Crest Ultrasonics water bath) to form stable dis-
rate to the sandpack holder. In addition, four different displacement
persions. HPAM powders were then dissolved at desired concentration
floodings were implemented into the porous medium in which 1.5 PVs
(2̗000 ppm) in SiO2 NPs dispersions in order to determine optimum
of the brine (30̗000 ppm) was followed by 2.5 PVs of HPAM solution,
concentration of SiO2 NPs. This optimum concentration of SiO2 NPs was
SiO2-HPAM, Al2O3-HPAM, and TiO2-HPAM dispersions until oil pro-
applied to other NPs' concentrations to study their influence on HPAM
duction was negligible in the effluent at ambient temperature. During
solution performance at same condition.
oil recovery flooding tests the volumes of produced oil were measured
as a function of time.
2.5. Nanoparticles characterization
3. Results and discussion
In order to determine the NPs compositions, sizes, and morpholo-
gies, two different tests were carried out. First, X-Ray Diffraction (XRD)
3.1. Nanoparticles characterization
analysis was conducted with a SIEMENS diffractometer D5000 with Cu
Kα radiation (λ = 0.15418 nm) at the 2θ angle scanning range of
The NPs sizes and morphologies were examined by XRD and TEM
10–65° to determine the sizes and compositions of the NPs by Scherrer's
analyses (Figs. 2 and 3). XRD results indicated Al2O3-NP has alpha (α)
formula (Equation (1)).
crystalline structure and pure composition. However, TiO2 and SiO2
0.9 λ NPs have shown semi-crystalline structures. The former NP has com-
dhkl =
β cos θβ (1) position of anatase mineral while the latter has quartz component.
From Equation (1) the size of Al2O3, TiO2 and SiO2 NPs were calculated
Whereas dhkl is the mean NP size (nm), β (radians) comprise the full
width at half maximum of the broadened diffraction line observed at Table 2
the 2θ angle range, λ is the wavelength of Cu Kα radiation Synthetic brine properties.
(λ = 0.15418 nm), and θβ is the Bragg angle of diffraction. Further-
Salt Percentage Purity, (%) Molecular weight, (g/mol) Brine type (ppm)
more, size and morphology of NPs were also determined by Transmis-
sion Electron Microscopy (TEM, model JEM-2100/, JEOL, SB1 SB2
Acc.200.00 kV) analysis.
Additionally, Fourier Transform Infrared (FTIR) test within the NaCl 99 58.44 30̗000 28̗500
CaCl2 78 110.99 0.0 1̗200
scanning range of 500 to 4000 cm−1 was conducted to determine the
MgCl2 95 95.21 0.0 300
functional groups and interaction between NPs and HPAM molecules in TDS 30̗000 30̗000
order to justify the viscosity improvement of NPs-HPAM hybrid

1059
M. Bashir Abdullahi et al. Journal of Petroleum Science and Engineering 172 (2019) 1057–1068

Fig. 1. Schematic of the displacement test apparatus.

