Download as pdf or txt
Download as pdf or txt
You are on page 1of 56

Journal Pre-proof

Phenomenological Study of the Micro- and Macroscopic Mechanisms during Polymer


Flooding with SiO2 Nanoparticles

Oveimar Santamaria, Sergio H. Lopera, Masoud Riazi, Mario Minale, Farid B. Cortés,
Camilo A. Franco
PII: S0920-4105(20)31189-X
DOI: https://doi.org/10.1016/j.petrol.2020.108135
Reference: PETROL 108135

To appear in: Journal of Petroleum Science and Engineering

Received Date: 13 June 2020


Revised Date: 10 November 2020
Accepted Date: 12 November 2020

Please cite this article as: Santamaria, O., Lopera, S.H., Riazi, M., Minale, M., Cortés, F.B., Franco,
C.A., Phenomenological Study of the Micro- and Macroscopic Mechanisms during Polymer Flooding
with SiO2 Nanoparticles, Journal of Petroleum Science and Engineering, https://doi.org/10.1016/
j.petrol.2020.108135.

This is a PDF file of an article that has undergone enhancements after acceptance, such as the addition
of a cover page and metadata, and formatting for readability, but it is not yet the definitive version of
record. This version will undergo additional copyediting, typesetting and review before it is published
in its final form, but we are providing this version to give early visibility of the article. Please note that,
during the production process, errors may be discovered which could affect the content, and all legal
disclaimers that apply to the journal pertain.

© 2020 Published by Elsevier B.V.


CRediT author statement

Conceptualization All authors

Methodology All authors

Formal análisis Oveimar Santamaria, Sergio H. Lopera, Mario Minale, Farid B. Cortés, and Camilo A.
Franco.

Investigation All authors

Writing – Oveimar Santamaria, Farid B. Cortés, Mario Minale

Writing - Review & Editing - Oveimar Santamaria, Mario Minale, Farid B. Cortés, and Camilo A.
Franco.

of
ro
-p
re
lP
na
ur
Jo
Phenomenological Study of the Micro- and Macroscopic Mechanisms during Polymer

Flooding with SiO2 Nanoparticles

Oveimar Santamaria1,2, Sergio H. Lopera2, Masoud Riazi3, Mario Minale4, Farid B. Cortés 1

and Camilo A. Franco1,*

1
Grupo de Investigación en Fenómenos de Superficie—Michael Polanyi, Departamento

de Procesos y Energía, Facultad de Minas, Universidad Nacional de Colombia, Sede

of
Medellín, 050034 Medellín, Colombia.

ro
2
Grupo de Investigación en Yacimientos de Hidrocarburos, Departamento de Procesos y
-p
Energía, Facultad de Minas, Universidad Nacional de Colombia, 050034 Medellín,
re
Colombia.
lP

3
Enhanced Oil Recovery (EOR) Research Centre, IOR/EOR Research Institute, Shiraz
na

University, Shiraz, Iran.


ur

4
Department of Engineering, University of Campania Luigi Vanvitelli, via Roma 29,
Jo

81031 Aversa (CE), Italy.

* Corresponding author: caafrancoar@unal.edu.co; fbcortes@unal.edu.co


Phenomenological Study of the Micro- and Macroscopic Mechanisms during Polymer

Flooding with SiO2 Nanoparticles

Oveimar Santamaria1,2, Sergio H. Lopera2, Masoud Riazi3, Mario Minale4, Farid B. Cortés 1

and Camilo A. Franco1,*

1
Grupo de Investigación en Fenómenos de Superficie—Michael Polanyi, Departamento

de Procesos y Energía, Facultad de Minas, Universidad Nacional de Colombia, Sede

of
Medellín, 050034 Medellín, Colombia.

ro
2
Grupo de Investigación en Yacimientos de Hidrocarburos, Departamento de Procesos y
-p
Energía, Facultad de Minas, Universidad Nacional de Colombia, 050034 Medellín,
re
Colombia.
lP

3
Enhanced Oil Recovery (EOR) Research Centre, IOR/EOR Research Institute, Shiraz
na

University, Shiraz, Iran.


ur

4
Department of Engineering, University of Campania Luigi Vanvitelli, via Roma 29,
Jo

81031 Aversa (CE), Italy.

* Corresponding author: caafrancoar@unal.edu.co; fbcortes@unal.edu.co

Abstract

The incorporation of SiO2 nanoparticles in polymer solutions for oil recovery has

generated considerable interest in recent research. Although this approach improves the

polymer performance, little evidence is available about the recovery mechanisms that

intervene in the polymeric nanofluid suspension or polymer nanofluid. Therefore, the


purpose of this study is to describe the microscopic and macroscopic mechanisms

involved during oil recovery with polymeric nanofluid flooding. Changes in the rheology

and capillary forces (contact angle and interfacial tension) in the polymeric solution due to

the incorporation of SiO2 nanoparticles were evaluated. Coreflooding tests were made to

evaluate the retention, apparent viscosity, and impact on the relative permeability. Finally,

macroscopic oil recovery was measured by displacement tests in a quarter 5-spot system,

and microscopic displacement was evaluated by a microfluidic test. Results indicated that

of
the conservation of a favorable mobility ratio in polymeric nanofluids along the porous

ro
medium is the phenomenon that most impacts macroscopic oil recovery. This behavior
-p
was associated with a reduction in loss retention and the conservation of the apparent
re
viscosity in the porous medium. A direct relationship was observed between the
lP

concentration of SiO2 nanoparticles and the viscoelastic behavior of the polymer and the
na

water-wet condition on the surface exposed to the nanofluid. Moreover, residual oil

distribution in microscopic displacement revealed that the reduction in residual oil mainly
ur

originated from the decrease in the size of the oil clusters. Consequently, the evidence
Jo

suggests that microscopic oil displacement in polymeric nanofluid is attributed to the

reduction in the capillary forces and increases in the viscoelastic nature of the polymer.

Results in this work indicated that the additional oil recovery obtained by polymeric

nanofluid flooding is generated by the improvement of macroscopic recovery mechanisms

and the activation of microscopic mechanisms.

Keywords: Polymeric nanofluids, SiO2 Nanoparticles, Microscopic oil displacement,

Macroscopic oil displacement.


Graphical abstract

of
ro
-p
re
lP
na
ur
Jo
1 1. Introduction

2 Partially hydrolyzed polyacrylamide (HPAM) is considered the polymer with better

3 performance in the implementation of enhanced oil recovery (EOR) processes.

4 Polyacrylamide (PAM) is partially hydrolyzed by reacting with a base such as sodium or

5 potassium hydroxide or sodium carbonate [1]. The process is controlled to incite partial

6 hydrolysis [1,2]. Hydrolysis converts some amide groups (CONH2) to carboxyl groups

of
7 (COO−) [1]. The lineal conformation allows for the most flexibility, which is critical for

ro
8 transport inside porous media. Atactic conformation in COO− groups generates self-

9
-p
repulsion between sections of the chain; as a consequence, the viscosifying power is
re
10 greater, with less adsorption over the rock surface [3]. As a consequence, oil is produced
lP

11 due to a favorable mobility ratio to improve sweep efficiency. However, there are many

12 different views concerning the recovery mechanisms of polymer flooding.


na

13 In immiscible fluid displacement in porous media, the micro- and macroscopic forces
ur

14 govern oil recovery in flooding processes. Two-phase displacement can be characterized in


Jo

15 homogeneous systems by two dimensionless numbers: the capillary number (Nc) and the

16 mobility ratio (M) [4,5]. The capillary number relates viscous to capillary forces, and the

17 mobility ratio M is defined as the ratio of the displacing phase mobility to the displaced

18 phase mobility. At the pore scale, displacement mechanisms under drainage conditions are

19 originated by a reduction in capillary forces such as interfacial tensions (IFT) reduction or

20 wettability alteration [4]. At macro conditions, uniform volumetric sweeping is reached at a

21 favorable mobility ratio resulting in a stable (non-fingering) oil sweep. For polymer

22 flooding, it is assumed that the oil recovered is uniquely due to volumetric sweeping; that
1 is, it cannot decrease the residual oil saturation within the swept area. However, several

2 experimental and field studies have confirmed that polymer solutions increase microscopic

3 oil displacement efficiency [6–9]. Wang Demin et al. [10] analyzed the mechanism of how

4 polymer solutions increase microscopic displacement efficiency and concluded that the

5 result was due to its viscoelastic effects.

6 The flow of a non-Newtonian fluid in a porous medium can be studied with an

of
7 averaged equation similar to that proposed by Darcy [11]. However, the non-linearity of a

ro
8 fluid constitutive equation introduces some difficulties in the modeling that are mainly

9 -p
related to the impossibility of decoupling the porous medium and fluid parameters from the
re
10 drag terms of the equation of motion [12]. Minale [13,14] showed that decoupling is

11 theoretically possible for a viscoelastic fluid showing constant viscosity and normal stress
lP

12 differences, and that an extra drag resistance tensor is introduced: a non-Newtonian


na

13 permeability tensor. For shear-thinning fluids, a general theoretical demonstration is


ur

14 missing. However, there is consensus on the use of a generalized Darcy’s equation where
Jo

15 the permeability is that defined for a Newtonian fluid and where all non-Newtonian effects

16 are arranged into a “porous medium viscosity” [15]. The porous medium viscosity, also

17 called apparent viscosity, is used in place of the fluid viscosity and will be described by the

18 same constitutive equation. The choice of the apparent shear rate in the porous medium will

19 affect the fluid constitutive parameters [16].

20 Nanotechnology has attracted widespread interest for use in recovery processes.

21 Nanoparticles (NPs) incorporate new properties into traditional technologies applied in

22 EOR [17–19]. Polymer-based nanofluids or polymeric nanofluids consist of a combination

23 of nanoparticles with polymeric solutions [20–28]. Polymeric nanofluids are of great


1 interest because they combine the advantages of inorganic NPs and organic polymers,

2 generating synergy between the best of each of the two materials [28]. Using Fourier

3 transform infrared (FTIR) spectra data, Hu et al. [26], proposed the formation of hydrogen

4 bonds between the carbonyl groups in the polymer and the silanol on the surface of SiO2

5 NPs, which improves the rheological properties of polymeric nanofluids at high

6 temperature and salinity. Numerous experimental works have corroborated that NPs

7 contribute to forming a more stable nanoparticle–polymer matrix [21,22,24,26,29–34].

of
8 Researchers have proposed that in a polymer chemically linked to NPs (polymer-grafted

ro
9 NPs, or “PGN”), repulsion (electrosteric or steric repulsion) reduced hydrophobic–

10
-p
hydrophobic interaction between the PGN and the rock surface [35–37]. Similar results
re
11 have been reported for polymeric nanofluid suspension (PNS) [38–41]. Although numerous
lP

12 studies focus on polymer solution stability in the presence of NPs, an inconclusive


na

13 understanding remains of the source of additional oil recovered during coreflooding tests

14 based on identifying the influence of micro- and macroscopic mechanisms. It is assumed


ur

15 that NPs only benefit the favorable mobility ratio, despite the widespread documentation
Jo

16 that nanofluids can increase the oil produced by enhanced fluid–fluid interactions such as

17 IFT reduction or rock–fluid interactions such as wettability alteration [18,19,42–45].

18 Hence, the purpose of this study is to describe the micro- and macroscopic mechanisms

19 that act in oil displacement during polymeric nanofluid flooding. Therefore, this work

20 evaluates the impact on rheology and capillary forces (contact angle and interfacial tension)

21 in HPAM solutions due to the incorporation of SiO2 nanoparticles. Coreflood tests were

22 performed to evaluate the retention, apparent viscosity, and effect on the relative

23 permeability. Finally, macroscopic oil recovery was measured by displacement tests in a


1 quarter 5-spot system, and microscopic displacement was assessed by a microfluidic test.

