Next Article in Journal
Epitaxial Growth and Characterization of Nanoscale Magnetic Topological Insulators: Cr-Doped (Bi0.4Sb0.6)2Te3
Next Article in Special Issue
Synthesis and Characterization of New Nanohybrids Based on Carboxymethyl Scleroglucan and Silica Nanoparticles
Previous Article in Journal
Computational Nanotoxicology Models for Environmental Risk Assessment of Engineered Nanomaterials
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Experimental Investigation of the Viscosity and Stability of Scleroglucan-Based Nanofluids for Enhanced Oil Recovery

by
Rubén H. Castro
1,*,
Laura M. Corredor
2,
Sebastián Llanos
3,
María A. Causil
1,
Adriana Arias
3,
Eduar Pérez
4,
Henderson I. Quintero
2,
Arnold R. Romero Bohórquez
3,
Camilo A. Franco
1 and
Farid B. Cortés
1
1
Grupo de Investigación en Fenómenos de Superficie—Michael Polanyi, Facultad de Minas, Universidad Nacional de Colombia—Sede Medellín, Medellín 050034, Colombia
2
Centro de Innovación y Tecnología—ICP, Ecopetrol S.A., Piedecuesta 681011, Colombia
3
Grupo de Investigación en Química Estructural (GIQUE), Escuela de Química, Universidad Industrial de Santander, Bucaramanga 680006, Colombia
4
Departamento de Ingeniería Mecánica, Universidad Francisco de Paula Santander, Ocaña 546551, Colombia
*
Author to whom correspondence should be addressed.
Nanomaterials 2024, 14(2), 156; https://doi.org/10.3390/nano14020156
Submission received: 30 September 2023 / Revised: 26 October 2023 / Accepted: 30 October 2023 / Published: 10 January 2024
(This article belongs to the Special Issue Advances in Nanomodified Polymers and in Polymer Science)

Abstract

:
Biopolymers emerge as promising candidates for enhanced oil recovery (EOR) applications due to their molecular structures, which exhibit better stability than polyacrylamides under harsh conditions. Nonetheless, biopolymers are susceptible to oxidation and biological degradation. Biopolymers reinforced with nanoparticles could be a potential solution to the issue. The nanofluids’ stability and performance depend on the nanoparticles’ properties and the preparation method. The primary objective of this study was to evaluate the effect of the preparation method and the nanoparticle type (SiO2, Al2O3, and TiO2) on the viscosity and stability of the scleroglucan (SG). The thickening effect of the SG solution was improved by adding all NPs due to the formation of three-dimensional structures between the NPs and the SG chains. The stability test showed that the SG + Al2O3 and SG + TiO2 nanofluids are highly unstable, but the SG + SiO2 nanofluids are highly stable (regardless of the preparation method). According to the ANOVA results, the preparation method and standing time influence the nanofluid viscosity with a statistical significance of 95%. On the contrary, the heating temperature and NP type are insignificant. Finally, the nanofluid with the best performance was 1000 ppm of SG + 100 ppm of SiO2_120 NPs prepared by method II.

Graphical Abstract

1. Introduction

Enhanced oil recovery (EOR) is becoming essential in the global oil supply because production from mature fields is declining and hydrocarbon discoveries are insufficient to meet the growing energy demand [1,2,3]. EOR contributes to maximizing oil reserves, extending the life of fields, and increasing the oil recovery factor. Flooding with water-soluble synthetic polymers, mainly with partially hydrolyzed polyacrylamide (HPAM), has been used to improve oil sweep efficiency by reducing water mobility and water permeability in the swept zone and by diverting the displacing fluid to unswept zones [1,4,5,6,7]. However, synthetic polymers are affected by reservoir temperature, formation water salinity, and hardness [8]. Polysaccharides such as xanthan gum (XG) [9], carboxymethylcellulose (CMC) [10], schizophyllan (SPG) [11], and scleroglucan (SG) [12] have emerged as an alternative to HPAM due to their remarkable rheological properties and resistance to hydrolysis, pH, electrolytes, mechanical shearing, and temperature [13,14].
Scleroglucan is a non-ionic, water-soluble polysaccharide produced by Sclerotium species [12,15,16,17,18]. It can be made with different branching frequencies, side-chain lengths, and molecular weights (ranging from 1.3–3.2·105 to 0.3–6.0·106 Da) depending on the fermentation conditions, the producing species (Sclerotium glucanicum, Sclerotium rolfsii, and Sclerotium delphinii), and the extraction methods [19,20,21]. Sclerotium glucanicum and Sclerotium rolfsii are the main species used for SG production. SG forms a triple-helical structure when dissolved in water. These solutions exhibit a shear thinning behavior but can tolerate high temperatures, a pH range of 1 to 11, and various electrolytes. The rate at which the viscosity of the SG solutions develops depends on purity grade, mixing, temperature, pH, and polymer concentration. Previous studies have shown that improper preparation of the SG solutions contributes to the negative performance of this biopolymer at laboratory and field scales (low viscosity, poor filterability, and formation damage) [15,16,22].
In the last decade, the combination of polymer flooding and nanoparticles (NPs) has been investigated as a promising method to enhance oil recovery through the improvement of the rheological properties of the injected fluid [23,24], reductions in polymer retention and oil-water interfacial tension [25,26], and wettability alteration [23,27,28,29,30,31,32]. The NPs that have shown great potential to enhance oil recovery for both light and heavy oil are SiO2, TiO2, and Al2O3 NPs [33,34]. The NPs can be dispersed in the polymer solution [34,35,36], or the polymer chains can be grafted onto the NP surface [33,37,38,39].
The nanofluids can be prepared by mixing following one-step or two-step methods. The one-step method consists of simultaneously synthesizing and dispersing the nanoparticles into a fluid. In the two-step method, the nanoparticles are first synthesized and then dispersed into the fluid [40]. Employing the mixing of biopolymers with metal oxide NPs for heavy oil recovery, Corredor et al. [23] reported that adding untreated silica, SiO2-MPS, and SiO2-OTES NPs improved the thickening behavior of XG solutions. In contrast, Fe(OH)3, Al2O3, and TiO2 NPs decreased the viscosity of the biopolymer solutions. At 0 ppm and 3000 ppm NaCl, the NPs increased the cumulative oil recovery between 3% and 9% and between 1% and 5%, respectively. However, at 10,000 ppm NaCl, only Fe(OH)3 and TiO2 NPs increased the cumulative oil recovery between 2% and 3%. The differences in the performance of the nanofluids were ascribed to the changes in the electrostatic interactions between NPs-XG-counterions-sand grains. Similarly, Saha and coworkers [41] found that the incorporation of hydrophilic SiO2 NPs into the XG solutions reduced the oil-water IFT, increased the viscosity of the biopolymer solution, stabilized the emulsions, and changed the wettability of the porous media, leading to an increment in the cumulative oil recovery of 20.82% at 30 °C and 18.44% at 80 °C as compared withwater flooding.
Rellegadla et al. [36] reported that nickel-assisted XG flooding yielded the highest recovery of 5.98% residual oil in place (ROIP), compared with 4.48% ROIP of XG flooding and 4.58% ROIP of NP flooding due to the higher intrinsic viscosity of the nickel-XG nanofluid. Orodu et al. [30,35] studied the effect of Al2O3 NPs (30–60 nm) on the performance of the Potato Starch (PSP) and the Gum Arabic (GA). After waterflooding, they reported an incremental oil recovery between 5–12% and 5–7% for the PSP-NPs and the GA-NPs samples. These results were attributed to the increment in the biopolymer’s viscosity and the improvement of their thermal stability caused by the alumina NPs.
In a later work, Rueda et al. [31] evaluated the effect of modified silica nanoparticles (1000 ppm of polymer-coated silica nanoparticles) on the performance of XG and SG solutions (160 and 250 ppm). The results showed that NP-assisted Xanthan flooding achieved the highest ultimate oil recovery at all evaluated conditions due to a more homogenous dispersion of the NPs in the XG solution and reduced polymer adsorption. The dispersion of the NPs in the SG solution was unsuccessful. Buitrago et al. [42] evaluated the effect of the sonication time, the addition order of the components, and the polymer hydration time on the rheological behavior of XG-hydrophilic SiO2 nanofluids. They concluded that the preparation method has little impact on the performance of the nanofluids because all of them exhibited similar rheological behavior and viscosity values.
The two-step method is used in industries to produce nanofluids on a large scale due to its lower production cost than the one-step method. However, this method is challenging to avoid the agglomeration of NPs. No literature investigations have described a specific method to prepare a scleroglucan-based nanofluid and nanoparticles by an easy and practical method that takes advantage of the mechanical effect to solubilize the biopolymer and disperse the nanoparticles [43]. Previous reports, such as Rueda [31] and Buitrago [42], have used lower-purity biopolymer solutions (SG and XG) to prepare nanofluids using magnetic effect and ultrasound (300 W) with long hydration times (24 h, full hydration for 7 days or longer) with no representative viscosity differences, in contrast to this study using a high-performance immersion blender at 20,000 rpm [44,45] to represent a practical setting on a possible field scale and to avoid low hydration effects in EOR polymer preparation [22,31,46]. For this reason, the primary objective of this study is to examine the feasibility of improving the stability and viscosity of SG-based nanofluids prepared by an easy two-step method for EOR applications by changing the preparation method and the NP type.