and found to be 26, 11.8, and 16.1 nm, respectively. Furthermore, the 1981). It can be attributed to the impairment of the hydrophobic effect
NPs size distribution in DIW was shown in Fig. 3(d) using a Malvern due to the higher mobility of the HPAM chains and the losing inter-
Nanosizer based on the Dynamic Light Scatting (DLS) method. The molecular connections (Yahya et al., 1995). It was found that at tem-
average particles size of Al2O3, TiO2, and SiO2 NPs were observed as 22 perature above 60 °C, the viscosity of HPAM solution was drastically
(10–42), 28 (15–42) and 20 (10–48) nm, respectively. declined due to the hydrolysis reaction of the acrylamide groups of
HPAM (Seright et al., 2010). The hydrolysis reaction products are de-
3.2. HPAM concentration and viscosity behaviour graded in the presence of inorganic salts in SB2 (NaCl, CaCl2, and
MgCl2).
Fig. 4 shows the effect of HPAM concentrations on solution viscosity
at 26 °C under different shear rates. The rheogram and flow curves
exhibit pseudoplastic behaviour over the range of shear rate measured 3.4. The behavior of HPAM solution viscosity in electrolyte solutions
as the HPAM concentrations increases at 26 °C. Therefore, Herschel-
Bulckley model is suggested to best fit the relationship for flow curves The effect of mono- and divalent ions on viscosity of 2̗000 ppm of
(pseudoplastic with yield stress, Fig. 4). The presence of the yield stress HPAM solutions were investigated and presented in Fig. 6. It was de-
might result from the absence of salts in the HPAM solution composi- tected that HPAM solution viscosity was gradually declined with the
tion. The salt content negatively affects not only the viscosity of the increase in salt concentration in SB1 which composed of 100% mono-
HPAM fluids but also the yield stress (Tovar et al., 2014; Wilton, 2015). valent ions (Na+ and Cl−). On the other hand, the combination of
It is a model (τ=τo + Kγn), that represents the behaviour of non-New- mono- and divalent ions (95% Na+, 4% Ca2+ and 1% Mg2+) showed
tonian fluids which is appropriate for concentrated solutions in the detrimental influence on HPAM solution viscosity which leads to rapid
absence of salt (Lopes et al., 2014). It was also noticed that increasing decline on viscosity curve compare to SB1 solution as the concentration
HPAM concentration directly increases HPAM solution viscosity not of salt increases in the HPAM solution. The molecules of polymer are in
only at low shear rate, but also at higher shear rate due to the rise in the form of lumps and flocks under high salinity brine while in low salinity
intermolecular entanglement and higher resistant to flow. HPAM so- brine, molecules are long chain and straight connected skeleton. The
lution viscosity is a function of shear rate for pseudoplastic fluids (Lopes higher the concentrations of salt in polymer suspension results in the
et al., 2014; Silveira et al., 2016). Hence, the viscosities were gradually lower reaction between the polymer molecules and the salt ions in the
reduced with an increase of the shear rate, showing shear-thinning solution. There is a swift reaction between cations (Na+, Ca2+, Mg2+)
characteristics. The curling and uncoiling of polymer skeleton structure from brines and C]O from polymer molecular chain structure that
under high shear rate can contribute to the pseudoplastic behaviour in cause shielding of the mutual repulsions between carboxylic groups
the HPAM solution. along the backbone of HPAM molecule which would lead to a decrease
of the hydrodynamic volume, resulting in additional viscosity reduction
3.3. Thermal degradation of HPAM solution at different salt composition (Fig. 7). Polymer molecules under high salinity brines are like cotton
and its tendency to form bond with monovalent or divalent ions is very
Thermal degradation of 2̗000 ppm HPAM solution was analysed at low, which would also leads to reduction of the viscosity. It was dis-
the presence of SB1 and SB2 at 26 °C, 50 °C, and 70 °C (Fig. 5). It was covered that the change in viscosity of polymer solution which contains
detected that increasing temperature dramatically decreases the visc- divalent ions shows higher viscosity loss compare to monovalent ions
osity of HPAM solution at low, medium and high shear rates. At the due to the presence of calcium and magnesium ions. It can be resulted
same salinity and shear rate, the intermolecular force of interaction to the deformation of HPAM molecules to become spherical and also
between polymer molecules was reduced due to the increase in thermal non-ionic molecules which contribute to HPAM solution viscosity loss
motion of the molecules, thereby solution viscosity reduced (Muller, (Maghzi et al., 2014).

Table 3
Sandpack models properties.
Model no. L (cm) ID (cm) BV (cc) Sand size (μm) PV (cc) Porosity (%) Permeability (D) Initial oil saturation (%) Swc (%)

#1 44 3.4 399.28 50–120 131.7624 33 3.9 86.5 13.5


#2 44 3.4 399.28 50–120 135.7552 34 3.7 87 13
#3 44 3.4 399.28 50–120 131.7624 33 3.9 85.9 14.1
#4 44 3.4 399.28 50–120 139.748 35 3.8 86.3 13.7
#5 44 3.4 399.28 50–120 139.748 34 3.7 86.8 12.2

1060
M. Bashir Abdullahi et al. Journal of Petroleum Science and Engineering 172 (2019) 1057–1068

Fig. 2. X-ray diffraction (XRD) analysis of (a) Al2O3, (b) TiO2, and (c) SiO2 NPs.