2 Experimental results show that a favorable mobility ratio is generated by SiO2 NPs and

3 conserved in the porous media as an effect of reduced retention. Also, microscopic oil

4 displacement was generated by a reduction in capillary forces and an increase in the

5 viscoelastic nature of the polymer.

6 2. Materials and Methods

of
7 2.1. Materials:

ro
8 2.1.1. Fluids and nanoparticle: -p
re
9 The polymer used was HPAM, which has a molecular weight between 6–8 MDa and a
lP

10 hydrolysis percentage of 30%. The sample of polymer was supplied by Nalco S.A

11 (Colombia). NPs are commercially available and were provided by Sigma-Aldrich Co.,
na

12 (St. Louis, MO, USA). The nanomaterial is a hydrophilic fumed silica NP (SiO2-NPs). The
ur

13 surface area, which was obtained using the Brunauer–Emmett–Teller theory (BET), is
Jo

14 389 m2 g−1, and the average particle size is 7 nm. Complementary information about the

15 nanoparticles was reported by Franco et al. [46,47]. Consistency with previous studies was

16 maintained by using the same materials and methods to minimize possible differences and

17 distortions of the measurements [31,48].

18 2.1.2. Reservoir fluids:

19 Samples and experimental conditions were obtained from a field in the Magdalena Medio

20 Valley of Colombia. Mature fields are located in this geological structure with the potential

21 for EOR processes. The oil used in this study has an API gravity of 21°, viscosity of
1 0.024 Pa s at 52 °C and the total acid number was 0.71 mg KOH·g−1. The content of

2 saturated, aromatics, resins, and asphaltenes were 34.3%, 36.7%, 24.6%, and 4.13%,

3 respectively. Synthetic brine was prepared with chloride salts equivalent to the reservoir

4 brine, which was a solution of 5000 mg L−1 KCl (98%, Sigma-Aldrich, St. Louis, MO,

5 USA) in deionized water. KCl provides a way to supply equivalent chloride salts and

6 minimize the impact of monovalent ions (Na+) on the stability of polymers [49,50].

7 Potassium thiocyanate (Merck KGaA, Colombia) was used as a tracer for the determination

of
8 of inaccessible pore volume by effluent analysis.

ro
9 2.1.3. Porous media: -p
re
10 Outcrop rocks and sandpacks were employed for estimating the impact of micro–macro
lP

11 forces that control oil displacement in polymeric nanofluid flooding. The main

12 characteristics of porous media are listed in Table 1. Artificial sandpacks were constructed
na

13 with Ottawa sand (Minercol S.A.S, Bogotá D.C., Colombia). The mineral composition is
ur

14 quartz, whose surfaces are made up of hydrated silica tetrahedra (Si-O-H) [51].
Jo

15 Consolidated samples were composed of sandstone with <5% clay (3% kaolinite and 2%

16 chlorite and illite), and outcrop rocks were obtained from the La Paz formation of the

17 Magdalena Medio Valley of Colombia.

18 Building sandpacks is an in-house process developed to guarantee similar

19 petrophysical characteristics and repeatability in the baseline [52]. A combination of grain

20 sizes sieved into 30–40, 50–60, and 100–120 were used in proportions of 40%, 30%, and

21 30%, respectively. With these proportions, sandpacks were adjusted to the porosity and

22 permeability representative of the reference reservoir. To eliminate impurities and promote

23 a neutral wettability, the sand was immersed in 5% HCl (Sigma-Aldrich, St. Louis, MO,
1 USA) for 24 h, after which it was thoroughly washed with water and dried in a ventilated

2 oven at 60 °C. Finally, the sand was packed and compacted at a pressure of 25 MPa.

3 Table 1. Displacement tests and characteristics of porous media.

Nanoparticle
Test Porous Permeability Porosity Soi*
Test concentration
No. medium (mD) (%) (%)
(mg·L-1)
Apparent viscosity in the
1 Sandpack 460 21 0 N/A
porous medium
Apparent viscosity in the

of
2 Sandpack 512 24.2 3000 N/A
porous medium

ro
Rock Polymer retention -
3 72.4 11 0 N/A
outcrop Inaccessible Pore Volume

4
Rock Polymer retention -
-p 72.4 11 3000 N/A
re
outcrop Inaccessible Pore Volume

Rock
lP

5 Relative permeability 123 16.2 0 77


outcrop
Rock
6 Relative permeability 123 16.2 3000 77.1
outcrop
na

Oil recovery
7 Sandpack** 661 25 From 0 to 2000 78
SiO2-water nanofluid
ur

Oil recovery
8 Sandpack** 672 26 From 0 to 2000 81.2
SiO2-polymer nanofluid
Jo

Oil recovery
9 Micromodel Polymeric solution and 5710 70.9 0 - 3000 100
polymeric nanofluids
4 * Initial oil saturation
5 ** Packed in a quarter 5-spot system.
6

7 2.2. Experimental set-up:

8 2.2.1. Coreflood system:

9 A conventional coreflood setup was employed for linear displacement tests (Fig. 1). The

10 system includes a positive displacement pump, accumulator cylinders, and a back-pressure


1 system. The injection system is place inside a heating oven to guarantee reservoir

2 temperature in the rock and fluids. The effective pressure ( = −

3 ) was fixed at 10.4 MPa.

dp1
dp1

Heated zone

of
Core 1

ro
Back
pressure
T T

Oil
-p
re
Overburden pump

Brine
lP
na

Effluent sample
Dual piston pump
ur

12,34

Balance
Jo

Polymer reservoir

Dual piston pump


12,34

Balance
4

5 Figure 1. Schematic representation of the coreflood system for linear displacement tests.

6 2.2.2. Quarter 5-Spot system:

7 Physical models for displacement tests are an experimental representation of any process in

8 reservoir conditions [53–55]. For oil displacement under reservoir flow conditions, quarter

9 5-spot systems have been used to estimate the macroscopic sweep displacement (Fig. 2)
1 [56,57]. The present study employed a cubic cell (250 × 150 × 50 mm), which represents a

2 quarter of a five-point arrangement with a producer well and an injector well. The sieved

3 quartz sand is packed and compacted into the cell and hermetically sealed. The

4 experimental configuration includes a constant rate pump (Marterflex Cole-Palmer), a

5 pressure transductor, and a system for the recollection of effluents. Displacement tests in

6 the radial flow system were performed at reservoir temperature (temperature = 52 °C).

of
dp1
dp1

ro
Heated zone

-p Injection point
re
Sandpack
lP

Production point Back


pressure
T T
Radial flow cell
na

Oil
ur

Brine
Jo

Effluent sample
Dual piston pump

12,34

Balance

Polymer reservoir

Dual piston pump


12,34

Balance
7

8 Figure 2. Schematic representation of the quarter 5-spot system for radial displacement

9 tests.
1 2.2.3. Microfluidic device:

2 Fig. 3 shows a schematic representation of microfluidic devices and the experimental setup.

3 The microfluidic porous media devices were fabricated with a chemically inert polymer,

4 polydimethylsiloxane (PDMS). Details of the fabrication process have been described in

5 previous work [58]. A microchannel network was designed using Layout Editor Software

6 (Germany). For the fabrication, a mixture of epoxy resin and a curing agent (Cristal-Tack,

of
7 Novarchem, Argentina) was poured onto a female photopolymer mold to replicate the

ro
8 design in high relief. The dimensions for the microfluidic devices were 26.5 × 12.7 mm for

9 -p
the length and width, respectively. The pattern and the physical descriptions of the
re
10 microfluidic devices have a pore throat size of 0.1–0.6 mm.
lP

Camera
dp1
na

Micromodel
ur
Jo

Effluent sample

12,34

Balance

OEM syringe pump


11

12 Figure 3. Schematic representation of the micromodel setup.

13
1 2.3. Methods:

2 The experiments were conducted to estimate the impact of SiO2-NPs on macro-microscopic

3 oil displacement at static conditions (without flow in the porous media) and dynamic

4 conditions (during a flow in the porous media). Fig. 4 illustrates the experimental workflow

5 followed in this research.

Static tests Coreflooding tests Oil recovery tests

of
ro
Apparent viscosity

Macro-Microscopic recovery mechanisms


Rheology evaluation -p
Polymer retention
Macroscopic oil
displacement:
Quarter 5-spot system
re
Contact angle
Inaccessible pore
lP

volume

Interfacial tension Residual resistance


na

factor
Microscopic oil
displacement:
Microfluidic test

Relative permeability
ur

curves
Jo

7 Figure 4. Experimental workflow for the evaluation of recovery mechanisms.

8 Displacement forces that occur in homogeneous porous media are characterized by two

9 dimensionless numbers, the capillary number (for microscopic displacement) and the

10 mobility ratio (for macroscopic displacement) [4,59]. The capillary number ( ) relates

11 viscous to capillary forces and is defined as = ⁄ , where ( ) and


!
12 ( · ) are the viscosity and velocity of the advancing non-wetting fluid,
!
13 respectively, and ( · ) is the interfacial tension between the non-wetting
1 (subscript n) and wetting fluid (subscript w). is the fluid–fluid contact angle. The

2 mobility ratio (M) is defined as the ratio of the displacing phase mobility (" ) to the

3 displaced phase mobility (" ):# = " ⁄" , where the subscripts u and d represent upstream

4 and downstream, respectively [5]. Favorable control mobility is reached at a mobility ratio

5 equal to or less than one [5], while an increase in the capillary numbers generates

6 mobilization of the oil residual by capillary forces [59]. Displacement tests were developed

7 in porous media with different geometries to corroborate the impact during a flow through

of
8 the pores.

ro
9 2.3.1. Polymeric nanofluid preparation: -p
re
10 Polymeric nanofluid samples are conformed by SiO2 nanoparticles and the polymeric
lP

11 solution. The order of addition was selected considering the contributions discussed by

12 Giraldo et al. [31]. Polymeric solutions were prepared following the API 63 Standard
na

13 “Practices for evaluation of polymers used in enhanced oil recovery operations” (API),
ur

14 regarding the recommendations for agitation speed and total solubilization time [60]. The
Jo

15 polymer concentration in solutions was 500 mg L−1 based on the oil viscosity and

16 permeabilities (quarter 5-spot system), obtaining a favorable mobility ratio (M ≈ 1). NPs

17 were uniformly dispersed in deionized water and mixed by ultrasonic agitation for 30 min.

18 Nanoparticle dispersed in water was carefully homogenized with the polymer solution, it

19 was stored in the absence of light and heat for 48 h to guarantee their homogeneity and

20 stability.