2. Materials and Methods

2.1. Materials and Reagents

The biopolymer employed was a commercial EOR-grade scleroglucan (SG, purity >99%, 5% humidity) with a molecular weight of 4 × 106 Da. For the preparation of the synthetic brine, 0.83 g/L sodium chloride (NaCl, 99.5% pure, Merck Millipore, Burlington, MA, USA), 0.04 g/L potassium chloride (KCl, 99.5% pure, Merck Millipore, USA), 0.07 g/L magnesium chloride (MgCl2.6H2O, 99% pure, Merck Millipore, USA), and 0.34 g/L calcium chloride (CaCl2.2H2O, 99% pure, Merck Millipore, USA), and type II water (pH ≈ 7) were used. Commercial nanoparticles of SiO2, Al2O3, and TiO2 of different nature, sizes, and surface areas was used for the nanofluid preparations as described in Table 1.

2.2. Methods

2.2.1. Nanofluid Preparation

The brine was prepared in deionized water and filtered through a 0.45 µm MCE membrane filter (Merck Millipore, USA) before use. The biopolymer solution was made as proposed by Abraham and Sumner [43] and Castro et al. [44,45]. The nanofluids were prepared at a fixed concentration of 1000 ppm SG and 100 ppm NPs, following four different methodologies (I, II, III, IV) to determine the effect of the nanofluid preparation on its performance. All nanofluids were stored at a temperature of 30 °C and duplicated in an oven at 60 °C. The methods are described in Table 2 [47].

2.2.2. Turbidity Measurements

The turbidity of the nanofluids determines the particle suspension stability. It was measured at 30 °C by the 8237-absorptiometry method using a Hach 2100P turbidimeter (HATCH, Houston, TX, USA). A standard reference suspension (1–10 NTU: accuracy ±0.1%). For higher values, an accuracy of ±10% was used for the turbidimeter calibration.

2.2.3. Viscosity Test

The viscosities of the samples heated at 30 and 60 °C in an oven for 21 days were measured at 30 °C in a DV3TTM rheometer (AMETEK Brookfield, Middleborough, MA, USA) with an Ultra Low Adapter (ULA, µ < 100 cP, Accuracy: ±1.0%, Repeatability: ±0.2%) by changing the spindle depending on the viscosity of the sample (4.24–106 1/s). The uncertainties in the viscosity results were 1% of the reported value, according to the oil standard reference. The nominal viscosities were determined according to the API RP63 standard [48] at 6 rpm, equivalent to 7.3 1/s for the ULA spindle ( γ ˙ = 1.224 rpm). This shear rate was selected for a sandstone formation, with shear rate values between 7 and 10 1/s [49]. Finally, Analysis of Variance (ANOVA) was used to evaluate the influence of continuous and categorical variables on the viscosity values of all nanofluids [50].

2.2.4. Rheological Behavior

The rheological behavior of the nanofluids was measured at 30 °C over the range of 1–100 1/s. All the viscosity data exhibit a good fit for the Carreau–Yasuda model [51,52].
μ = η + ( η 0 η ) 1 + ( λ γ ˙ ) α ( n 1 ) / α
This model describes the behavior of non-Newtonian fluids [36] as a function of the zero shear viscosity η 0 (cP), the infinite shear viscosity η (cP), the effective shear rate, γ ˙ (1/s), the relaxation parameter λ (s) (which limits the transition zone between dilatant and pseudoplastic behavior), the power law exponent n (dimensionless), and the transition parameter α (dimensionless, describes the transition of the behavior at time zero and the critical point of the shear rate) [53].

3. Results

3.1. Nanofluid’s Viscosity

The SG concentration used in all the experiments was set at 1000 ppm, according to the results previously reported by the authors [44,45]. Table 3 displays the viscosity values of the nanofluids prepared by method I. It is observed that the viscosity of the SG solutions and the nanofluids were not affected by temperature (30 and 60 °C, Table 3), which was expected due to the high thermal stability of the biopolymer. Adding all NPs positively affects the viscosity of the SG solutions (increments up to 11.3%, 6.2% average). The increment in viscosity can be attributed to the interactions between the glycosidic groups of the SG with the OH groups on the surface of the NPs through hydrogen bonding (Figure 1) and the hydrophobic interactions between the APTES on the NPs (Figure 2) and the backbone of the biopolymer. Furthermore, the NPs act as crosslinkers between the SG chains.
Despite the viscosity results, the nanofluids containing TiO2_65, Al2O3_120, Al2O3_180, and Al2O3_35 NPs should not be considered as EOR additives due to their low dispersity into the biopolymer solutions, attributed to the low density of silanol groups on their surface [44,45].
The viscosity values of the nanofluids prepared by method II are presented in Table 4. Adding all NPs to the SG solution increased its viscosity up to 15.2% (8.9% average). At Theating = 30 °C, the NPs with the highest viscosifying effect were alumina Al2O3_120, Al2O3_180, and SiO2_380. However, Al2O3_120, Al2O3_180, Al2O3_35, and TiO2_65 NPs were unstable in the SG solution. At Theating = 60 °C, the highest viscosity values were obtained with all SiO2 NPs due to their higher stability than the Al2O3 and TiO2 NPs.
Table 5 presents the viscosity values of the nanofluids prepared by method III. Adding the NPs increased the viscosity of the SG solution up to 12.7% (5.4% average, Table 5). As in methods I and II, Al2O3_120, Al2O3_180, Al2O3_35, and TiO2_65 NPs were unstable in the SG solution.
In method IV, the increments in the SG viscosity by adding the NPs reached up to 10.8% (3.7% average, Table 6). The SG + Al2O3 and SG + TiO2 nanofluids were unstable as in the previous preparation methods.
All of the nanofluids tested have a decrease in viscosity (measured at 30 °C) after heating the samples for 0, 7, 14, and 21 days at 60 °C. This effect is caused by the weakening of the intermolecular forces between the NPs and the SG chains. Furthermore, the agglomeration of the NPs over time increases their particle size, reducing their Brownian velocity. When the Brownian velocity reaches terminal settling velocity, the NPs cannot overcome the gravitational force and precipitate in the nanofluids, causing viscosity reduction over time [54].
By comparing all viscosity data, it can be concluded that the nanofluids prepared by methods I and II exhibited the highest viscosity values. When the NPs are dispersed in the SG solution instead of in water (methods III and IV), the stability of the suspension increases because the viscosity of the dispersion medium is higher (Stoke’s law) [55].
Comparing methods I and II, it is observed that the dispersion of the NPs is affected by the stirring speed. At high stirring speed (method I @ 20,000 rpm × 5 min vs. method II @ 500 rpm × 60 min), the NPs move at the side of the beaker wall without being distributed throughout the biopolymer solution. It reduces the interaction with NP-SG, leading to lower viscosity increments for the nanofluids prepared by method I. Accordingly, method II is the one recommended for preparing SG-based nanofluids.