3.5. Effect of the SiO2 NPs concentration on the viscosity of NPs-HPAM has higher viscosity than NPs-HPAM hybrid dispersions at higher shear
hybrid dispersions rate ranges (0.6–10 s−1). Zeyghami et al. (2014) also found adsorption
interaction of HPAM molecules on the surface of NPs which present in
The influence of SiO2 NPs concentration at different shear rates was the liquid. The adsorption interaction of HPAM molecules is caused by
investigated. Eight SiO2-HPAM hybrid dispersions were prepared in hydrogen bonding between HPAM molecules and the surface of the NPs
different concentrations from 0, 0.05 wt%, 0.075 wt%, 0.1 wt%, which can be either electrostatic or hydrophobic adsorption
0.125 wt%, 0.15 wt%, 0.2 wt% and 0.25 wt% in 2̗000 ppm HPAM so- (Kawaguchi, 1994). Samoshina et al. (2003) also reported that there is a
lution at 26 °C, 50 °C, and 70 °C. In these reported experiments, the SiO2 non-electrostatic attraction between PAM molecules and negatively si-
NPs threshold concentration for 2̗000 ppm HPAM solution was 0.1 wt% lica surface area (0.1–0.3 mg/m2). This experiment proves that the
because before and after this critical concentration, it can be seen that presence of NPs in NPs-HPAM hybrid dispersions performs two main
the viscosity of SiO2-HPAM hybrid dispersions were decreased. For functions. They can increase viscosity of HPAM solution at quite low
SiO2-HPAM hybrid dispersion flooding to be effective, SiO2 NPs con- shear rate and also control rheological characteristics of NPs-HPAM
centration must meet critical concentration and for this experiment, the hybrid dispersions in the electrolyte solutions at medium or higher
effective concentration was 0.1 wt% even if the temperature was varied shear rates (0.6–10 s−1).
from room to elevated temperatures, as shown in Fig. 8. Although ad-
dition of SiO2 NPs to HPAM solution can increase viscosity, it could not
3.7. NPs-HPAM hybrid dispersions rheological properties in electrolyte
prevent viscosity loss in the HPAM solution. It should be mentioned
solution
that the effective NP concentration of NPs-HPAM hybrid dispersion
depends on HPAM concentration in the dispersion and varies from one
As shown in Fig. 10(a) HPAM solutions and NPs-HPAM Hybrid
experiment to another.
dispersions had similar rheological characters in DIW which was better
than that in the electrolyte solution. As it was mentioned in section 3.4,
3.6. Effect of NPs type on rheological properties of HPAM solution introducing mono- and divalent ions to HPAM solutions deteriorated
the viscosity of the solution (Fig. 7). The results in Fig. 10(b) indicate
HPAM solution viscosity can be detracted by shear rate due to the degradation of HPAM molecules in SB2 (30̗000 ppm) due to the cation-
mechanical degradation along the polymer molecules. It was observed dipole binding/interaction between the ions and oxygen molecules.
that the viscosity of the 2̗000 ppm of HPAM solution and 0.1 wt% of Hence, binding forces between cations and amino (NH2) group were
NPs-HPAM hybrid dispersions (Al2O3, SiO2, and TiO2) sharply declined quite weaker than interaction between oxygen and cations. Divalent
above 0.612 s−1 and 0.4 s−1, respectively (Fig. 9). However, HPAM ions such as Ca2+ and Mg2+ have higher charge densities than mono-
solution has shown lower viscosity than NPs-HPAM hybrid dispersions valent ions (Na+) and formed stronger binding with amide group than
at lower shear rates (0.1–0.6 s−1). On the other hand, HPAM solution monovalent ions which weakened the NeH and the C]O bonds in

1061
M. Bashir Abdullahi et al. Journal of Petroleum Science and Engineering 172 (2019) 1057–1068

Fig. 3. TEM images of (a) SiO2, (b) Al2O3, (c) TiO2 NPs, and (d) the particle size distribution of NPs in distilled water measured by the DLS method.