21 2.3.2. Rheology tests:


1 Rheology measurements were performed in a Kinexus pro+ rotary rheometer (Malvern

2 Instruments, Worcestershire, UK) equipped with a Peltier cartridge for temperature control

3 (52 °C). All rotational rheology measurements were performed at shear rates from 1 to

4 100 s−1, applied for 600 s. The range of shear rates is based on typical values in the well-

5 face (~100 s−1) and the depth formation (~10 s−1) [30]. Concentric cylinders were

6 employed, with a radius for bod (OD) and cup (ID) of 13.329 and 14.463 mm, respectively.

7 Although the Carreau model is commonly used to describe the rheology of polymer

of
8 solutions [61], our samples did not show the Newtonian region in the experimental

ro
9 conditions we used, and only the shear-thinning region was observed. Thus, the power-law

10
-p
model (or Ostwald–de Waele model) was used to analyze the data [62]. The equation for
re
11 interpolation is shown below:
lP

(1)
na

12 Where k is called the consistency index, and n is the power-law index. The root-mean-
ur

13 square error (RSME %) was employed to estimate the fit model [48].
Jo

14 For oscillatory rheology tests, parameters for measurements were selected as proposed

15 by Aliabadian et al. [20,29], who demonstrated that for HPAM solutions under these

16 conditions, flow instability is not presented, nor torque overload or wall-slip effects. For

17 frequency sweep tests, the angular frequency varied between 0.1 and 10 rad s−1. Based on

18 the strain sweep results, a strain amplitude equal to 0.1% was selected to be in the linear

19 viscoelastic region (LVR).

20 2.3.3. Interfacial tension (IFT):


1 IFTs were measured for polymeric nanofluids with different doses of NPs and compared

2 with SiO2-water nanofluids. The measurement sequence was defined as suggested by

3 Llanos et al. [48]. Initially, the equipment is calibrated through the measurement of a

4 pattern with a digital tensiometer (KRÜSS GmbH K20, Germany), which uses the Du

5 Noüy ring mechanism. The standard fluid is deionized water (tension = 72 ± 2 mN m−1).

6 After each measurement, the ring is carefully subjected to high temperature for five minutes

7 to remove impurities or traces of previous fluids. Oil-aqueous phase IFTs were monitored

of
8 for 24 h until they reached constant values. In this condition, the system can be considered

ro
9 to be in thermodynamic equilibrium.
-p
re
10 2.3.3. Contact angle:
lP

11 Contact angles were determined to quantify the impact of nanofluids on rock wettability

12 and capillary forces. Outcrop rock samples with a diameter of 3.8 cm and a length of 2 cm
na

13 were cleaned with toluene and methanol to remove impurities. After cleaning, the samples
ur

14 were dried at 70 °C for 48 h. The rock samples were originally water-wet and were restored
Jo

15 to an oil-wet state by aging with oil in the method reported in previous studies [63].

16 Restored samples were submerged in nanofluids for 48 h at 70 °C with constant stirring

17 (500 rpm). The excess was removed by washing and drying for two hours and dried for 24

18 hours at 70 °C. Water contact angles (CA) on the surface were measured; all measurements

19 of CA were performed in quintuplicate and processed using the LayOut 2015 software.

20 2.3.3. Displacement tests in porous media:

21 Coreflood, quarter 5-spot, and microfluidic tests were employed to estimate the impact and

22 stability of micro–macro displacement for nanofluids. From coreflood tests, intensive


1 properties were determined such as apparent viscosity, retention, inaccessible pore volume,

2 and relative permeability. A quarter five-spots pattern was used to generate radial flow

3 geometries and evaluate the volumetric sweep efficiency. Finally, it was included a

4 displacement test in a microfluidic device for measurement microscopic oil displacement.

5 2.3.4. Dynamic adsorption and inaccessible pore volume:

6 HPAM is formed in its structure by amide groups (CONH2) and carboxyl groups (COO−)

of
7 [1,2]; in consequence, the level of adsorption and filterability is controlled by the charge

ro
8 distribution in the polymer and the surface of the solid [51]. Dynamic adsorption and

9
-p
inaccessible pore volume involve the stability of the polymer during the flow through
re
10 porous media, and these were evaluated for the polymeric solution and nanofluid (Tests 3
lP

11 and 4, Table 1). The experimental procedure was developed following the standard practice

12 API 63 [60], in which the basic principle is UV–vis spectrophotometry. Two polymer slugs
na

13 are injected while effluents are collected, and the mass of the polymer in the downstream
ur

14 line is determined and the mass balance of the polymer retained is identified. The polymer
Jo

15 in the effluents reacts with sodium hypochlorite (NaClO2) to form insoluble

16 polyacrylamide. Turbidity in the sample is proportional to the polymer amount, which was

17 measured by spectrophotometry (Genesys 10S UV-VIS spectrophotometer, Thermo

18 Scientific, Waltham, MA).

19 Inaccessible pore volume is an effect of blocked pores or pores that are too small

20 compared to the polymer molecular size [64], and the depleted layer is a thin layer of

21 polymer-free liquid resulting from steric expulsion of large molecules from the pore walls

22 [65,66]. During flow, the concentration front (polymer and tracer) though the accessible
1 pore volume appears perfectly normal [64]. However, salt fronts are delayed by the transfer

2 of salt into the water located in the inaccessible pore spaces. According to the technical

3 protocol to determine inaccessible pore volume [60], potassium thiocyanate (tracer at 3%)

4 is dispersed in the polymeric solution. Any such interaction would cause changes in the

5 effluent polymer concentration coincident or delayed with the changes in the salt

6 concentration [67]. Tracer concentration in the effluent is determined by titration, and the

7 inaccessible pore volume is obtained from tracer production profiles. Although small size

of
8 changes cannot be perceived from technique, this provides reasonable results when the

ro
9 aggregate size change is significant [68].
-p
re
10 2.3.5. Apparent viscosity polymer and Residual resistance factor:
lP

11 In the pore space network, deformations can be large, rapid, and non-unidirectional. The

12 field of velocities then has a divergent and convergent [69]. Flow resistance by
na

13 deformations tested in porous media was evaluated with the apparent viscosity for the
ur

14 polymeric solution and the polymeric nanofluid (Tests 1 and 2, Table 1). The procedure
Jo

15 consists of a successive injection of polymeric solution at different flow rates until it

16 reaches a steady-state. Each flow rate corresponds to an apparent shear rate and imposes a

17 flow resistance in the porous medium. The apparent viscosity of the polymer is calculated

18 using the generalized Darcy’s equation [1], and the shear rate in the porous medium ($% & ) is

19 calculated as follows [70]:

(2)
1 where = ' ⁄( ) is superficial velocity, Q the flow rate, S the core section, ) the

2 porosity, and r is the average pore radius (taken as = 789 ⁄):; ). α is a parameter that

3 depends on the homogeneity of the porous medium: for regular-homogeneous porous

4 systems, α assumes a value of one and increases as heterogeneity increases. The flow rates

5 were set by scaling the typical shear rates at the well-face (~100 s−1) and depth conditions

6 in the formation (~10 s−1) [60]. For this study, these values were 0.067, 0.167, 0.240, 0.287,

of
7 and 0.333 cm3·min−1.

ro
8 The residual resistance factor (RRF) is defined as the additional flow opposition that

9
-p
water experiences after a rock has been exposed to the polymer [2], which is evidence of
re
10 changes in the flow lines by decreases in the cross-sectional flow [71] area or the rock–
lP

11 fluid interaction [72]. This is calculated from the comparison of the water permeability

12 before and after the polymer injection. For the RRF measurement, brine was injected before
na

13 and after polymer/nanofluid flooding (Tests 1 and 2, Table 1). For every flow rate, the
ur

14 pressure drop was determined when the flow reached stability (this was near to 5 PVs
Jo

15 injected).

16 2.3.6. Relative permeability curves:

17 Relative permeability curves relate flow capacity to the saturation of fluids in a porous

18 media [69]. Implicitly, relative permeability shows the impact of capillary forces. In this

19 study, the non-stable state method proposed by Johnson, Bossler, and Naumann [73] was

20 used (Tests 5 and 6, Table 1). For this method, the volume of produced fluids and drop

21 pressure are monitored and calculated to determine the relative permeability for each phase

22 and saturation. Brine was injected until the oil residual saturation was reached, following
1 0.5 PV of the polymeric solution/nanofluid. Finally, the brine was pumped until the oil

2 production was zero. All the sequences of injection emulated a recovery process in an oil

3 field, that is, waterflooding, polymer/nanofluid flooding, and post-polymer water.

4 2.3.7. Oil recovery – macroscopic oil displacement:

5 Volumetric sweep efficiency was determined in a quarter 5-spot setup where the radial flow

6 geometric was sensible to changes in the mobility ratio. Polymer-SiO2 and water-SiO2

of
7 nanofluids were evaluated in sequential injections with the polymeric solution and

ro
8 waterflooding, respectively (Tests 7 and 8, Table 1). The flooding process in the sandpack

9
-p
is based on the fluids’ migration process. Accordingly, about 10 PV of brine and oil was
re
10 injected to saturate the system at a rate of 0.2 cm3 min−1. In the beginning, brine was
lP

11 injected (waterflooding stage) until the oil production remained equal to zero. The flow rate

12 for all recovery stages was 0.05 cm3 min−1, which is equivalent to 1 ft day−1 in the reservoir.
na

13 Subsequently, in Test 8, polymer solution was pumped (polymer flooding). Once the oil
ur

14 production was zero, nanofluids were pumped (water-SiO2 nanofluid for Test 7, and
Jo

15 polymer-SiO2 nanofluid for Test 8). The concentration of SiO2 nanoparticles was gradually

16 increased from 200 to 2000 mg L−1. At each stage, it was expected that the water cut would

17 reach 100% again, and the production of fluids and pressure drop were constantly

18 monitored. The concentration of NPs was limited by the poor capacity of water to maintain

19 elevated concentrations in solution.

20 2.3.8. Oil recovery—microscopic oil displacement

21 At pore-scale, the main distribution forms of residual oil after waterflooding are oil in

22 “dead end” oil film, oil droplets, and oil clusters [10]. Oil in the throats of the pores can be
1 produced by the alteration of at least one type of residual oil. Microscopic oil displacement

2 and its relationship with the type of residual oil affected were studied with a micromodel

3 test. Initially, the microdevice was saturated with oil, and the waterflooding stage (brine

4 injection) was started until zero oil production was reached. The next stage was polymer

5 flooding, where the injection was developed until only polymer was produced (oil

6 production = 0). Finally, polymeric nanofluid was injected. It was expected that any

7 additional oil production during this stage corresponds to the action of recovery

of
8 mechanisms that are not present in polymer flooding. All displacement stages were

ro
9 developed at 1ft d−1. (0.19 μL min−1) and room temperature (25 °C).
-p
re
10 2.3.9. Image analysis
lP

11 To calculate the oil recovery percentage of the flooding experiments, standard image

12 analysis using Fiji -Image J software [74] was used. Due to the high contrast between the
na

13 oil and the injection fluids in the obtained images, it was possible to differentiate the two
ur

14 phases by defining a proper threshold range and converting them to binary images [75].
Jo

15 Thus, the difference between the initial state of the black pixels and the final state was

16 interpreted as recovered oil. Also, the breakthrough time (tb) was calculated from the

17 images. The tb is defined as the earliest time at which the injecting fluid reaches the outlet

18 of the chip in the microfluidic system [75].