3.2. Nanofluid’s Stability

Visual observation and turbidity measurements were used to study the stability of the SG-based nanofluids. During the visual stability test (see the images of the nanofluids in Appendix A, Figure A1, Figure A2, Figure A3 and Figure A4), the SG + TiO2 and SG + Al2O3 nanofluids exhibited rapid agglomeration and settlement. In contrast, the SG + SiO2 nanofluids were stable for more than 21 days.
Accordingly, the SG + SiO2 nanofluids exhibited the lowest turbidity values because the good dispersion of the SiO2 NPs in the SG solution reduced the amount of light scattered (Table 7). The SG + TiO2 and SG + Al2O3 nanofluids showed the highest turbidity values because the low interaction between the NPs and the SG chains causes the agglomeration of the NPs and higher light scattering in the nanofluid. The SG + TiO2 and SG + Al2O3 nanofluids prepared by method II showed lower turbidity values than the other methods due to the rapid precipitation of the NPs observed during the visual stability test. From these results, it can be concluded that SG + SiO2 are the most stable nanofluids, regardless of the preparation method (Appendix A, Figure A5).

3.3. Statistical Analysis

The statistical decision tree determined the best preparation method and nanofluid (Appendix A, Figure A6). The results show methods I and II, and SG + SiO2_120 and SG + SiO2_APTES_120 nanofluids (Figure A6, blue line).
The analysis of the experimental data was conducted following a 2k factorial design using R statistical software (version 4.2.2) to examine the effects. The interactions of preparation method, standing time, heating temperature, and nanoparticle type on the viscosity of the nanofluids within an empirically selected range of high (1) and low (−1) levels are summarized in Table 8. The standing time and heating temperature are continuous variables, while the preparation method and NP type are discrete variables.
The highest viscosity values are obtained with method II (Figure 3a) using SiO2_120 NPs (Figure 3d) after 21 days of standing time (high level, Figure 3b) at 30 °C (low level, Figure 3c).
The previous analysis provided the optimal levels of each factor but could not determine which factors impact the response variable (viscosity) most. To achieve this, an analysis of variance (ANOVA [56]) was performed. According to the ANOVA results (Table 9), the preparation method and standing time influence the nanofluid viscosity with a statistical significance of 95% [50]. On the contrary, the heating temperature and NP type are insignificant (p > 0.05) [57].
In this study, it is also observed that nanoparticles SiO2_120 and SiO2_APTES_120, spherical with average sizes of 20 nm and a superficial area of 120 m2/g, are present in the solutions with higher viscosity. Similarly, nanoparticles SiO2_640, amorphous porous nanoparticles of 20 nm but with a higher superficial area (640 m2/g), generated greater dispersions, increasing the variability in the viscosity data. On the other hand, SiO2 spherical nanoparticles of size 20 nm significantly increase viscosity (Figure A6). According to Keblinski et al. [58], a low nanoparticle size (at constant concentration) will augment the viscosity value of nanofluids due to an interparticle spacing decrease, which intensifies interparticle interaction and generates greater aggregate structure.

3.4. Rheological Behavior of Nanofluid

Figure 4 shows that SG and nanofluid SG + SiO2_120 solutions exhibited a shear thinning behavior with pseudo-plasticity indices (n) less than one [59], where hydroxyl groups over silica nanoparticles surface could contribute to improving the interactions with OH groups around the rod-like structure of the Scleroglucan [60].
For a rheological explanation of the increase in bulk viscosity with the addition of the nanofluid prepared by method II, the rheological parameters of the Carreau–Yasuda model, such as relaxation time, viscoelasticity index, and viscosity at zero and infinite time, are essential to establishing the nanoparticle effect. The infinite shear viscosity and the relaxation parameter were fixed at 0.458 cP and 1.5 s, respectively [8]. The Carreau–Yasuda model parameters of SG and nanofluid SG + SiO2_120 are shown in Table 10. Nanofluid SG + SiO2_120 viscosity parameters at zero times ( η 0 ) and relaxation times ( λ ) are higher than the SG solution. The pseudo-plasticity index (n) is slightly higher for nanofluid SG + SiO2_120 due to its higher pseudo-plasticity compared with the SG solution related to the rising in the molecular entanglement [61,62].

4. Conclusions

This paper provides insights into the effect of the preparation method and NP type on the stability and viscosity of SG-based nanofluids. The addition of all NPs improved the thickening behavior of the SG solution due to the formation of three-dimensional structures between the NPs and the SG polymeric chains. These structures are formed through hydrogen bonding between the glycosidic groups of the SG and the silanol groups on the surface of the NPs or through hydrophobic interactions between the APTES on the NPs and the backbone of the SG.
From the visual observation and the turbidity monitoring, the SG + SiO2 exhibited the lowest turbidity values and good dispersion because hydroxyl groups over the surface of silica nanoparticles could contribute to improving the OH groups interactions with the structure of Scleroglucan, regardless of the preparation method. Furthermore, the SG + Al2O3 and SG + TiO2 nanofluids showed the highest turbidity values because the low interaction between the NPs and the SG chains caused the agglomeration of the NPs and higher light scattering in the nanofluid, attributed to the low density of silanol groups on their NP surface. The selected method for the nanofluid preparation was method II, and the nanofluid with a higher viscosity increment was 1000 ppm of SG + 100 ppm of SiO2_120 NPs.
Finally, it can be concluded that the nanofluid SG + SiO2_120 exhibits higher zero-shear viscosity, pseudo-plasticity index, and relaxation times due to its higher molecular entanglement compared with the SG solution.

Author Contributions

Conceptualization, R.H.C., L.M.C., C.A.F. and F.B.C.; methodology, L.M.C., A.R.R.B., C.A.F. and F.B.C.; validation, L.M.C., H.I.Q., A.R.R.B., C.A.F. and F.B.C.; formal analysis, R.H.C., L.M.C. and F.B.C.; investigation, R.H.C., S.L., M.A.C., A.A., E.P. and F.B.C.; resources, L.M.C., C.A.F. and F.B.C.; data curation, S.L., E.P. and R.H.C.; writing—original draft preparation, R.H.C., L.M.C. and S.L.; writing—review and editing, R.H.C., L.M.C., C.A.F. and F.B.C.; visualization, R.H.C. and F.B.C.; supervision, L.M.C., H.I.Q., A.R.R.B., C.A.F. and F.B.C. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Ecopetrol S.A., Universidad Nacional de Colombia—Sede Medellín, and MinCiencias by Agreement FP44842-326-2017 (D.Nal Empresa 758 of 2016).