HPAM molecular structure. Thus, HPAM solution was degraded che- interaction (Fig. 10(c)). Al2O3-HPAM hybrid dispersion has higher
mically but solution viscosity decreased. Nevertheless, as shown in viscosity than SiO2-HPAM and TiO2-HPAM hybrid dispersion due to the
Fig. 10(b), NPs-HPAM hybrid dispersions have performed better visc- strong ion-dipole bond formed between cations in electrolyte and
osity behaviour than HPAM solution in the presence of electrolyte so- oxygen atoms from Al2O3 NPs. Thus, the amount of cations which at-
lution (SB2). Ion-dipole interactions between SiO2, Al2O3, and TiO2 NPs tack to the HPAM molecules has been reduced to some degree and
with ions (Ca2+, Mg2+, and Na+) are stronger than HPAM-cation chemical degradation on HPAM solution viscosity has been minimized

Fig. 4. (a) Rheogram and (b) flow curves for HPAM concentrations at 26 °C.

1062
M. Bashir Abdullahi et al. Journal of Petroleum Science and Engineering 172 (2019) 1057–1068

Fig. 5. Dependence of HPAM viscosity on salt composition at (a) 26 °C, (b) 50 °C, and (c) 70 °C.

in the presence of NPs. In addition, FTIR analysis deeply described the also might form between the oxygen/nitrogen of amide groups from
mechanism in which the viscosity of HPAM was improved in the pre- HPAM with eOH groups from hydrogen of the SiO2 surface
sence of NPs. (SiOeH⋯NeH or SiOeH⋯OeCNH2). Moreover, it might form between
the oxygen from the SiO2 NPs surface and the hydrogen from HPAM
(SiO⋯HNHeCOeC). The broader peak area appeared in pure HPAM
3.7.1. FTIR tests analysis for HPAM and NPs-HPAM hybrid samples and SiO2-HPAM hybrid samples from 3̗100 to 3̗800 cm−1 which in-
FTIR spectroscopy analysis for HPAM and SiO2-HPAM hybrid sam- dicates the vibration of the eOH, SiOeH, and NeH groups. Therefore,
ples are presented in Fig. 11(a). The results were in the ranges of the possible intermolecular interaction between the HPAM and SiO2
500–4̗000 cm−1 confirmed the potential bonding effects between SiO2 NPs has led to viscosity improvement of the SiO2-HPAM hybrid dis-
NPs and HPAM polymer which improve the viscosity of NPs-HPAM persion.
hybrid dispersions. The peaks at 1̗080 and 800 cm−1 for pure SiO2 NPs The potential bonding between Al2O3 NPs and HPAM polymer
were due to the asymmetric stretching and bending vibrations of monomers was noticed from FTIR spectroscopy result at the ranges of
SieOeSi band as shown in Fig. 11(a). However, no peaks were ob- 500–4̗000 cm−1 (Fig. 12(a)). The AleO stretching in octahedral struc-
served in pure HPAM polymer at the corresponding wavelength of pure ture were discovered at the peaks at 594, 642 and 682 cm−1 for pure
SiO2 NPs. The noticeable shifts were detected on the SiO2-HPAM hybrid Al2O3 sample as shown in Fig. 12(a). Pure HPAM sample does not have
sample at 1̗098 and 792 cm−1 due to the asymmetric stretching and any peak at the corresponding wavelengths. Whereas Al2O3-HPAM
bending vibrations of SieOeSi band which shows interaction effects hybrid sample shows three main peaks at 598, 648 and 700 cm−1
between the SiO2 NPs and HPAM polymer due to hydrolysable covalent which can be assigned due to the asymmetric stretching of AleO band
cross-links in the system (Gao et al., 2016). However, peak at 952 cm−1 in octahedral structure. Another two peaks at 718 cm−1 and 1̗122 cm−1
was found in SiO2-HPAM hybrid sample which was appeared in neither are caused by AleO stretching in tetrahedral structure and symmetric
SiO2 NPs nor HPAM due to the bending vibration of SieOeH by the bending vibration of AleOeH by the hydroxyl on alumina surfaces,
hydroxyl on SiO2 NPs surface as shown in Fig. 11(b). Hydrogen bonding