19

20 3. Results and Discussion

21 3.1. Rheology measurements:


1 Data obtained in previous studies [20,21,23–25,30], using rheology tests indicated that the

2 incorporation of SiO2 nanoparticles into polymeric solution increased the apparent viscosity

3 and the viscoelastic nature. According to Hu et al. [26], silanol groups of NPs form

4 hydrogen bonds with carboxylic groups on polymers, strengthening the polymer

5 microstructure. Viscosity and viscoelastic modules were obtained for the polymer at a fixed

6 concentration, and different doses of NPs were dispersed. Fig. 5 shows the results obtained

7 using rotational (Fig. 5a and 5b), and oscillatory (Fig. 5c) rheological measurement.

of
ro
8 Fig. 5(a) indicates an increase in viscosity according to NP concentration. Comparing

9 -p
behavior for the polymeric solution and nanofluid at 5000 mg L−1, a significant
re
10 improvement was obtained as well as a preservation in the shear-thinning behavior. Table 2

11 contains the consistency coefficient and power-law index. The power-law model provides
lP

12 reasonable predictions that are parallel to the experimental data for the shear-thinning
na

13 region. The viscosity, at 10 s−1 for the solutions, is reported in Fig. 5(b), which strongly
ur

14 confirms the model predictions. Results demonstrate that the presence of the NPs within the
Jo

15 HPAM network induces an increase in the system viscosity and an increase in shear-

16 thinning behavior. This observation is consistent with the results obtained by Aliabadian et

17 al. [20], which suggests that NPs enhance the polymer microstructure response to the flow.

18 For NP dispersion in water or SiO2-water nanofluids, viscosity is reported at a fixed

19 shear rate in Fig. 5(b), and the increase in viscosity is marginal compared with polymeric

20 nanofluids. According to the Einstein model [76], the increment of viscosity from the

21 particles of dispersion should be less than 0.5% when the NP concentration is 5000mg L−1.

22 Note, however, that experimental results at an elevated concentration in water dispersion


1 display considerable uncertainty due to the high precipitation tendency. For concentrations

2 used in this work, changes in the water viscosity can be considered negligible.

of
ro
-p
re
lP
na
ur

3
Jo

4 Figure 5. Rheological measurements for polymeric solution/nanofluid at 52 °C and the

5 impact of silica nanoparticle. (a)Viscosity plotted as a function shear rate and adjusted to

6 the power-law model. (b) Increase in viscosity measured at fixed shear rate of 10 s−1. (c)

7 Storage modulus (G′) as a function of angular frequency under a controlled strain of $< =

8 0.1%.

9
1 The rigidity of the HPAM networks was examined with the elastic modulus G′ and the

2 capability to dissipate stress was quantified by the viscous modulus Gʺ, which implies that

3 the structure of the polymer chain dramatically affects the linear viscoelastic properties

4 [29,77]. The viscoelastic modulus over a range of 0.1 to 10 rad s−1 was evaluated to identify

5 changes in the HPAM network with the incorporation of SiO2 NPs. Only G′ has been

6 reported for clarity. Fig. 5(c) displays the results of the frequency sweep test for HPAM

7 solutions, where significant differences in their microstructures is seen. The elastic modulus

of
8 G′ for HPAM solutions increases with the incorporation of SiO2 nanoparticles. Similar

ro
9 results were reported by Aliabadian et al. [20,29], who observed resistance increases in the

10
-p
microstructure of the polymer. The plateau region is clear and pronounced in higher
re
11 molecular weight polymers, concentrated solutions, the presence of long-chain branches, or
lP

12 in hydrogen bonding [78,79]. In line with the previous discussion by Hu et al. [26], an
na

13 increase in viscoelastic behavior may have its origin in the positive interaction between Si-

14 OH (on the nanoparticle) and COO− (on the polymer chain) groups. On the other hand,
ur

15 Wang et al. [10] demonstrated that viscoelastic behavior in polymer solutions increases the
Jo

16 microscopic oil displacement efficiency, which implies that this is a mechanism that must

17 be considered and is not involved in the capillary number expression.

18 Table 2. Power-law parameters for HPAM solutions: behavior with dispersed

19 SiO2 nanoparticles.

Power-law model

SiO2 Consistency Power law


RSME (%)
concentration index index
0 0.190 0.51 1.5%
500 0.195 0.51 2.1%
1000 0.210 0.52 3.2%
2000 0.260 0.56 1.9%
3000 0.310 0.59 2.3%
4000 0.400 0.65 3.2%
5000 0.540 0.71 3.1%

2 3.2. Interfacial tension (IFT):

3 Fig. 6(a) displays the IFT for crude oil with the aqueous phase (i.e., water, polymer, SiO2-

of
4 water nanofluids, and SiO2-polymer nanofluids). The IFT data in Fig. 6(a) indicates a

ro
5 strong decline in SiO2-water nanofluid relative to the concentration of nanoparticles. The

6 -p
drop in the oil–water interfacial tension due to the dispersion of SiO2 nanoparticles in water
re
7 has been widely reported [19,42]. This behavior is attributed to the positioning of NPs on
lP

8 the oil–water interface, reducing Gibbs energy. However, in the case of the polymer,

9 changes in the oil-polymer were substantially fewer. These results are consistent with those
na

10 obtained by Sharma et al. [32], who suggested that appreciable changes in oil–polymer IFT
ur

11 are only obtained with NPs for polymer-surfactant (SP) systems. It is reasonable to suppose
Jo

12 that polymer-NP networks could limit nanoparticle positioning at the interface, thus IFT

13 changes are usually minimal.


Oil-Water IFT
Oil-P olymer IFT (a) Surface treated with polymer (b)
Surface treated with water
30 100

25 80

Contact angle for water (°)


20
IFT (mN/m)

60

15
40
10

20
5

0 0

of
0 200 500 1000 1500 2000 3000 No 0 200 500 1000 1500 2000 3000
treatment
NPs concentration in the aqueous phase (mg∙L−1) NPs concentration in the treating phase (mg·L−1)

ro
1

2 Figure 6. (a) IFTs measured for oil–aqueous phase (water or polymer) and the
-p
re
3 effect of increasing the nanoparticle concentration at 25 °C. (b) Contact angles for
lP

4 water, in rocks exposed to water and polymer, and the effect of increasing
na

5 nanoparticle concentration at 25 °C.


ur

6 3.3. Contact angle:


Jo

7 Wettability is a key parameter that controls the distribution of fluids in porous media and

8 influences the capillary pressure and relative permeability curves that govern oil recovery

9 [69]. Water CAs were measured in sandstone rock cores to identify the effect SiO2 NPs on

10 the wettability of surfaces. Fig. 6(b) displays water CAs for samples that have been exposed

11 to both polymers and deionized water at different dosages of SiO2 NPs. Initially, the angle for

12 no-treatment rock was close to 90°, which indicates a water-wet condition attributed to the

13 wettability restoration process. It was observed that the angle contact decreased after the

14 rocks were exposed to water and the polymer, implying that contact with the aqueous phase
1 manages to reverse the surface wettability slightly. The NP concentration was accompanied

2 with a drop in the water contact angle. However, surfaces treated with polymer solution

3 showed the highest rate of decrease. For both fluids, an asymptotic value was observed close

4 to 22°.

5 The results for SiO2-water nanofluids are parallel to experimental data from previous

6 research [19,43,80,81], where experimental evidence has been accompanied by theoretical

7 support. For instance, Wasan et al. [80], showed that the movement of particles by

of
8 Brownian motion originates an ordering of the NPs that forms a two-dimensional layered

ro
9 structure in the confines of a three-phase contact region (solid-oil-aqueous). The ordered

10
-p
microstructures exert excess pressure in the film called structural disjoining pressure. As
re
11 this structural disjoining pressure increases, the film tension toward the vertex of the wedge
lP

12 increases, and an increase in the spreading coefficient of the nanofluids occurs, thus
na

13 causing the wetting of the surface [27,80].

14 Few researchers have addressed the wettability alteration by polymeric nanofluids in


ur

15 sandstone; however, results obtained here are in good agreement with existing polymeric
Jo

16 nanofluid flooding results [82–86]. Experimental observations indicate an alteration in the

17 wettability of the rock surface to water-wet conditions by the interaction with polymeric

18 nanofluids. The contact angle reduction has been attributed to the increased hydrophilicity

19 on the surface of the SiO2 NPs caused by the presence of the polyacrylamide solution [27].

20 Maurya et al. [84] suggested that structural disjoining pressure determines wettability

21 alteration in polymeric nanofluids, though little evidence is available about the components

22 (Van der Waals, electrostatic, or solvation forces) that play a determining role in the

23 disjoining pressure [27,82].


1 3.4. Polymer retention in porous media and inaccessible pore volume

2 Fig. 7 displays polymer and tracer production profiles for coreflooding tests (Tests 3 and 4)

3 based on the methodology proposed by Zaitoun et al. [87]. The area between the curves (1°

4 and 2° slug) is proportional to the polymer mass retained in the rock surface. The figure

5 shows a significant reduction in area, and the retention calculated was 59.8 μg∙g−1 and

6 34.9 μg∙g−1 for the polymeric solution and polymeric nanofluid, respectively. Advances in

of
7 polymeric nanofluids briefly address the impact on retention, though experimental evidence

ro
8 has shown a tendency toward reduction in the adsorption of polymeric nanofluid

9 -p
suspensions on sandstone [38–41], which is in line with the results obtained in this work.
re
10 The conceptual models propose that polymer chains on the nanoparticle surface provide

11 electrostatic stabilization by repulsion, and consequently the low adsorption on the rock
lP

12 surface is seen in the steric repulsion between nanoclusters and the surface [35–37]. Note
na

13 that the oil phase was not considered—the oil film could decrease the impact, particularly
ur

14 in strongly oil-wet systems [88].


Jo

15 The fraction of inaccessible pore volume is determined from the second polymer slug

16 and the tracer. The tracer transits in pores where polymer chains cannot enter, causing a

17 delay in the production profile [87]. The inaccessible pore volume calculated for both

18 systems was equal (20 ± 1% ). Therefore, differences in the filterability of polymer in the

19 porous medium were not observed. Although small-sized changes cannot be perceived from

20 this technique, reasonable results are obtained in short systems where the aggregate change

21 in size is significant [68]. It can be inferred, therefore, that the NPs presence in the

22 microstructure does not generate an appreciable change in the hydrodynamic radius of the
1 polymer. These results provide compelling evidence that, for the evaluated conditions, there

2 is no loss injectivity of the porous medium.