Institutional Review Board Statement

Not applicable.

Data Availability Statement

The data presented in this study are available on request from the corresponding author.

Acknowledgments

The authors thank Ecopetrol S.A, the Universidad Nacional de Colombia—Sede Medellín, and MinCiencias for the logistical and financial support.

Conflicts of Interest

The authors declare no conflict of interest.

Appendix A

Figure A1. Visual stability test and turbidity values of the nanofluids prepared by method I after heating the samples for 0 and 21 days at 30 °C.
Figure A1. Visual stability test and turbidity values of the nanofluids prepared by method I after heating the samples for 0 and 21 days at 30 °C.
Nanomaterials 14 00156 g0a1
Figure A2. Visual stability test and turbidity values of the nanofluids prepared by method II after heating the samples for 0 and 21 days at 30 °C.
Figure A2. Visual stability test and turbidity values of the nanofluids prepared by method II after heating the samples for 0 and 21 days at 30 °C.
Nanomaterials 14 00156 g0a2
Figure A3. Visual stability test and turbidity values of the nanofluids prepared by method III after heating the samples for 0 and 21 days at 30 °C.
Figure A3. Visual stability test and turbidity values of the nanofluids prepared by method III after heating the samples for 0 and 21 days at 30 °C.
Nanomaterials 14 00156 g0a3
Figure A4. Visual stability test and turbidity values of the nanofluids prepared by method IV after heating the samples for 0 and 21 days at 30 °C.
Figure A4. Visual stability test and turbidity values of the nanofluids prepared by method IV after heating the samples for 0 and 21 days at 30 °C.
Nanomaterials 14 00156 g0a4
Figure A5. Comparison of the turbidity values of the nanofluids at 0 days and 30 °C.
Figure A5. Comparison of the turbidity values of the nanofluids at 0 days and 30 °C.
Nanomaterials 14 00156 g0a5
Figure A6. Statistical decision tree.
Figure A6. Statistical decision tree.
Nanomaterials 14 00156 g0a6