1063
M. Bashir Abdullahi et al. Journal of Petroleum Science and Engineering 172 (2019) 1057–1068

Fig. 6. HPAM solution viscosity at 400 s−1 in different salt compositions (SB1 & SB2) and temperatures at (a) 26 °C, (b) 50 °C, and (c) 70 °C.

from HPAM (AlO⋯HNHeCOeC). The obvious shifts in the Al2O3-


HPAM hybrid sample (598, 648 and 700 cm−1) compare to that from
pure Al2O3 (594, 642 and 682 cm−1) represent interaction effects be-
tween the Al2O3 NPs and HPAM polymer due to hydrolysable covalent
cross-links in the system (Gao et al., 2016). Consequently, the possible
intermolecular interaction between the HPAM and Al2O3 NPs has im-
proved the viscosity of the Al2O3-HPAM hybrid sample.
FTIR spectrum result was used to establish the possibility of inter-
molecular interaction between TiO2 NPs and HPAM molecules in TiO2-
HPAM hybrid sample as in Fig. 13. The broad peak from 3̗200 to 3̗750
Fig. 7. Ionic interaction between cations (Na+, Ca2+, Mg2+) from brines and in the TiO2-HPAM hybrid sample and HPAM polymer are due to the
C]O from acrylamide of HPAM molecular chain. stretching of the eOH and eNH (amines) groups. The detected peaks
from 1600 cm−1–1750 cm−1 are assigned to the bending vibration of
respectively (Djebaili et al., 2015; Hu et al., 2017). Hydrogen bonding the eOH group. In the spectrum curve of pure TiO2, the peak at
was formed between the oxygen/nitrogen of amide groups from HPAM 545.44 cm−1 shows stretching vibration of TieO (Sala and Trifirò,
with eOH groups from hydrogen of the Al2O3 surface (AlOeH⋯NeH or 1974) and peak at 1406 cm−1 reveal the stretching vibrations of
AlOeH⋯OeCNH2) in Al2O3-HPAM hybrid sample (Fig. 12(b)). It might TieOeTi. The TiO2-HPAM hybrid sample has demonstrated two peaks
formed between the oxygen from the Al2O3 surface and the hydrogen at 539 cm−1 which shows stretching vibration of TieO and 1̗406 cm−1

1064
M. Bashir Abdullahi et al. Journal of Petroleum Science and Engineering 172 (2019) 1057–1068

Fig. 8. Effect of SiO2 NPs Concentrations on SiO2-HPAM hybrid dispersion viscosity at (a) 26 °C, (b) 50 °C, and (c) 70 °C.

caused by stretching vibrations of TieOeTi which the first peak has


been shifted noticeably compare to that of pure TiO2 NPs. The other
peak, on other hand has not been shifted. The shift at 539 cm−1 can be
contributed to hydrolysable covalent cross-links in the TiO2-HPAM
hybrid sample.

3.8. Displacement experiment

Final objective of this experimental study is to determine the per-


formance of NPs-HPAM hybrid dispersions as an EOR technique and
comparing their recoveries with HPAM solution flooding at room
temperature. Parameters such as flow rate and pore volume injected
remain the same for all flooding processes. As explained in the dis-
placement test section 4 PVs of brine was flooded into the sandpack
holder #1 to simulate secondary oil recovery. As it is shown in Fig. 14,
after injecting 1.5 PVs of brine (30̗000 ppm) the total oil recovery was
not increased more than 25.1% of initial oil in place which is due to the
water break through that occurred at 1.5 PVs. The next step was in-
Fig. 9. Effect of NPs type on rheological properties of HPAM solution.
jecting 1.5 PVs brine followed by 2.5 PVs of HPAM solution (2̗000 ppm)
and NPs-HPAM hybrid dispersions into the sandpack holders #2, #3,
#4, and #5, respectively. The results of displacement test from sand-
pack holder #2 demonstrated that utilizing HPAM solution caused

1065
M. Bashir Abdullahi et al. Journal of Petroleum Science and Engineering 172 (2019) 1057–1068

Fig. 10. Viscosity Vs Shear Rate for (a) HPAM and NPs-HPAM Hybrid Dispersions without electrolyte (DIW), (b) at 3 wt% electrolyte solution (SB2), and (c) potential
interaction between NPs and mono- and divalent ions at 26 °C.