Polymeric nanofluid - 1º slug


Polymeric solution - 1º slug (a) Tracer - 1º slug
Tr acer - 1º slug
Polymeric nanofluid - 2º slug
(b)
Polymeric solution - 2º slug

1.2 1.2

1 1

Concentration ratio (C/Co)


Concentration ratio (C/Co)

0.8 0.8

0.6 0.6

of
0.4 0.4

ro
0.2 0.2

3
0
0 1 2
Injected pore volume (PV)
3
-p
4
0
0 1 2
Injected pore volume (PV)
3 4
re
4 Figure 7. Polymer retention in the porous medium determined from the production profiles
lP

5 at reservoir conditions (effective pressure = 1500 psi and temperature= 52ºC). (a)Polymeric
na

6 solution (HPAM at 500 mg∙L-1). (b)Polymeric nanofluid (HPAM at 500 mg∙L-1 with SiO2

7 nanoparticles at 3000 mg∙L-1).


ur
Jo

9 3.5. Apparent viscosity polymer and residual resistance factor:

10 The apparent viscosity in a porous medium for the polymeric solution and polymeric

11 nanofluid is reported in Fig. 8 (coreflood Tests 1 and 2). Similar behavior was observed for

12 the polymeric nanofluid in porous media and rheometer. The power-law parameters in

13 porous media were 0.66 and 0.37 (for k and n), which indicate a noticeable consistency

14 with results obtained by the rheometer (Table 2). On the contrary, the polymeric solution
1 showed a significant drop in viscosity response in porous media. The power-law parameters

2 (k = 0.07and n = 0.29) indicate a loss in the viscosity and thinning behavior.

3 In porous media, flow resistance is determined by deformations due to the contraction–

4 expansion nature of the flow in a porous medium, as well as the number of free polymer

5 molecules, which are sheared near the channel walls and stretched along the flow axis [20].

6 The evidence suggests that a drop in the viscosity of a polymer solution is related to the

of
7 retention behavior (Fig. 7). In contrast, the stability in the polymer network from the SiO2-

ro
8 NPs was crucial in reducing retention, which was derived from the conservation of

9 viscosity and mobility control in the porous medium. -p


re
Polymer ic nanofluid (measur ed in r heometer )
Polymeric solution (measured in rheometer)
Polymeric solution (measured in porous medium) (a) Polymer ic nanofluid (measur ed in porous me dium) (b)
lP

Powe r-law model


Powe r-law model
0.2
0.2
na
Apparent viscosity (Pa·s)

Apparent viscosity (Pa·s)


ur

0.04
0.04
Jo

0.008 0.008
1 5 25 125 1 5 25 125
Shear rate (s−1) Shear rate (s−1)
10

11 Figure 8. Apparent viscosity of the polymer measured in a porous medium and by

12 rheometer at 52 °C: (a) Polymeric solution (HPAM at 500 mg L−1) (b) Polymeric nanofluid

13 (HPAM at 500 mg L−1 with SiO2 nanoparticles at 3000 mg L−1).

14
1 As mentioned above, the RRF compares the water permeability before and after the

2 polymer contacts the rock surface [2], which is evidence of changes in the flow lines by

3 decreases in the cross-sectional flow [71] area or the rock–fluid interaction [72]. RRFs data

4 in Fig. 9 indicates an average value of 2.5 for the polymeric solution, which does not

5 change with an increase in the shear rate experimented in porous media. On the contrary,

6 values for the polymeric nanofluid were not constant. At low shear rate, the RRF was 5.5,

7 and with a high shear rate, RRF was 3. Additional flow restrictions for water after the

of
8 polymer flooding is evidence that the condition in the pores has changed. The previous

ro
9 section demonstrated that dynamic retention is less for polymeric nanofluids. According to

10
-p
recent research [35–37], polymer chains on the NP surface provide electrostatic repulsion
re
11 and a weak interaction with anionic sandstone surfaces, with low retention in the porous
lP

12 medium as a result [35–41]. Therefore, the flow restriction is generated by the rock–fluid
na

13 interaction. Based on the contact angle behavior, it can be deduced that behavior in the RRF

14 is due to an increase in the water-wet condition. To the knowledge of the authors, the data
ur

15 in Figs. 8–9 is the first of its kind for polymeric nanofluids.


Jo
of
ro
1

2
-p
Figure 9. The residual resistance factor as a function of shear rate tested for
re
3 fluid through porous media. Curves for the polymeric solution and polymeric
lP

4 nanofluid are measured at the reservoir conditions (effective pressure = 1500 psi and
na

5 temperature = 52ºC).
ur

6
Jo

7 3.6. Relative permeability curves:

8 Fig. 10 contains the relative permeability curves for the sequential injection of

9 waterflooding, polymer flooding (polymeric solution or nanofluid), and water drive. A

10 noticeable reduction in the permeability relative to water (Krw) is observed from Fig. 10 in

11 the polymer flooding stage, which is accompanied by a diminution in the residual oil

12 saturation (Sor). At final conditions, Krw and Sor for the coreflood test with a polymeric

13 solution were 0.1% and 17%, respectively, while for the polymeric nanofluid,

14 Krw = 0.074%, and Sor = 8% (Table 3). Comparing Krw and Sor values for flooding with and
1 without SiO2 nanoparticles, it is clear that the significant reduction in the water flow

2 capacity was accompanied by additional oil recovery. The diminution in the endpoint for

3 permeability relative to water and residual oil saturation for permeability relative curves is

4 shown by the decrease in the capillary forces [69], that is, changes in the rock–fluid

5 interaction (wettability condition), and the fluid–fluid interaction (IFT) in the pores

6 structure (size pore distribution). According to the CA and IFT results (Fig. 6), changes in

7 the endpoint for Krw and Sor are due to a wettability shift from intermediate-wet to strongly

of
8 water-wet. Similar results were reported by Sharma et al. [32], who observed that NP

ro
9 (polymer with NPs) and NSP (polymer-surfactant with NPs) nanofluids generate a strongly

10
-p
water-wet condition in quartz sandpacks, which was evidenced by the decrease in the
re
11 endpoint for Krw and shifting crossover points to the right.
lP

Kro_in Waterflooding (PS) Kro_i n Waterflooding (PN) Krw_in Waterflooding (PS)


Krw_in Polymer flooding (PS) Krw_in Polymer flooding (PN) Krw_in Water drive (PS)
na

Krw_in Water drive (PN) Krw_in Waterflooding (PN)


1.00 1.00
Kro
0.90 Polymer 0.90
ur

Waterflooding Water drive


flooding

0.80 0.80
Jo

0.70 0.70

0.60 0.60
Kro

K rw

0.50 0.50

0.40 0.40

0.30 0.30
0.1457
Krw 0.1000
0.20 0.20
0.0740

0.10 0.10

0.00 0.00
0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0
Sw
PS = Polymeric solution test
PN = Polymeric nanofluid test
12
1 Figure 10. Relative permeability curves during the stages of waterflooding,

2 polymer flooding (polymeric solution or polymeric nanofluid), and water drive;

3 measured at reservoir conditions (effective pressure = 1500 psi and temperature =

4 52ºC).

5 Table 3. Endpoints and residual saturations for the polymeric solution and

6 polymeric nanofluid tests (fraction).

of
Polymeric solution Polymeric nanofluid

ro
Polymer Polymer
Waterflooding Water drive Waterflooding Water drive
flooding flooding
Kro
Krw
1
0.145
-
0.0563
-
0.1
-p 1
0.145
-
0.05
-
0.074
re
Sor 0.28 0.22 0.17 0.28 0.2 0.08
Swr 0.23 0.23 0.23 0.23 0.23 0.23
lP

7
na

8 3.7. Capillary number and mobility ratio:


ur
Jo

9 It was suggested in the previous section that SiO2 nanoparticles alter the capillary forces

10 and rheology behavior of HPAM polymer solutions. Overall, water-wet conditions, viscous

11 and viscoelastic behavior follow similar patterns, increasing as a function of nanoparticle

12 concentration. The capillary number and mobility ratio were calculated to characterize the

13 displacement forces present in the porous medium (Table 4). The capillary number

14 provides information about the microscopic oil displacement, while the mobility ratio

15 estimates the macroscopic displacement. As shown in Table 4, in SiO2-water nanofluids,

16 the capillary number was considerably affected by nanoparticles. The growth in the

17 capillary number was caused by the increase in the water-wet condition and a reduction in
1 the IFT. Meanwhile, an unfavorable mobility ratio was observed for all cases evaluated

2 with SiO2-water. For polymeric nanofluids, Nc increased with the concentration of

3 nanoparticles; however, their behavior is more conservative, which is a consequence of a

4 wettability shift to water-wet. On the contrary, a favorable mobility ratio was intensified in

5 every NP concentration, which is attributed to the viscosity polymer increasing.

6 In porous media, it was observed that a reduction in the viscosity loss by retention was

of
7 the determinant for macroscopic displacement. This reduction in the viscosity loss results in

ro
8 the apparent viscosity of the polymeric nanofluid to be conserved during the transit in

9 -p
pores. The results seem to indicate that the conservation of the favorable mobility ratio is
re
10 the determinant mechanism for polymeric nanofluid flooding for macroscopic oil

11 displacement.
lP

12
na

13 Table 4. The capillary number and mobility ratio for SiO2-water and SiO2-
ur

14 polymer nanofluids as a function of nanoparticle concentration.


Jo

SiO2-water nanofluid SiO2-polymer nanofluid


NPs Capillary number Mobility ratio Capillary number Mobility ratio
concentration (Nc) (M) (Nc) (M)
0 2.4 x 10-7 80 1.2 x 10-5 1.08
-7 -5
200 5.7 x 10 79.9 1.3 x 10 0.98
500 8.6 x 10-7 79.9 1.4 x 10-5 0.95
1000 1.2 x 10-6 79.9 1.9 x 10-5 0.94
1500 1.4 x 10-6 79.9 2.1 x 10-5 0.92
-6
2000 1.5 x 10 79.8 3.0 x 10-5 0.92
-6 -5
3000 1.5 x 10 79.6 3.6 x 10 0.79

15
1 The wettability shift from intermediate-wet to water-wet obtained from CAs was

2 confirmed with the reduction in flow capacity of the aqueous phase by the incrusting of

3 NPs in the polymer. Guo et al. [59] presented a wide discussion about the capillary number

4 and impact in the recovery oil, suggesting that small changes (near to one decade) in the Nc

5 could mobilize residual oil. Moreover, Manrique et al. [89] proposed that significant oil

6 recovery is achieved at low and ultra-low IFTs; however, capillary forces can act even

7 under low Nc changes over trapped oil in pores. It can be inferred that part of the oil

of
8 produced by polymeric nanofluid is a consequence of a reduction in the capillary forces.

ro
9 -p
Another mechanism that could be of considerable importance in microscopic oil
re
10 displacement and is not considered in the capillary number is the viscoelasticity of the

11 polymer. A frequency sweep showed a relationship between the content of NPs and the
lP

12 viscoelastic nature of the polymeric nanofluids. According to Want et al. [10], at the micro-
na

13 scale, residual oil is not pushed out by the polymer solution but pulled out. The larger the
ur

14 viscoelastic behavior, the stronger the capability of the polymer solution to “sweep out” the
Jo

15 residual oil [9,10].

16 3.8. Oil recovery—macroscopic oil displacement

17 Fig. 11 displays oil recovery for water (Test 7) and polymer flooding (Test 8) during

18 sequential injection at an incremental concentration of SiO2 NP dispersions (from 0 to

19 2000mg L−1). Displacement tests were realized in a quarter 5-spot system to induce

20 macroscopic oil displacement by volumetric sweeping and radial flow. The sweeping

21 efficiency of the polymer by the favorable mobility ratio was clearly observed during the

22 first two pore volumes (without nanoparticles, Fig. 11(a)). Oil recovered was 70% and 43%
1 for polymer flooding and waterflooding, respectively. According to theory, the additional

2 oil recovered corresponds to the homogeneous flow front of the polymer, which sweeps

3 zones where the water did not flow [5]. Sequential injections of SiO2-water and SiO2-

4 polymer nanofluids at incremental doses of NPs revealed small additional oil produced,

5 which is not clearly observed by the scale of the graphic.