References

  1. Li, X.; Zhang, F.; Liu, G. Review on polymer flooding technology. IOP Conf. Ser. Earth Environ. Sci. 2021, 675, 012199. [Google Scholar] [CrossRef]
  2. Babadagli, T. Development of mature oil fields—A review. J. Pet. Sci. Eng. 2007, 57, 221–246. [Google Scholar] [CrossRef]
  3. Manrique, E.J.; Thomas, C.P.; Ravikiran, R.; Izadi Kamouei, M.; Lantz, M.; Romero, J.L.; Alvarado, V. EOR: Current status and opportunities. In Proceedings of the SPE Improved Oil Recovery Symposium, Tulsa, OK, USA, 24–28 April 2010; Society of Petroleum Engineers: Dallas, TX, USA, 2010. [Google Scholar]
  4. Manrique, E.; Ahmadi, M.; Samani, S. Historical and recent observations in polymer floods: An update review. CTF-Cienc. Tecnol. Y Futuro 2017, 6, 17–48. [Google Scholar] [CrossRef]
  5. Alvarado, V.; Manrique, E. Enhanced oil recovery: An update review. Energies 2010, 3, 1529–1575. [Google Scholar] [CrossRef]
  6. Alvarado, V.; Manrique, E. Enhanced Oil Recovery: Field Planning and Development Strategies; Gulf Professional Publishing: Sebastopol, CA, USA, 2010. [Google Scholar]
  7. Xiangguo, L.; Bao, C.; Kun, X.; Weijia, C.; Yigang, L.; ZHANG, Y.; Xiaoyan, W.; ZHANG, J. Enhanced oil recovery mechanisms of polymer flooding in a heterogeneous oil reservoir. Pet. Explor. Dev. 2021, 48, 169–178. [Google Scholar]
  8. Jouenne, S. Polymer flooding in high temperature, high salinity conditions: Selection of polymer type and polymer chemistry, thermal stability. J. Pet. Sci. Eng. 2020, 195, 107545. [Google Scholar] [CrossRef]
  9. Jang, H.Y.; Zhang, K.; Chon, B.H.; Choi, H.J. Enhanced oil recovery performance and viscosity characteristics of polysaccharide xanthan gum solution. J. Ind. Eng. Chem. 2015, 21, 741–745. [Google Scholar] [CrossRef]
  10. Rellegadla, S.; Jain, S.; Agrawal, A. A holistic approach to determine the enhanced oil recovery potential of hydroxyethylcellulose, tragacanth gum and carboxymethylcellulose. J. Mol. Liq. 2021, 341, 117334. [Google Scholar] [CrossRef]
  11. Ahmad Wazir, N.; Abdul Manap, A.A. Schizophyllan as Potential Environmental Friendly EOR Polymer for A Selected Malaysian Field. In Proceedings of the Offshore Technology Conference Asia, Kuala Lumpur, Malaysia, 2–6 November 2020; OnePetro: Richardson, TX, USA; Society of Petroleum Engineers: Dallas, TX, USA, 2020. [Google Scholar]
  12. Kozlowicz, B.; Jensen, T.; Khambete, M.; Kadhum, M.J.; Garshol, F.K.; Vik, E.A.; Tomczak-Wandzel, R.; Wever, D.A.; Lomans, B.P.; Ray, C. Scleroglucan Polymer Stability: Thermal, Chemical, and Microbial. In Proceedings of the SPE Improved Oil Recovery Conference, Tulsa, OK, USA, 31 August–4 September 2020. [Google Scholar]
  13. Pu, W.; Shen, C.; Wei, B.; Yang, Y.; Li, Y. A comprehensive review of polysaccharide biopolymers for enhanced oil recovery (EOR) from flask to field. J. Ind. Eng. Chem. 2018, 61, 1–11. [Google Scholar] [CrossRef]
  14. Clinckspoor, K.-J.; Sousa-Ferreira, V.-H.d.; Barros-Zanoni-Lopes-Moreno, R.J.C.; F-Ciencia, T.y.F. Bulk rheology characterization of biopolymer solutions and discussions of their potential for enhance. CTF-Cienc. Tecnol. Y Futuro 2021, 11, 123–135. [Google Scholar] [CrossRef]
  15. Fournier, R.; Tiehi, J.-E.; Zaitoun, A. Laboratory study of a new EOR-Grade Scleroglucan. In Proceedings of the SPE EOR Conference at Oil and Gas West Asia, Muscat, Oman, 26–28 March 2018; OnePetro: Richardson, TX, USA, 2018. [Google Scholar]
  16. Rivenq, R.; Donche, A.; Nolk, C. Improved scleroglucan for polymer flooding under harsh reservoir conditions. SPE Reserv. Eng. 1992, 7, 15–20. [Google Scholar] [CrossRef]
  17. Kalpakci, B.; Jeans, Y.; Magri, N.; Padolewski, J. Thermal stability of scleroglucan at realistic reservoir conditions. In Proceedings of the SPE/DOE Enhanced Oil Recovery Symposium, Tulsa, OK, USA, 22–25 April 1990; Society of Petroleum Engineers: Dallas, TX, USA, 1990. [Google Scholar]
  18. Rodríguez-Mateus, Z.-P.; Angarita, R.-C.; Niño-Gómez, J.-A.; Corredor, L.-M.; Llanos Gallo, S.; Quintero, H.; Castro-García, R.-H.J.C.; F-Ciencia, T.y.F. Biodegradation and toxicity of scleroglucan for enhanced oil recovery. CTF-Cienc. Tecnol. Futuro 2022, 12, 5–12. [Google Scholar] [CrossRef]
  19. Viñarta, S.C.; François, N.J.; Daraio, M.E.; Figueroa, L.I.; Fariña, J. Sclerotium rolfsii scleroglucan: The promising behavior of a natural polysaccharide as a drug delivery vehicle, suspension stabilizer and emulsifier. Int. J. Biol. Macromol. 2007, 41, 314–323. [Google Scholar] [CrossRef] [PubMed]
  20. Viñarta, S.C.; Molina, O.; Figueroa, L.; Fariña, J. A further insight into the practical applications of exopolysaccharides from Sclerotium rolfsii. Food Hydrocoll. 2006, 20, 619–629. [Google Scholar] [CrossRef]
  21. Castillo, N.A.; Valdez, A.L.; Fariña, J.I. Microbial production of scleroglucan and downstream processing. Front. Microbiol. 2015, 6, 1106. [Google Scholar] [CrossRef]
  22. Seright, R.S.; Wavrik, K.E.; Zhang, G.; AlSofi, A.M. Stability and behavior in carbonate cores for new enhanced-oil-recovery polymers at elevated temperatures in hard saline brines. SPE Reserv. Eval. Eng. 2021, 24, 1–18. [Google Scholar] [CrossRef]
  23. Corredor, L.M.; Husein, M.M.; Maini, B.B. Effect of hydrophobic and hydrophilic metal oxide nanoparticles on the performance of xanthan gum solutions for heavy oil recovery. Nanomaterials 2019, 9, 94. [Google Scholar] [CrossRef]
  24. Corredor-Rojas, L.M.; Sarapardeh, A.H.; Husein, M.M.; Dong, P.M.; Maini, B.B. Rheological behavior of surface modified silica nanoparticles dispersed in Partially Hydrolyzed Polyacrylamide and Xanthan Gum solutions: Experimental measurements, mechanistic understanding, and model development. Energy Fuels 2018, 32, 10628–10638. [Google Scholar] [CrossRef]
  25. Cheraghian, G.; Khalili Nezhad, S.S.; Kamari, M.; Hemmati, M.; Masihi, M.; Bazgir, S. Adsorption polymer on reservoir rock and role of the nanoparticles, clay and SiO2. Int. Nano Lett. 2014, 4, 114. [Google Scholar] [CrossRef]
  26. Kadhum, M.J.; Swatske, D.; Weston, J.; Resasco, D.E.; Shiau, B.; Harwell, J.H. Polymer-stabilized multi-walled carbon nanotube dispersions in high-salinity brines. Energy Fuels 2017, 31, 5024–5030. [Google Scholar] [CrossRef]
  27. Yuan, H.; Zhang, Z.; Mi, Y.; Ye, F.; Liu, W.; Kuan, J.; Jiang, X.; Luo, Y. Demulsification of water-containing crude oil driven by environmentally friendly SiO2@ CS composite materials. Energy Fuels 2020, 34, 8316–8324. [Google Scholar] [CrossRef]