Fig. 11. (a) FTIR spectra for 0.1 wt% SiO2 NPs, 0.2 wt% HPAM polymer, and 0.2 wt% HPAM/0.1 wt% SiO2 Hybrid and (b) potential interaction between SiO2 NP and
HPAM molecule.

1066
M. Bashir Abdullahi et al. Journal of Petroleum Science and Engineering 172 (2019) 1057–1068

Fig. 12. (a) FTIR spectra for 0.1 wt% Al2O3 NPs, 0.2 wt% HPAM polymer, and 0.2 wt% HPAM/0.1 wt% Al2O3 Hybrid and (b) potential interaction between Al2O3 NP
and HPAM molecule.

50.2% oil recovery where the ultimate oil recovery was 75.3%. It was
found that using SiO2-HPAM hybrid dispersion in the sandpack holder
#3 increased oil recovery 58.9% more than waterflooding and total
recovery was 84%. The fourth displacement test was done in the
sandpack holder #4 via Al2O3-HPAM hybrid dispersion and the total oil
recovery was 90%. The last displacement test was performed by TiO2-
HPAM dispersion with the ultimate oil recovery of 80%. The details of
oil recoveries after each displacement test are drawn in Fig. 14.
The performance of HPAM solution flooding depends on the stabi-
lity of HPAM solution in harsh condition. As mentioned earlier, the
presence of mono- and divalent ions in the solution spoils the rheolo-
gical properties of the HPAM solution. Therefore, the efficiency of
HPAM solution flooding is lessened and ineffective. It was clearly de-
tected and found that employing NPs in HPAM solution not only es-
calates the salt tolerance of HPAM molecules, but also could increase
the oil recovery compare to HPAM solution.

4. Conclusion

This experimental study explored the impact of different type of NPs


(SiO2, Al2O3, and TiO2) on HPAM solution rheological properties under
different salinities and temperatures. Fig. 14. Residual oil recoveries for water, HPAM, and NPs-HPAM hybrid dis-
persions floodings.
1. There was a critical concentration for SiO2 NPs established for this

Fig. 13. FTIR spectra for 0.1 wt% TiO2 NPs, 0.2 wt% HPAM polymer, and 0.2 wt% HPAM/0.1 wt% TiO2 Hybrid.