6 Fig. 11(b) displays the diminution of residual oil saturation normalized (( A ) as a

of
7 function of NPs concentration, which can be calculated as ( ABC − ( A ⁄( ABC , where

ro
8 ( ABC and ( A are the residual oil saturation at the waterflooding stage and at the

9 -p
evaluating condition, respectively. Incremental oil recovered was observed for two
re
10 displacement tests; however, for polymeric nanofluids, the reduction in the residual oil

11 saturation was more substantial. In accordance with previous results, additional oil
lP

12 recovered can be attributed to the conservation of a favorable mobility ratio in porous


na

13 media and the microscopic oil sweep. Although they cannot be directly differentiated, it can
ur

14 be said that from each mechanism, capillary forces, and viscoelastic behavior, part of the
Jo

15 oil recovery is provided from microscopic oil displacement. In contrast, additional oil

16 obtained in SiO2-water nanofluid flooding was only provided by the improvement of

17 microscopic oil mechanisms (IFT reduction and water-wet condition).


SiO
(a) SiO2-polymer
SiO2-Polymernanofluid
nanofluid (b)
2-Polymer nanofluid
SiO2-polymer nanofluid SiO2-Water nanofluid
SiO2-water nanofluid
Pressure_Polymer Pressure _Water SiO2-water
SiO2-Waternanofluid
nanofluid
1.0 0.06 50%

NPs=0 mg·L-1 NPs=200 mg·L-1 NPs=500 mg·L-1 NPs=1000 mg·L-1 NPs=1500 mg·L-1 NPs=2000 mg·L-1

0.05
0.8 40%

Sorn = (Sor WF − Sori)/Sor WF


0.04

Pressure drop (MPa)


0.6 30%
Oil recovery

0.03

0.4 20%
0.02

0.2 10%
0.01

0.0 0 0%
0 2 4 6 8 10 12 0 500 1000 1500 2000
Injected pore volume (PV) NPs concentration (mg·L−1)

of
1

ro
2 Figure 11. Oil recovery in quarter 5-Spot system for SiO2-water and SiO2-

3 polymer nanofluids during sequential injection at incremental concentrations of SiO2 -p


re
4 nanoparticles; temperature = 52 °C. (a) Oil recovery as a function of the injected
lP

5 pore volume. (b) Diminution of residual oil saturation normalized as a function of

6 NPs concentration.
na

7
ur
Jo

8 3.9. Oil recovery - microscopic oil displacement:

9 Fig. 12 shows the microscopic sweep efficiency for sequential injections of polymer and

10 nanofluid. Initially, the microdevice was saturated with injected oil and brine. Types of

11 residual oil observed after waterflooding were oil droplets, oil film coats, dead-end, and

12 cluster types according to models of residual oil distribution [10]. Oil lodges at the rock

13 crevices and the “dead ends” of flow channels. In this study, however, case oil film and oil

14 droplets trapped between grains were more frequent. These residual oil types are observed

15 from capillary forces commonly seen in strongly oil-wet rocks [6,10]. In contrast, oil

16 clusters are trapped in microscopic pores when the porous media has small-scale
1 heterogeneity. Shapes in the residual oil indicate a surface with a tendency to oil-wet,

2 perhaps by the nature of the material (epoxy resin). For this reason, efficiency in oil

3 recovery was low in the waterflooding stage (46.7%). The recovery obtained by the

4 flooding of the polymeric solution and polymeric nanofluid were 52.2% and 56.4%,

5 respectively. Residual oil distribution after the polymeric solution and polymeric nanofluid

6 flooding revealed that there is a widening of the flow channel formed by the water and a

7 reduction in the size of the oil cluster.

of
ro
Waterdrive
Polymer flooding
Polymeric nanofluid flooding
1. 0

0. 9

0. 8

0. 7
-p
re
0.564
0. 6 0.522
Oil recovery

0.467
0. 5
lP

0. 4

0. 3

0. 2

0. 1
na

0. 0
0 1 2 3 4 5 6 7
Injected pore volume (PV)
ur

2
1
3
Jo

Oil saturation Waterflooding Polymer flooding Polymeric nanofluid flooding


Types of residual oil

1. Oil droplet 3. Cluster -type

2. Dead end 4. Oil film


8

9 Figure 12. Microscopic oil displacement for sequential flooding of the polymeric

10 solution and polymeric nanofluid

11 Micro forces are caused by the change of kinetic energy that simultaneously occurs

12 with the velocity changes of the flooding fluid. For instance, Xia et al. [7] proposed that
1 these forces act on the protruding portion of the oil blob, making it change shape and move.

2 Structural disjoining pressure could reduce capillary forces in polymeric nanofluids, which

3 contributes to the disruption of oil clusters. However, several authors have proposed that

4 microscopic oil displacement in HPAM injection is due to its viscous-elastic effects

5 [6,7,9,10]. The driving force can be increased due to the increase in micro force; therefore,

6 micro displacement efficiency is enhanced and decreased for all types of residual oil. [7].

7 Consequently, the micro displacement of oil in polymeric nanofluid flooding can be

of
8 attributed to an increase in the viscoelasticity and decreasing capillary forces occasioned by

ro
9 the SiO2 NPs.
-p
re
10
lP

11 4. Conclusions
na

12 In this work, displacement mechanisms were described that act in oil recovery during

13 polymeric nanofluid flooding. Our results provide evidence that SiO2 nanoparticles
ur

14 improve the macro-microscopic mechanisms of oil displacement in a polymer solution.


Jo

15 At the macro condition, the conservation of a favorable mobility ratio along the porous

16 medium was obtained that was determinant in oil displacement by volumetric sweeping.

17 Although there was a small increase in the viscosity at low concentrations of nanoparticles,

18 the control of the loss of mass by retention resulted in the conservation of the apparent

19 viscosity of the polymeric nanofluids in porous media as well as in mobility control. On the

20 contrary, a direct relationship was seen between the concentration of nanoparticles and the

21 viscoelastic behavior of the polymer and water-wet condition on the sandstone surface

22 exposed to the nanofluid. In consequence, microscopic oil displacement was mainly caused
1 by a reduction in the size of oil clusters in the residual oil distribution, which can be

2 attributed to a reduction in the capillary forces and an increase in the viscoelasticity nature

3 of the polymer.

4 Future work should focus on the discretization of the percentage contribution of all the

5 mechanisms demonstrated in this research. Combined with the results in this study, this

6 data could eventually lead to the identification of appropriate implementation methods in

of
7 the reservoir to maximize the capabilities of polymer nanofluids.

ro
8 Acknowledgments

9 The authors would like to


-p
acknowledge AGENCIA NACIONAL DE
re
10 HIDROCARBUROS, COLCIENCIAS, and Universidad Nacional de Colombia for
lP

11 financial and logistical support in this research, provided in Agreement 721 of 2015. The
na

12 authors also thank M.Sc. Lady Giraldo and M.Sc. Sebastián Llanos for their fruitful

13 support.
ur
Jo

14 Nomenclature:

15 Nc = Capillary number

16 = Viscosity of the advancing non-wetting fluid

17 = Velocity of the advancing non-wetting fluid

18 = Fluid-fluid contact angle

19 M = Mobility ratio

20 " = Advancing non-wetting fluid mobility

21 " = Displaced wetting fluid mobility

22 PGN = Polymer-grafted nanoparticle


1 SiO2 -NPs = hydrophilic fumed silica nanoparticle

2 NPs = Nanoparticles

3 D = Apparent viscosity

4 k = Consistency index

5 n = Power-law index

6 RSME % = Root-mean-square error

7 LVR = Linear viscoelastic region

of
8 IFT = Interfacial tension

ro
9 CA = Water contact angle

10 $% E = Shear rate in porous medium -p


re
11 = Superficial velocity

12 Q = Flow rate
lP

13 S = Core section
na

14 α = Parameter depending of the homogeneity in the rock

15 r = Average pore radius


ur

16 RRF = Residual resistance factor


Jo

17 Sor = Residual oil saturation

18 WF= Waterflooding

19 SiO2-water nanofluids = Nanofluids formed by the dispersion of SiO2 nanoparicles in


20 deionized water.

21 SiO2-polymer nanofluids = Nanofluids formed by the dispersion of SiO2 nanoparicles


22 in polymeric solution (also named as polymeric nanofluids).

23

24 References:

25 1. Sheng, J.J. Polymer Flooding. In Modern Chemical Enhance Oil Recovery; 2011;
26 pp. 101–206 ISBN 9781856177450.
1 2. Sorbie, K.S. Polymer-improved oil recovery; Springer Science+Business Media New
2 York: India, 1991; ISBN 9789401053549.

3 3. Wever, D.A.Z.Z.; Picchioni, F.; Broekhuis, A.A. Polymers for enhanced oil
4 recovery: A paradigm for structure-property relationship in aqueous solution. Prog.
5 Polym. Sci. 2011, 36, 1558–1628.

6 4. Zhang, C.; Oostrom, M.; Wietsma, T.W.; Grate, J.W.; Warner, M.G. Influence of
7 viscous and capillary forces on immiscible fluid displacement: Pore-scale
8 experimental study in a water-wet micromodel demonstrating viscous and capillary
9 fingering. Energy and Fuels 2011, 25, 3493–3505.

of
10 5. Sheng, J.J. Mobility Control Requirement in EOR Processes. Mod. Chem. Enhanc.

ro
11 Oil Recover. 2011, 79–100.

12
13
6. -p
Yin, H.; Wang, D.; Zong, H. Study on flow behaviors of viscoelastic polymer
solution in micropore with dead end. SPE Annu. Technol. Conf. 2006, 1–10.
re
14 7. Xia, H.; Wang, D.; Wang, G.; Ma, W.; Deng, H.W.; Liu, J. Mechanism of the Effect
lP

15 of Micro-Forces on Residual Oil in Chemical Flooding. SPE Symp. Improv. Oil


16 Recover. 20-23 April. Tulsa, Oklahoma, USA 2008, 1–10.
na

17 8. Friedrich, K.; Zhang, M.Q. Nonlinear phenomena in flows of viscoelastic polymer


18 fluids; 1995; Vol. 6; ISBN 9789401112581.
ur

19 9. Sheng, J.J. Polymer Viscoelastic Behavior and Its Effect on Field Facilities and
Jo

20 Operations. In Modern-Chemical-Enhanced-Oil-Recovery; 2011; pp. 207–238 ISBN


21 978-1-85617-745-0.

22 10. Wang, D.; Xia, H.; Liu, Z.; Yang, Q. Study of the Mechanism of Polymer Solution
23 With Visco-Elastic Behavior Increasing Microscopic Oil Displacement Efficiency
24 and the Forming of Steady “Oil Thread” Flow Channels. SPE Asia Pacific Oil Gas
25 Conf. Exhib. 2001, 1–9.

26 11. Venkataraman, P.; Rao, P.R. Non-Darcy Flow Through Large Granular Media. ISH
27 J. Hydraul. Eng. 1998, 4, 39–48.