  28. Maurya, N.K.; Kushwaha, P.; Mandal, A. Studies on interfacial and rheological properties of water soluble polymer grafted nanoparticle for application in enhanced oil recovery. J. Taiwan Inst. Chem. Eng. 2017, 70, 319–330. [Google Scholar] [CrossRef]
  29. Ali, J.A.; Kolo, K.; Manshad, A.K.; Mohammadi, A.H. Recent advances in application of nanotechnology in chemical enhanced oil recovery: Effects of nanoparticles on wettability alteration, interfacial tension reduction, and flooding. Egypt. J. Pet. 2018, 27, 1371–1383. [Google Scholar] [CrossRef]
  30. Orodu, K.B.; Afolabi, R.O.; Oluwasijuwomi, T.D.; Orodu, O.D. Effect of aluminum oxide nanoparticles on the rheology and stability of a biopolymer for enhanced oil recovery. J. Mol. Liq. 2019, 288, 110864. [Google Scholar] [CrossRef]
  31. Rueda, E.; Akarri, S.; Torsæter, O.; Moreno, R.B. Experimental investigation of the effect of adding nanoparticles to polymer flooding in water-wet micromodels. Nanomaterials 2020, 10, 1489. [Google Scholar] [CrossRef]
  32. Corredor, L.M.; Husein, M.M.; Maini, B.B. Impact of PAM-grafted nanoparticles on the performance of hydrolyzed polyacrylamide solutions for heavy oil recovery at different salinities. Ind. Eng. Chem. Res. 2019, 58, 9888–9899. [Google Scholar] [CrossRef]
  33. Corredor, L.M.; Husein, M.M.; Maini, B.B. A review of polymer nanohybrids for oil recovery. Adv. Colloid Interface Sci. 2019, 272, 102018. [Google Scholar] [CrossRef]
  34. Khan, M.B.; Khoker, M.F.; Husain, M.; Ahmed, M.; Anwer, S. Effects of nanoparticles on rheological behavior of polyacrylamide related to enhance oil recovery. Acad. J. Polym. Sci 2018, 1, 555573. [Google Scholar]
  35. Orodu, O.D.; Orodu, K.B.; Afolabi, R.O.; Dafe, E.A. Rheology of Gum Arabic Polymer and Gum Arabic Coated Nanoparticle for enhanced recovery of Nigerian medium crude oil under varying temperatures. Data Brief 2018, 19, 1773–1778. [Google Scholar] [CrossRef]
  36. Rellegadla, S.; Bairwa, H.K.; Kumari, M.R.; Prajapat, G.; Nimesh, S.; Pareek, N.; Jain, S.; Agrawal, A. An effective approach for enhanced oil recovery using nickel nanoparticles assisted polymer flooding. Energy Fuels 2018, 32, 11212–11221. [Google Scholar] [CrossRef]
  37. Cao, J.; Song, T.; Zhu, Y.; Wang, S.; Wang, X.; Lv, F.; Jiang, L.; Sun, M. Application of amino-functionalized nanosilica in improving the thermal stability of acrylamide-based polymer for enhanced oil recovery. Energy Fuels 2018, 32, 246–254. [Google Scholar] [CrossRef]
  38. Ye, Z.; Qin, X.; Lai, N.; Peng, Q.; Li, X.; Li, C. Synthesis and performance of an acrylamide copolymer containing nano-SiO2 as enhanced oil recovery chemical. J. Chem. 2013, 2013, 437309. [Google Scholar] [CrossRef]
  39. Lai, N.; Guo, X.; Zhou, N.; Xu, Q. Shear resistance properties of modified Nano-SiO2/AA/AM copolymer oil displacement agent. Energies 2016, 9, 1037. [Google Scholar] [CrossRef]
  40. Urmi, W.T.; Rahman, M.; Kadirgama, K.; Ramasamy, D.; Maleque, M.A. An overview on synthesis, stability, opportunities and challenges of nanofluids. Mater. Today Proc. 2021, 41, 30–37. [Google Scholar] [CrossRef]
  41. Saha, R.; Uppaluri, R.V.; Tiwari, P. Silica Nanoparticle Assisted Polymer Flooding of Heavy Crude Oil: Emulsification, Rheology, and Wettability Alteration Characteristics. Ind. Eng. Chem. Res. 2018, 57, 6364–6376. [Google Scholar] [CrossRef]
  42. Rincon, D.B.; Sadtler, V.; Ojeda, R.M.; Carmes, T.R.; Marchal, P.; Avella, J.P.; Lemaitre, C. Effect of Silica Nanoparticles in Xanthan Gum Solutions: Rheological Behavior and Preparation Methods of Nanofluids. Chem. Eng. Trans. 2021, 86, 1171–1176. [Google Scholar]
  43. Abraham, T.W.; Sumner, E.S. Method for Solubilizing Biopolymer Solids for Enhanced Oil Recovery Applications. U.S. Patent 16/089,740, 18 April 2019. [Google Scholar]
  44. RubCastro-García, R.H.; LLanos-Gallo, S.; Rodríguez-Ardila, J.L.; Quintero-Perez, H.I.; Zapata, J.F.; Manrique, E. Heavy Oil and High-Temperature Polymer EOR Applications. Cienc. Tecnol. Futuro—CTF. J. Oil Gas Altern. Energy Sources 2020, 10, 73–83. [Google Scholar] [CrossRef]
  45. Castro, R.H.; Llanos, S.; Rodríguez, J.; Quintero, H.I.; Manrique, E. Polymers for EOR Application in High Temperature and High Viscosity Oils: Rock–Fluid Behavior. Energies 2020, 13, 5944. [Google Scholar] [CrossRef]
  46. Gao, X.; Huang, L.; Xiu, J.; Yi, L.; Zhao, Y. A Study on Viscosity Increasing and Rheological Properties of Novel Biopolymers in Extreme Reservoirs. Preprints 2023, 2023100462. [Google Scholar] [CrossRef]
  47. Giraldo, L.J.; Giraldo, M.A.; Llanos, S.; Maya, G.; Zabala, R.D.; Nassar, N.N.; Franco, C.A.; Alvarado, V.; Cortés, F.B. The effects of SiO2 nanoparticles on the thermal stability and rheological behavior of hydrolyzed polyacrylamide based polymeric solutions. J. Pet. Sci. Eng. 2017, 159, 841–852. [Google Scholar] [CrossRef]
  48. American Petroleum Institute. Recommended Practices for Evaluation of Polymers Used in Enhanced Oil Recovery Operations; American Petroleum Institute: Singapore, 1990. [Google Scholar]
  49. Wassmuth, F.; Green, K.; Arnold, W.; Cameron, N. Polymer flood application to improve heavy oil recovery at East Bodo. J. Can. Pet. Technol. 2009, 48, 55–61. [Google Scholar] [CrossRef]
  50. van Wageningen-Kessels, F.; Van Lint, H.; Vuik, K.; Hoogendoorn, S. Genealogy of traffic flow models. EURO J. Transp. Logist. 2015, 4, 445–473. [Google Scholar] [CrossRef]
  51. Carreau, P.J. Rheological equations from molecular network theories. Trans. Soc. Rheol. 1972, 16, 99–127. [Google Scholar] [CrossRef]
  52. Osswald, T.; Rudolph, N. Polymer Rheology; Carl Hanser: München, Germany, 2015. [Google Scholar]
  53. Yasuda, K. Investigation of the Analogies between Viscometric and Linear Viscoelastic Properties of Polystyrene Fluids; Massachusetts Institute of Technology: Cambridge, MA, USA, 1979. [Google Scholar]
  54. Hemadri, V.; Mane, N. Study of the effect of preparation parameters on thermal conductivity of metal oxide nanofluids using Taguchi method. J. Energy Syst. 2021, 5, 149–164. [Google Scholar] [CrossRef]
  55. Larsson, M.; Hill, A.; Duffy, J. Suspension stability; why particle size, zeta potential and rheology are important. Annu. Trans. Nord. Rheol. Soc. 2012, 20, 209–214. [Google Scholar]
  56. Barkhordar, A.; Ghasemiasl, R.; Armaghani, T. Statistical study and a complete overview of nanofluid viscosity correlations: A new look. J. Therm. Anal. Calorim. 2022, 147, 7099–7132. [Google Scholar] [CrossRef]