1067
M. Bashir Abdullahi et al. Journal of Petroleum Science and Engineering 172 (2019) 1057–1068

experiment (0.1 wt %) in 2̗000 ppm HPAM under different tem- Bifunctional ultraviolet/ultrasound responsive composite TiO 2/polyelectrolyte mi-
peratures and shear rates. It was noticed that below and beyond the crocapsules. Nanoscale 8, 5170–5180. https://doi.org/10.1039/C5NR06666B.
Giese, S.W., Powers, S.E., 2002. Using polymer solutions to enhance recovery of mobile
critical NPs concentration, viscosity of SiO2-HPAM hybrid disper- coal tar and creosote DNAPLs. J. Contam. Hydrol. 58, 147–167.
sion is low due to low concentration of NPs in the dispersion and Hu, Z., Haruna, M., Gao, H., Nourafkan, E., Wen, D., 2017. Rheological properties of
adsorption interaction between HPAM molecules and SiO2 NPs partially hydrolyzed polyacrylamide seeded by nanoparticles. Ind. Eng. Chem. Res.
56, 3456–3463. https://doi.org/10.1021/acs.iecr.6b05036.
surface, respectively. Among the NPs-HPAM hybrid dispersions, Kawaguchi, M., 1994. Rheological properties of silica suspensions in polymer solutions.
Al2O3-HPAM hybrid dispersion showed higher viscosity followed by Adv. Colloid Interface Sci. 53, 103–127. https://doi.org/10.1016/0001-8686(94)
SiO2-HPAM and TiO2-HPAM hybrid dispersions under high salinity 00214-2.
Lopes, L.F., Silveira, B.M.O., Moreno, R.B.Z.L., 2014. Rheological evaluation of HPAM
and high temperature conditions. fluids for EOR applications. Int. J. Eng. Technol. IJET-IJENS 14, 35–41. https://doi.
2. It was found that NPs increased HPAM solution thermal stability and org/10.1.1.659.6456.
salt tolerance. The results indicated that NPs were able to control Maghzi, A., Kharrat, R., Mohebbi, A., Ghazanfari, M.H., 2014. The impact of silica na-
noparticles on the performance of polymer solution in presence of salts in polymer
viscosity loss by forming strong intermolecular interaction with
flooding for heavy oil recovery. Fuel 123, 123–132.
cations and strengthen polymer molecules repulsion thereby visc- Maghzi, A., Mohebbi, A., Kharrat, R., Ghazanfari, M.H., 2013. An experimental in-
osity loss reduced. vestigation of silica nanoparticles effect on the rheological behavior of poly-
3. The FTIR spectral data confirmed that the formation of hydrogen acrylamide solution to enhance heavy oil recovery. Petrol. Sci. Technol. 31, 500–508.
https://doi.org/10.1080/10916466.2010.518191.
bonds between the carbonyl groups in HPAM and the NPs functional Muller, G., 1981. Thermal stability of high-molecular-weight polyacrylamide aqueous
groups on the surface of SiO2, Al2O3, and TiO2 NPs which caused solutions. Polym. Bull. 5, 31–37. https://doi.org/10.1007/BF00255084.
improvement in rheological performance of HPAM solution. Nguyen, P., Fadaei, H., Sinton, D., 2014. Pore-scale assessment of nanoparticle-stabilized
CO2 foam for enhanced oil recovery. Energy Fuels 28, 6221–6227.
4. It was discovered that NPs-HPAM hybrid dispersion with excellent Ogolo, N., Olafuyi, O., Onyekonwu, M., 2012. Enhanced oil recovery using nanoparticles.
performance in harsh condition can produce more amount of oil Saudi Arab. Sect. Tech. Symp. Exhib. 9. https://doi.org/10.2118/160847-MS.
from the porous medium. The ultimate oil recoveries achieved in Olajire, A.A., 2014. Review of ASP EOR (alkaline surfactant polymer enhanced oil re-
covery) technology in the petroleum industry: prospects and challenges. Inside
porous media at room temperature by Al2O3-HPAM, SiO2-HPAM, Energy. https://doi.org/10.1016/j.energy.2014.09.005.
and TiO2-HPAM hybrid dispersions were 90%, 84%, and 80%, re- Ragab, A.M.S., Hannora, A.E., 2015. An experimental investigation of silica nano particles
spectively. for enhanced oil recovery applications. In: SPE North Africa Tech. Conf. Exhib, .
https://doi.org/10.2118/175829-MS.
Sala, F., Trifirò, F., 1974. Oxidation catalysts based on tin-antimony oxides. J. Catal. 34,
Acknowledgment 68–78. https://doi.org/10.1016/0021-9517(74)90013-X.
Samoshina, Y., Diaz, A., Becker, Y., Nylander, T., Lindman, B., 2003. Adsorption of ca-
tionic, anionic and hydrophobically modified polyacrylamides on silica surfaces.
The authors would like to thank the Ministry of Higher Education of
Colloids Surfaces A Physicochem. Eng. Asp. 231, 195–205. https://doi.org/10.1016/
Malaysia for their financially support under the project numbers of j.colsurfa.2003.09.005.
R.J130000.7846.4F946. The authors are also grateful to department of Seright, R.S., Campbell, A., Mozley, P., Han, P., 2010. Stability of partially hydrolyzed
Petroleum Engineering at Universiti Teknologi Malaysia for providing polyacrylamides at elevated temperatures in the absence of divalent cations. SPE J.
15, 341–348. https://doi.org/10.2118/121460-PA.
adequate facilities, guidance and mentoring to conduct this study and ShamsiJazeyi, H., Miller, C.A., Wong, M.S., Tour, J.M., Verduzco, R., 2014.
also to the Research Management Center (RMC) of Universiti Teknologi Polymer‐coated nanoparticles for enhanced oil recovery. J. Appl. Polym. Sci. 131.
Malaysia for financial supports under the project number of Sheng, J.J., 2014. Critical review of low-salinity waterflooding. J. Petrol. Sci. Eng.
https://doi.org/10.1016/j.petrol.2014.05.026.
Q.J130000.2456.14H50. Silveira, B.M.O., Lopes, L.F., Moreno, R.B.Z.L., 2016. Rheological approach of HPAM
solutions under harsh conditions for EOR applications. Int. J. Eng. Technol. IJET-
References IJENS 16, 1–8.
Tovar, F.D., Barrufet, M.A., Schechter, D.S., Baker, H., bp, Chevron, ConocoPhillips, Sasol,
et al., 2014. Long term stability of acrylamide based polymers during chemically
Ahmad Ramazani, S.A., Nourani, M., Emadi, M.A., Esfad, N.J., 2010. Analytical and ex- assisted CO2WAG EOR. In: 19th SPE Improved Oil Recovery Symposium, IOR 2014,
perimental study to predict the residual resistance factor on polymer flooding process pp. 310–318.
in fractured medium. Transport Porous Media 85, 825–840. https://doi.org/10. Wilton, R.R., 2015. Rheology and Flow Behaviour of Non-newtonian, Polymeric Gluids in
1007/s11242-010-9594-8. Capillaric and Porous Media: Aspects Related to Polymer Flooding for Enchance
Alomair, O.A., Matar, K.M., Alsaeed, Y.H., 2015. Experimental study of enhanced-heavy- Recovery of Heavy Oil.
oil recovery in berea sandstone cores by use of nanofluids applications. SPE Reservoir Yahya, G.O., Ali, S.A., Al‐Naafa, M.A., Hamad, E.Z., 1995. Preparation and viscosity
Eval. Eng. 18, 387–399. https://doi.org/10.2118/171539-PA. behavior of hydrophobically modified poly(vinyl alcohol) (PVA). J. Appl. Polym. Sci.
Bayat, A.E., Junin, R., Piroozian, A., Hokmabadi, M., 2014. Impact of metal oxide na- 57, 343–352. https://doi.org/10.1002/app.1995.070570311.
noparticles on enhanced oil recovery from limestone media at several temperatures. Yousefvand, H., Jafari, A., 2015. Enhanced oil recovery using polymer/nanosilica.
Energy Fuels 28, 6255–6266. https://doi.org/dx.doi.org/10.1021/ef5013616. Procedia Mater. Sci. 11, 565–570. https://doi.org/10.1016/j.mspro.2015.11.068.
Bhardwaj, P., Singh, S., Singh, V., Aggarwal, S., Mandal, U.K., 2008. Nanosize poly- Zargartalebi, M., Kharrat, R., Barati, N., 2015. Enhancement of surfactant flooding per-
acrylamide/SiO2 composites by inverse microemulsion polymerization. Int. J. Polym. formance by the use of silica nanoparticles. Fuel 143, 21–27. https://doi.org/10.
Mater. Polym. Biomater. 57, 404–416. https://doi.org/10.1080/ 1016/j.fuel.2014.11.040.
00914030701729156. Zeyghami, M., Kharrat, R., Ghazanfari, M.H., 2014. Investigation of the applicability of
Cheraghian, G., 2016. Effect of nano titanium dioxide on heavy oil recovery during nano silica particles as a thickening additive for polymer solutions applied in EOR
polymer flooding. Petrol. Sci. Technol. 34, 633–641. https://doi.org/10.1080/ processes. Energy Sources, Part A Recover. Util. Environ. Eff. 36, 1315–1324. https://
10916466.2016.1156125. doi.org/10.1080/15567036.2010.551272.
Djebaili, K., Mekhalif, Z., Boumaza, A., Djelloul, A., 2015. XPS, FTIR, EDX, and XRD Zhang, H., Nikolov, A., Wasan, D., 2014. Enhanced oil recovery (EOR) using nanoparticle
analysis of Al2O3 scales grown on PM2000 alloy. J. Spectrosc. 2015. https://doi. dispersions: underlying mechanism and imbibition experiments. Energy Fuels 28,
org/10.1155/2015/868109. 3002–3009.
Gao, H., Wen, D., Tarakina, N.V., Liang, J., Bushby, A.J., Sukhorukov, G.B., 2016.

1068

Vous aimerez peut-être aussi