28 12. Larson, R.G. DERIVATION OF GENERALIZED DARCY EQUATIONS FOR


29 CREEPING FLOW IN POROUS MEDIA. Ind. Eng. Chem. Fundam. 1981, 20, 132–
30 137.

31 13. Minale, M. Modelling the flow of a second order fluid through and over a porous
1 medium using the volume averages. II. The stress boundary condition. Phys. Fluids
2 2016, 28, 023103.

3 14. Minale, M. Modelling the flow of a second order fluid through and over a porous
4 medium using the volume averages. I. The generalized Brinkman’s equation. Phys.
5 Fluids 2016, 28, 023102.

6 15. Tosco, T.; Marchisio, D.L.; Lince, F.; Sethi, R. Extension of the Darcy-Forchheimer
7 Law for Shear-Thinning Fluids and Validation via Pore-Scale Flow Simulations.
8 Transp. Porous Media 2013, 96, 1–20.

9 16. Pearson, J.R.A.; Tardy, P.M.J. Models for flow of non-Newtonian and complex

of
10 fluids through porous media. J. Nonnewton. Fluid Mech. 2002, 102, 447–473.

ro
11 17. Agista, M.N.; Guo, K. A State-of-the-Art Review of Nanoparticles Application in
12 -p
Petroleum with a Focus on Enhanced Oil Recovery. Appl. Sci. 2018.

13 18. Chegenizadeh, N.; Saeedi, A.; Quan, X. Application of Nanotechnology for


re
14 Enhanced Oil Recovery: A Review. Defect Diffus. Forum 2016, 367, 149–156.
lP

15 19. Franco, C.A.; Zabala, R.; Cortés, F.B. Nanotechnology applied to the enhancement
16 of oil and gas productivity and recovery of Colombian fields. J. Pet. Sci. Eng. 2017,
na

17 157, 39–55.

18 20. Aliabadian, E.; Sadeghi, S.; Kamkar, M.; Chen, Z.; Sundararaj, U. Rheology of
ur

19 fumed silica nanoparticles/partially hydrolyzed polyacrylamide aqueous solutions


Jo

20 under small and large amplitude oscillatory shear deformations. J. Rheol. (N. Y. N.
21 Y). 2018, 62, 1197–1216.

22 21. Zhu, D.; Han, Y.; Zhang, J.; Li, X.; Feng, Y. Enhancing rheological properties of
23 hydrophobically associative polyacrylamide aqueous solutions by hybriding with
24 silica nanoparticles. J. Appl. Polym. Sci. 2014, 131, n/a-n/a.

25 22. Zheng, C.; Cheng, Y.; Wei, Q.; Li, X.; Zhang, Z. Suspension of surface-modified
26 nano-SiO2in partially hydrolyzed aqueous solution of polyacrylamide for enhanced
27 oil recovery. Colloids Surfaces A Physicochem. Eng. Asp. 2017, 524, 169–177.

28 23. Zeyghami, M.; Kharrat, R.; Ghazanfari, M.H.H. Investigation of the Applicability of
29 Nano Silica Particles as a Thickening Additive for Polymer Solutions Applied in
30 EOR Processes. Energy Sources Part a-Recovery Util. Environ. Eff. 2014, 36, 1315–
31 1324.
1 24. Maurya, N.K.; Mandal, A. Studies on behavior of suspension of silica nanoparticle
2 in aqueous polyacrylamide solution for application in enhanced oil recovery. Pet.
3 Sci. Technol. 2016, 34, 429–436.

4 25. Maghzi, A.; Mohebbi, A.; Kharrat, R.; Ghazanfari, M.H. An experimental
5 investigation of silica nanoparticles effect on the rheological behavior of
6 polyacrylamide solution to enhance heavy oil recovery. Pet. Sci. Technol. 2013, 31,
7 500–508.

8 26. Hu, Z.; Haruna, M.; Gao, H.; Nourafkan, E.; Wen, D. Rheological Properties of
9 Partially Hydrolyzed Polyacrylamide Seeded by Nanoparticles. Ind. Eng. Chem. Res.

of
10 2017, 56, 3456–3463.

ro
11 27. Gbadamosi, A.O.; Junin, R.; Manan, M.A.; Yekeen, N.; Agi, A.; Oseh, J.O. Recent
12 advances and prospects in polymeric nanofluids application for enhanced oil
13 -p
recovery. J. Ind. Eng. Chem. 2018, 66, 1–19.
re
14 28. Corredor, L.M.; Husein, M.M.; Maini, B.B. A review of polymer nanohybrids for oil
15 recovery. Adv. Colloid Interface Sci. 2019, 272, 102018.
lP

16 29. Aliabadian, E.; Kamkar, M.; Chen, Z.; Sundararaj, U. Prevention of network
17 destruction of partially hydrolyzed polyacrylamide (HPAM): Effects of salt,
na

18 temperature, and fumed silica nanoparticles. Phys. Fluids 2019, 31, 013104.
ur

19 30. Rezaei, A.; Abdi-Khangah, M.; Mohebbi, A.; Tatar, A.; Mohammadi, A.H.A.H.
20 Using surface modified clay nanoparticles to improve rheological behavior of
Jo

21 Hydrolized Polyacrylamid (HPAM) solution for enhanced oil recovery with polymer
22 flooding. J. Mol. Liq. 2016, 222, 1148–1156.

23 31. Giraldo, L.J.L.J.; Giraldo, M.A.; Llanos, S.; Maya, G.; Zabala, R.D.R.D.; Nassar,
24 N.N.N.; Franco, C.A.C.A.; Alvarado, V.; Cortés, F.B.F.B. The effects of
25 SiO2nanoparticles on the thermal stability and rheological behavior of hydrolyzed
26 polyacrylamide based polymeric solutions. J. Pet. Sci. Eng. 2017, 159, 841–852.

27 32. Sharma, T.; Iglauer, S.; Sangwai, J.S.J.S. Silica Nanofluids in an Oilfield Polymer
28 Polyacrylamide: Interfacial Properties, Wettability Alteration, and Applications for
29 Chemical Enhanced Oil Recovery. Ind. Eng. Chem. Res. 2016, 55, 12387–12397.

30 33. Hu, X.; Ke, Y.; Zhao, Y.; Lu, S.; Yu, C.; Peng, F. Synthesis and characterization of a
31 β-cyclodextrin modified polyacrylamide and its rheological properties by hybriding
32 with silica nanoparticles. Colloids Surfaces A Physicochem. Eng. Asp. 2018, 548,
1 10–18.

2 34. Nguyen, B.D.B.D.; Ngo, T.K.T.K.; Bui, T.H.T.H.; Pham, D.K.D.K.; Dinh,
3 X.L.X.L.; Nguyen, P.T.P.T. The impact of graphene oxide particles on viscosity
4 stabilization for diluted polymer solutions using in enhanced oil recovery at HTHP
5 offshore reservoirs. Adv. Nat. Sci. Nanosci. Nanotechnol. 2015, 6, 15012.

6 35. He, F.; Zhao, D.; Liu, J.; Roberts, C.B. Stabilization of Fe - Pd nanoparticles with
7 sodium carboxymethyl cellulose for enhanced transport and dechlorination of
8 trichloroethylene in soil and groundwater. Ind. Eng. Chem. Res. 2007, 46, 29–34.

9 36. Bagaria, H.G.; Xue, Z.; Neilson, B.M.; Worthen, A.J.; Yoon, K.Y.; Nayak, S.;

of
10 Cheng, V.; Lee, J.H.; Bielawski, C.W.; Johnston, K.P. Iron oxide nanoparticles

ro
11 grafted with sulfonated copolymers are stable in concentrated brine at elevated
12 temperatures and weakly adsorb on silica. ACS Appl. Mater. Interfaces 2013, 5,
13 3329–3339. -p
re
14 37. Schwenke, K.; Isa, L.; Del Gado, E. Assembly of nanoparticles at liquid interfaces:
15 Crowding and ordering. Langmuir 2014, 30, 3069–3074.
lP

16 38. Cheraghian, G. Evaluation of clay and fumed silica nanoparticles on adsorption of


17 surfactant polymer during enhanced oil recovery. J. Japan Pet. Inst. 2017, 60, 85–
na

18 94.
ur

19 39. Cheraghian, G.; Shahram, S.; Nezhad, K.; Kamari, M.; Hemmati, M.; Masihi, M.;
20 Bazgir, S.; Khalili Nezhad, S.S.; Kamari, M.; Hemmati, M.; et al. Adsorption
Jo

21 polymer on reservoir rock and role of the nanoparticles, clay and SiO 2. Int. Nano
22 Lett. 2014, 4, 114.

23 40. Najafiazar, B.; Yang, J.; Simon, C.R.; Karimov, F.; Torsæter, O.; Holt, T. Transport
24 Properties of Functionalised Silica Nanoparticles in Porous Media.; Society of
25 Petroleum Engineers, 2016.

26 41. Khalilinezhad, S.S.S.S.; Cheraghian, G.; Karambeigi, M.S.M.S.M.S.; Tabatabaee,


27 H.; Roayaei, E. Characterizing the Role of Clay and Silica Nanoparticles in
28 Enhanced Heavy Oil Recovery During Polymer Flooding. Arab. J. Sci. Eng. 2016,
29 41, 2731–2750.

30 42. Sun, X.; Zhang, Y.; Chen, G.; Gai, Z. Application of nanoparticles in enhanced oil
31 recovery: A critical review of recent progress. Energies 2017, 10.

32 43. Li, S.; Hendraningrat, L.; Torsaeter, O. Improved Oil Recovery by Hydrophilic
1 Silica Nanoparticles Suspension: 2-Phase Flow Experimental Studies. Int. Pet.
2 Technol. Conf. 2013, 3, 1–15.

3 44. Hendraningrat, L.; Torsæter, O. A Stabilizer that Enhances the Oil Recovery Process
4 Using Silica-Based Nanofluids. Transp. Porous Media 2015, 108.

5 45. Guo, K.; Li, H.; Yu, Z. Metallic Nanoparticles for Enhanced Heavy Oil Recovery:
6 Promises and Challenges. In Proceedings of the Energy Procedia; 2015; Vol. 75.

7 46. Franco, C.A.; Montoya, T.; Nassar, N.N.; Pereira-Almao, P.; Cortés, F.B.
8 Adsorption and Subsequent Oxidation of Colombian Asphaltenes onto Nickel and/or
9 Palladium Oxide Supported on Fumed Silica Nanoparticles. Energy & Fuels 2013,

of
10 27, 7336–7347.

ro
11 47. Franco, C.A.; Nassar, N.N.; Montoya, T.; Ruíz, M.A.; Cortés, F.B. Influence of
12
13
-p
Asphaltene Aggregation on the Adsorption and Catalytic Behavior of Nanoparticles.
Energy & Fuels 2015, 29, 1610–1621.
re
14 48. Llanos, S.; Giraldo, L.J.; Santamaria, O.; Franco, C.A.; Cortés, F.B. Effect of
lP

15 Sodium Oleate Surfactant Concentration Grafted onto SiO<inf>2</inf>


16 Nanoparticles in Polymer Flooding Processes. ACS Omega 2018, 3.
na

17 49. Seright, R.S.; Campbell, A.R.; Mozley, P.S.; Han, P. Stability of partially
18 hydrolyzed polyacrylamides at elevated temperatures in the absence of divalent
ur

19 cations. SPE J. 2010, 15, 341–348.


Jo

20 50. Caulfield, M.J.; Hao, X.; Qiao, G.G.; Solomon, D.H. Degradation on
21 polyacrylamides. Part I. Linear polyacrylamide. Polymer (Guildf). 2003, 44, 1331–
22 1337.

23 51. Graveling, G.J.; Vala Ragnarsdottir, K.; Allen, G.C.; Eastman, J.; Brady, P. V.;
24 Balsley, S.D.; Skuse, D.R. Controls on polyacrylamide adsorption to quartz,
25 kaolinite, and feldspar. Geochim. Cosmochim. Acta 1997, 61, 3515–3523.

26 52. Cortés, F.B.; Lozano, M.; Santamaria, O.; Marquez, S.B.; Zapata, K.; Ospina, N.;
27 Franco, C.A. Development and evaluation of surfactant nanocapsules for chemical
28 Enhanced Oil Recovery (EOR) applications. Molecules 2018, 23.

29 53. Khaledi, R. Optimized Solvent for Solvent Assisted-Steam Assisted Gravity


30 Drainage (SA-SAGD) Recovery Process. Exxonmob. Upstream Res. Co. 2015, 9–11.

31 54. Sharma, A.P.; Rao, D.N. Scaled physical model experiments to characterize the gas-
1 assisted gravity drainage EOR process. SPE - DOE Improv. Oil Recover. Symp.
2 Proc. 2008, 2, 773–795.

3 55. Shen, P.; Wang, J.; Yuan, S.; Zhong, T.; Jia, X. Study of enhanced-oil-recovery
4 mechanism of alkali/surfactant/polymer flooding in porous media from experiments.
5 SPE J. 2009, 14, 237–244.

6 56. Farzaneh, S.A.; Dehghan, A.A.; Kharrat, R.; Ghazanfari, M.H. A comparative study
7 of WAS, SWAS and solvent-soak scenarios applied to heavy oil reservoirs using 5-
8 spot glass micromodels. In Proceedings of the Canadian International Petroleum
9 Conference 2009, CIPC 2009; Petroleum Society of Canada (PETSOC), 2009.

of
10 57. Kianinejad, A.; Ghazanfari, M.H.; Masihi, M.; Rashtchian, D.; Saidian, M.

ro
11 Visualization of the displacement mechanisms during worm-like surfactants flooding
12 in heavy oil fractured 5-spot models. In Proceedings of the 73rd European
13
14
-p
Association of Geoscientists and Engineers Conference and Exhibition 2011:
Unconventional Resources and the Role of Technology. Incorporating SPE
re
15 EUROPEC 2011; Society of Petroleum Engineers, 2011; Vol. 5, pp. 3300–3304.
lP

16 58. Olmos, C.M.; Vaca, A.; Rosero, G.; Peñaherrera, A.; Perez, C.; de Sá Carneiro, I.;
17 Vizuete, K.; Arroyo, C.R.; Debut, A.; Pérez, M.S.; et al. Epoxy resin mold and
18 PDMS microfluidic devices through photopolymer flexographic printing plate.
na

19 Sensors Actuators, B Chem. 2019, 288, 742–748.


ur

20 59. Guo, H.; Song, K.; Hilfer, R. A critical review of capillary number and its
21 application in enhanced oil recovery. Proc. - SPE Symp. Improv. Oil Recover. 2020,
Jo

22 2020-Augus.

23 60. API Recommended Practices for Evaluation of Polymers Used in Ehanced Oil
24 Recovery Operations. API Recom. Pract. 63 1990.

25 61. Carreau, P.J. Rheological Equations From Molecular Network Theories. Trans Soc
26 Rheol 1972, 16, 99–127.

27 62. Barnes, H.A. A HANDBOOK OF ELEMENTARY RHEOLOGY; First.; Published by


28 The University of Wales Institute of Non-Newtonian Fluid Mechanics, Department
29 of Mathematics, University of Wales Aberystwyth, Penglais, Aberystwyth, Dyfed,
30 Wales, SY23 3BZ., 2000; ISBN 0953803201.

31 63. Franco-Aguirre, M.; Zabala, R.D.; Lopera, S.H.; Franco, C.A.; Cortés, F.B.
32 Interaction of anionic surfactant-nanoparticles for gas - Wettability alteration of
1 sandstone in tight gas-condensate reservoirs. J. Nat. Gas Sci. Eng. 2018, 51, 53–64.

2 64. Dawson, R.; Lantz, R.B. Inaccessible pore volumen in polymer flooding. Soc Pet
3 Eng J 1972, 12, 448–452.

4 65. Sorbie, K.S. Depleted layer effects in polymer flow through porous media. I. Single
5 capillary calculations. J. Colloid Interface Sci. 1990, 139, 299–314.

6 66. Omari, A.; Moan, M.; Chauveteau, G. Wall effects in the flow of flexible polymer
7 solutions through small pores. Rheol. Acta 1989, 28, 520–526.

8 67. Al-Hajri, S.; Mahmood, S.M.S.M.; Abdulelah, H.; Akbari, S. An overview on

of
9 polymer retention in Porous media. Energies 2018, 11.

ro
10 68. Dawson, R.; Lantz, R.B. Inaccessible Pore Volume in Polymer Flooding. Soc. Pet.
11 Eng. J. 1972, 12, 448–452.

12 69.
-p
Tiad, D. Petrophysics Theory and Practice of Measuring Reservoir Rock and Fluid
re
13 Transport Properties; Gulf Professional Publishing, 2004;
lP

14 70. Zaitoun, A.; Makakou, P.; Blin, N.; Al-Maamari, R.; Al-Hashmi, A.-A.A.R.; Abdel-
15 Goad, M.; Al-Sharji, H.H. Shear Stability of EOR Polymers. SPE J. 2012, 17, 335–
na

16 339.

17 71. Grattoni, C.A.; Luckham, P.F.; Jing, X.D.; Norman, L.; Zimmerman, R.W. Polymers
ur

18 as relative permeability modifiers: Adsorption and the dynamic formation of thick


19 polyacrylamide layers. J. Pet. Sci. Eng. 2004, 45, 233–245.
Jo

20 72. Martin, F.D.; Sherwood, N.S. The Effect Solution of Hydrolysis of Polyacrylamide
21 Viscosity, Polymer Retention and Resistance Properties. Soc. Pet. Eng. J. 1975.

22 73. Johnson, E.F.; Bossler, D.P.; Naumann, V.O. Calculation of Relative Permeability
23 from Displacement Experiments. SPE 1023 1959.

24 74. Schindelin, J.; Arganda-Carreras, I.; Frise, E.; Kaynig, V.; Longair, M.; Pietzsch, T.;
25 Preibisch, S.; Rueden, C.; Saalfeld, S.; Schmid, B.; et al. Fiji: An open-source
26 platform for biological-image analysis. Nat. Methods 2012, 9, 676–682.

27 75. Kargozarfard, Z.; Riazi, M.; Ayatollahi, S. Viscous fingering and its effect on areal
28 sweep efficiency during waterflooding: an experimental study. Pet. Sci. 2019, 16,
29 105–116.

30 76. Einstein, A. Eine neue Bestimmung der Moleküldimensionen. Ann. Phys. 1906, 14,
1 229–247.

2 77. Colby, R.H. Structure and linear viscoelasticity of flexible polymer solutions:
3 Comparison of polyelectrolyte and neutral polymer solutions. Rheol. Acta 2010, 49,
4 425–442.

5 78. Shabbir, A.; Goldansaz, H.; Hassager, O.; Van Ruymbeke, E.; Alvarez, N.J. Effect
6 of Hydrogen Bonding on Linear and Nonlinear Rheology of Entangled Polymer
7 Melts. Macromolecules 2015, 48, 5988–5996.

8 79. Deshpande, A.; Krishnan, M.; Kumar, S. Rheology of Complex Fluids; Springer,
9 2010; Vol. 53; ISBN 978-1-4419-6493-9.

of
10 80. Lim, S.; Horiuchi, H.; Nikolov, A.D.; Wasan, D. Nanofluids alter the surface

ro
11 wettability of solids. Langmuir 2015, 31, 5827–5835.

12 81.
-p
Zhang, H.; Nikolov, A.; Wasan, D. Enhanced oil recovery (EOR) using nanoparticle
13 dispersions: Underlying mechanism and imbibition experiments. Energy and Fuels
re
14 2014, 28.
lP

15 82. Wasan, D.; Nikolov, A.; Kondiparty, K. The wetting and spreading of nanofluids on
16 solids: Role of the structural disjoining pressure. Curr. Opin. Colloid Interface Sci.
na

17 2011, 16, 344–349.

18 83. Li, R.; Jiang, P.; Gao, C.; Huang, F.; Xu, R.; Chen, X. Experimental investigation of
ur

19 silica-based nanofluid enhanced oil recovery: The effect of wettability alteration.


Jo

20 Energy and Fuels 2017, 31, 188–197.

21 84. Maurya, N.K.; Kushwaha, P.; Mandal, A. Studies on interfacial and rheological
22 properties of water soluble polymer grafted nanoparticle for application in enhanced
23 oil recovery. J. Taiwan Inst. Chem. Eng. 2017, 70, 319–330.

24 85. Ju, B.; Fan, T. Experimental study and mathematical model of nanoparticle transport
25 in porous media. Powder Technol. 2009, 192, 195–202.

26 86. Ju, B.; Dai, S.; Luan, Z.; Zhu, T.; Su, X.; Qiu, X. A Study of Wettability and
27 Permeability Change Caused by Adsorption of Nanometer Structured Polysilicon on
28 the Surface of Porous Media. Spe 2002.

29 87. Zaitoun, A.; Kohler, N. The Role of Adsorption in Polymer Propagation Through.
30 SPE 1987.
1 88. Broseta, D.; Medjahed, F.; Lecourtier, J.; Robin, M. Polymer Adsorption / Retention
2 in Porous Media: Effects of Core Wettability and Residual Oil. 1995.

3 89. Manrique, E.; De Carvajal, G.; Anselmi, L.; Romero, C.; Chacón, L.
4 Alkali/surfactant/polymer at VLA 6/9/21 field in Maracaibo Lake: Experimental
5 results and pilot project design. SPE - DOE Improv. Oil Recover. Symp. Proc. 2000.

of
ro
-p
re
lP
na
ur
Jo
Highlights

• The additional oil recovery obtained by polymeric nanofluid flooding is generated

by improving macroscopic recovery mechanisms and the activation of microscopic

mechanisms.

• Macroscopic oil displacement was due to the control of the loss of mass by

retention, resulting in preserving the apparent viscosity of the polymeric nanofluids

of
in porous media and mobility control.

ro
• Microscopic oil displacement was mainly caused by a reduction in the size of oil

-p
clusters in the residual oil distribution, which could be attributed to a reduction in

the capillary forces and an increase in the polymer's viscoelasticity nature.


re
• The combination of the micro−macroscopic mechanisms contributed more than
lP

30% of additional oil due to SiO2 nanoparticles' action in the polymeric solution.
na
ur
Jo
Declaration of interests

☒ The authors declare that they have no known competing financial interests or personal relationships
that could have appeared to influence the work reported in this paper.

☐The authors declare the following financial interests/personal relationships which may be considered
as potential competing interests:

of
ro
-p
re
lP
na
ur
Jo

You might also like