  57. Hu, X.; Yin, D.; Chen, X.; Xiang, G. Experimental investigation and mechanism analysis: Effect of nanoparticle size on viscosity of nanofluids. J. Mol. Liq. 2020, 314, 113604. [Google Scholar] [CrossRef]
  58. Keblinski, P.; Phillpot, S.; Choi, S.; Eastman, J. Mechanisms of heat flow in suspensions of nano-sized particles (nanofluids). Int. J. Heat Mass Transf. 2002, 45, 855–863. [Google Scholar] [CrossRef]
  59. Barnes, H.A. A Handbook of Elementary Rheology; University of Wales, Institute of Non-Newtonian Fluid Mechanics: Cardiff, UK, 2000. [Google Scholar]
  60. Dehkordi, S.A.H.; Golbodaqi, M.; Mortazavi-Manesh, A.; Safari, N.; Bahadoran, F.; Haghighat, M.F. Dimethyl ether from methanol on mesoporous γ-alumina catalyst prepared from surfactant free highly porous pseudo-boehmite. Mol. Catal. 2023, 538, 113004. [Google Scholar] [CrossRef]
  61. Hernandez, F.A.T.; Niño, J.C.L.; Moreno, R.L. Effects of salts and temperature on rheological and viscoelastic behavior of low molecular weight HPAM solutions. Rev. Fuentes 2018, 16, 19–35. [Google Scholar] [CrossRef]
  62. Chhabra, R.P. Non-Newtonian fluids: An introduction. In Rheology of Complex Fluids; Springer: New York, NY, USA, 2010; pp. 3–34. [Google Scholar]
Figure 1. Schematic of the interaction between the SG chain and the hydrophilic NPs in the aqueous phase.
Figure 1. Schematic of the interaction between the SG chain and the hydrophilic NPs in the aqueous phase.
Nanomaterials 14 00156 g001
Figure 2. Schematic of the interaction between the SG chain and the SiO2_APTES_120 NPs in the aqueous phase.
Figure 2. Schematic of the interaction between the SG chain and the SiO2_APTES_120 NPs in the aqueous phase.
Nanomaterials 14 00156 g002
Figure 3. Viscosity as a function of (a) preparation method, (b) standing time, (c) temperature, and (d) nanoparticle type.
Figure 3. Viscosity as a function of (a) preparation method, (b) standing time, (c) temperature, and (d) nanoparticle type.
Nanomaterials 14 00156 g003
Figure 4. Rheological behavior of the SG and nanofluid SG + SiO2_120 solutions prepared by method II at 30 °C.
Figure 4. Rheological behavior of the SG and nanofluid SG + SiO2_120 solutions prepared by method II at 30 °C.
Nanomaterials 14 00156 g004
Table 1. Nanoparticles are used to prepare the SG-based nanofluids.
Table 1. Nanoparticles are used to prepare the SG-based nanofluids.
NameDescriptionSupplier
SiO2_120SiO2 (20 nm, 120 m2/g, spherical, hydrophilic)Nanostructured & Amorphous Materials Inc., Los Alamos, NM, USA
SiO2_APTES_120SiO2 (20 nm, 120 m2/g, spherical, amphiphilic, surface coated by (3-aminopropyl) triethoxysilane 2%—APTES
SiO2_640SiO2 (20 nm, 640 m2/g, amorphous porous)
Al2O3_120Al2O3 (10 nm, 120 m2/g, spherical, gamma, hydrophilic)
Al2O3_180Al2O3 (20–30 nm, 180 m2/g, nearly spherical, gamma, hydrophilic)
Al2O3_35Al2O3 (27–43 nm, 35 m2/g, mainly alpha contains 5–10% gamma, hydrophilic)
SiO2_380SiO2 (12–15 nm, 380 m2/g, amorphous, hydrophilic)Evonik industries, Allentown, PA, USA
SiO2_200SiO2 (12 nm, 200 m2/g, amorphous, hydrophilic)
TiO2_65Titanium (IV) oxide (21 nm, 35–65 m2/g)Sigma Aldrich, St. Louis, MO, USA
Table 2. Description of the preparation methods of the Scleroglucan-based nanofluids.
Table 2. Description of the preparation methods of the Scleroglucan-based nanofluids.
MethodStep 1Step 2Step 3
IDissolve the SG powder into the brine under mechanical stirring at 500 rpm for 10 min. Then, stir the sample at 800 rpm and 40 °C for 10 min. Finally, homogenize the solution for 5 min using a high-performance immersion blender (IKA™ T 25 Digital Ultra-Turrax)Add the NPs to the SG solutionStir the nanofluid with the Ultra-Turrax at 20,000 rpm for 5 min
IISame as described in method I (step 1)Add the NPs to the SG solution Stir the nanofluid with the propeller agitator at 500 rpm for 60 min
IIIDisperse the NPs in brine and ultrasonicate the dispersions for 1 h Same as described in method I (step 1)-
IVAdd the SG powder and the NPs into the brine simultaneously. Stir the sample with a metallic blade for 10 min at 500 rpm. Then, stir the sample at 800 rpm and 40 °C for 10 min. Finally, stir the dispersion with the Ultra-Turrax at 20,000 rpm for 5 min.--
Table 3. Viscosity measurements of the SG-based nanofluid prepared by method I at 7.3 s−1 and 30 °C after heating the samples for 0, 7, 14, and 21 days at 30 °C and 60 °C.
Table 3. Viscosity measurements of the SG-based nanofluid prepared by method I at 7.3 s−1 and 30 °C after heating the samples for 0, 7, 14, and 21 days at 30 °C and 60 °C.
Heating Temperature (°C)SampleViscosity, cPViscosity Changes of the SG Solution
071421071421
30SiO2_12063.7667.1467.9466.984.5%10.0%11.3%9.8%
SiO2_APTES_12063.5664.6666.9667.384.2%6.0%9.7%10.4%
SiO2_64063.7064.4466.9666.904.4%5.6%9.7%9.6%
Al2O3_12063.9265.6264.1267.364.8%7.5%5.1%10.4%
Al2O3_18061.8066.9064.5067.741.3%9.6%5.7%11.0%
Al2O3_3564.6465.2065.8666.245.9%6.9%7.9%8.6%
SiO2_38062.9865.7065.8667.183.2%7.7%7.9%10.1%
SiO2_20065.2466.2265.3867.226.9%8.5%7.1%10.2%
TiO2_6563.7064.7064.3665.564.4%6.0%5.5%7.4%
SG61.0261.0062.9061.700.0%0.0%3.1%1.1%
60SiO2_12065.9064.9865.0264.528.0%6.5%6.6%5.7%
SiO2_APTES_12065.8866.3866.0665.948.0%8.8%8.3%8.1%
SiO2_64064.5864.4064.3064.025.8%5.5%5.4%4.9%
Al2O3_12064.5866.8065.3064.945.8%9.5%7.0%6.4%
Al2O3_18064.1665.0667.0865.705.1%6.6%9.9%7.7%
Al2O3_3563.9863.9864.8464.444.9%4.9%6.3%5.6%
SiO2_38064.6863.8065.1063.666.0%4.6%6.7%4.3%
SiO2_20064.1065.4465.6064.125.0%7.2%7.5%5.1%
TiO2_6562.5663.3263.4663.582.5%3.8%4.0%4.2%
SG61.0261.8262.6660.880.0%1.3%2.7%−0.2%
Table 4. Viscosity measurements (cP) of the SG-based nanofluid prepared by method II at 7.3 s−1 and 30 °C after heating the samples for 0, 7, 14, and 21 days at 30 °C and 60 °C.
Table 4. Viscosity measurements (cP) of the SG-based nanofluid prepared by method II at 7.3 s−1 and 30 °C after heating the samples for 0, 7, 14, and 21 days at 30 °C and 60 °C.
Heating Temperature (°C)SampleViscosity, cPViscosity Changes of the SG Solution
071421071421
30SiO2_12067.568.969.367.610.6%12.9%13.6%10.8%
SiO2_APTES_12067.168.268.668.910.0%11.8%12.4%12.8%
SiO2_64066.967.967.568.79.6%11.2%10.7%12.6%
Al2O3_12068.768.968.068.012.5%13.0%11.4%11.4%
Al2O3_18070.170.369.769.514.8%15.2%14.2%13.9%
Al2O3_3566.468.268.068.48.8%11.8%11.5%12.1%
SiO2_38068.669.469.769.412.4%13.7%14.3%13.7%
SiO2_20066.067.067.167.08.1%9.8%10.0%9.8%
TiO2_6565.865.566.666.27.8%7.3%9.2%8.5%
SG61.061.062.961.70.0%0.0%3.1%1.1%
60SiO2_12067.568.568.868.410.6%12.3%12.7%12.2%
SiO2_APTES_12066.465.767.565.78.8%7.7%10.6%7.7%
SiO2_64067.266.969.566.710.2%9.6%13.9%9.3%
Al2O3_12066.465.867.965.68.8%7.8%11.3%7.4%
Al2O3_18064.663.865.863.85.8%4.6%7.8%4.6%
Al2O3_3563.363.365.263.03.7%3.7%6.9%3.2%
SiO2_38065.265.367.365.46.9%7.0%10.3%7.2%
SiO2_20065.465.267.465.17.2%6.9%10.5%6.6%
TiO2_6563.764.165.965.24.4%5.0%8.0%6.9%
SG61.061.262.760.90.0%0.3%2.7%−0.2%
Table 5. Viscosity measurements of the SG-based nanofluid prepared by method III at 7.3 s−1 and 30 °C after heating the samples for 0, 7, 14, and 21 days at 30 °C and 60 °C.
Table 5. Viscosity measurements of the SG-based nanofluid prepared by method III at 7.3 s−1 and 30 °C after heating the samples for 0, 7, 14, and 21 days at 30 °C and 60 °C.
Heating Temperature (°C)SampleViscosity, cPViscosity Changes of the SG Solution
071421071421
30SiO2_12062.166.965.764.31.8%9.7%7.7%5.4%
SiO2_APTES_12063.967.966.667.34.7%11.3%9.1%10.3%
SiO2_64064.267.466.166.35.2%10.4%8.3%8.7%
Al2O3_12063.467.467.067.03.9%10.5%9.7%9.7%
Al2O3_18063.768.367.366.84.4%12.0%10.4%9.5%
Al2O3_3564.766.765.966.86.0%9.3%8.1%9.5%
SiO2_38064.867.867.668.86.2%11.2%10.8%12.7%
SiO2_20063.664.662.463.04.3%5.9%2.3%3.3%
TiO2_6562.665.463.465.92.6%7.1%3.9%8.0%
SG61.061.062.961.70.0%0.0%3.1%1.1%
60SiO2_12065.263.863.263.36.9%4.6%3.6%3.7%
SiO2_APTES_12064.963.963.964.06.3%4.7%4.7%4.9%
SiO2_64064.364.363.862.85.4%5.3%4.5%3.0%
Al2O3_12063.663.963.862.74.2%4.7%4.6%2.7%
Al2O3_18063.164.364.162.73.3%5.3%5.1%2.7%
Al2O3_3563.864.864.463.44.6%6.2%5.5%4.0%
SiO2_38063.361.564.062.03.7%0.8%4.9%1.6%
SiO2_20063.662.263.562.74.3%1.9%4.1%2.8%
TiO2_6562.462.963.963.12.2%3.1%4.7%3.5%
SG61.061.662.760.90.0%1.0%2.7%−0.2%
Table 6. Viscosity measurements of the SG-based nanofluid prepared by method IV at 7.3 s−1 and 30 °C after heating the samples for 0, 7, 14, and 21 days at 30 °C and 60 °C.
Table 6. Viscosity measurements of the SG-based nanofluid prepared by method IV at 7.3 s−1 and 30 °C after heating the samples for 0, 7, 14, and 21 days at 30 °C and 60 °C.
Heating Temperature (°C)SampleViscosity, cPViscosity Changes of the SG Solution
071421071421
30SiO2_12065.767.665.265.97.6%10.8%6.9%7.9%
SiO2_APTES_12063.064.264.064.53.3%5.1%4.9%5.7%
SiO2_64065.164.665.364.66.6%5.9%6.9%5.9%
Al2O3_12064.365.066.566.55.3%6.5%9.0%9.0%
Al2O3_18062.764.464.765.42.7%5.5%6.0%7.2%
Al2O3_3564.464.665.965.05.5%5.8%8.1%6.5%
SiO2_38065.064.867.664.76.5%6.2%10.8%6.0%
SiO2_20065.466.062.465.67.2%8.1%2.3%7.5%
TiO2_6561.463.563.463.70.7%4.1%3.9%4.4%
SG61.061.062.961.70.0%0.0%3.1%1.1%
60SiO2_12063.563.062.161.44.1%3.3%1.8%0.6%
SiO2_APTES_12062.660.960.460.12.6%−0.3%−1.0%−1.5%
SiO2_64064.162.162.061.05.0%1.8%1.5%0.0%
Al2O3_12063.162.863.963.23.4%2.9%4.8%3.5%
Al2O3_18063.263.163.563.23.6%3.4%4.0%3.5%
Al2O3_3562.461.962.062.42.2%1.5%1.6%2.3%
SiO2_38063.161.562.261.73.5%0.8%2.0%1.1%
SiO2_20061.362.160.360.60.4%1.7%−1.2%−0.8%
TiO2_6561.162.161.562.90.2%1.7%0.8%3.0%
SG61.060.862.760.90.0%−0.4%2.7%−0.2%
Table 7. Turbidity of the SG-based nanofluid a 30 °C.
Table 7. Turbidity of the SG-based nanofluid a 30 °C.
MethodSampleTurbidity
0 min after Preparation21 Days after Preparation
ISiO2_12010.010.6
SiO2_APTES_1209.28.9
SiO2_64012.77.9
Al2O3_12038.137.2
Al2O3_18062.045.2
Al2O3_35160.0133.0
SiO2_3804.03.6
SiO2_2007.05.8
TiO2_65140.0107.0
SG2.62.7
IISiO2_12013.010.6
SiO2_APTES_1209.07.2
SiO2_64010.06.3
Al2O3_1205.03.7
Al2O3_1807.06.1
Al2O3_357.05.1
SiO2_3802.02.4
SiO2_2003.03.1
TiO2_654.03.4
SG2.62.7
IIISiO2_12013.112.3
SiO2_APTES_12011.410.9
SiO2_6407.06.5
Al2O3_12081.282.8
Al2O3_18084.380.0
Al2O3_35144.0130.0
SiO2_3806.06.0
SiO2_2006.87.7
TiO2_65516.0502.0
SG2.62.7
IVSiO2_12011.85.7
SiO2_APTES_1209.54.6
SiO2_6407.95.5
Al2O3_12058.448.1
Al2O3_18059.144.2
Al2O3_35137.0111.0
SiO2_3804.04.8
SiO2_2006.06.2
TiO2_65593.0350.0
SG2.62.7
Table 8. Experiments and response variable.
Table 8. Experiments and response variable.
ExperimentPreparation MethodStanding TimeHeating TemperatureNanoparticle TypeViscosity, cP
1I−1−1SiO2_12064.5
2II−1−1SiO2_12065.7
3I1−1SiO2_12066.3
4II1−1SiO2_12067.4
5I−11SiO2_12065.9
6II−11SiO2_12067.5
7I11SiO2_12064.5
8II11SiO2_12067.1
9I−1−1SiO2_12063.8
10II−1−1SiO2_APTES_12067.5
11I1−1SiO2_APTES_12067
12II1−1SiO2_APTES_12067.6
13I−11SiO2_APTES_12065.9
14II−11SiO2_APTES_12067.5
15I11SiO2_APTES_12064.5
16II11SiO2_APTES_12068.4
17I−1−1SiO2_12066.3
18II−1−1SiO2_12065.8
19I1−1SiO2_12067.06
20II1−1SiO2_12068.5
21I−11SiO2_12065.9
22II−11SiO2_12066.4
23I11SiO2_12065.9
24II11SiO2_12066.2
25I−1−1SiO2_12063.6
26II−1−1SiO2_APTES_12067.1
27I1−1SiO2_APTES_12067.4
28II1−1SiO2_APTES_12068.9
29I−11SiO2_APTES_12065.9
30II−11SiO2_APTES_12066.4
31I11SiO2_APTES_12065.9
32II11SiO2_APTES_12065.7
Table 9. ANOVA results.
Table 9. ANOVA results.
ItemDegree of FreedomSum of SquaresMean SquareF-Valuep-Value
Preparation method117.02417.02415.8820.00046
Standing time15.0095.0094.6730.03967
Temperature10.7380.7380.6890.41391
Nanoparticle type10.1080.1080.1010.75324
Residual2728.941.072
Table 10. Carreau–Yasuda model parameters of SG and nanofluid SG + SiO2_120 at 30 °C.
Table 10. Carreau–Yasuda model parameters of SG and nanofluid SG + SiO2_120 at 30 °C.
ParameterSG SolutionNanofluid SG + SiO2_120
η 0 (cP)132.36149.22
λ (s)0.33580.4607
n 0.35300.3728
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Castro, R.H.; Corredor, L.M.; Llanos, S.; Causil, M.A.; Arias, A.; Pérez, E.; Quintero, H.I.; Romero Bohórquez, A.R.; Franco, C.A.; Cortés, F.B. Experimental Investigation of the Viscosity and Stability of Scleroglucan-Based Nanofluids for Enhanced Oil Recovery. Nanomaterials 2024, 14, 156. https://doi.org/10.3390/nano14020156

AMA Style

Castro RH, Corredor LM, Llanos S, Causil MA, Arias A, Pérez E, Quintero HI, Romero Bohórquez AR, Franco CA, Cortés FB. Experimental Investigation of the Viscosity and Stability of Scleroglucan-Based Nanofluids for Enhanced Oil Recovery. Nanomaterials. 2024; 14(2):156. https://doi.org/10.3390/nano14020156

Chicago/Turabian Style

Castro, Rubén H., Laura M. Corredor, Sebastián Llanos, María A. Causil, Adriana Arias, Eduar Pérez, Henderson I. Quintero, Arnold R. Romero Bohórquez, Camilo A. Franco, and Farid B. Cortés. 2024. "Experimental Investigation of the Viscosity and Stability of Scleroglucan-Based Nanofluids for Enhanced Oil Recovery" Nanomaterials 14, no. 2: 156. https://doi.org/10.3390/nano14020156

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop