Rheological Behavior of Surface Modified Silica Nanoparticles

Sep 20, 2018 - Polymer solutions are designed to develop a favorable mobility ratio between the injected polymer solution and the oil–water bank bei...
0 downloads 0 Views 4MB Size
Subscriber access provided by University of South Dakota

Fossil Fuels

Rheological behavior of surface modified silica nanoparticles dispersed in Partially Hydrolyzed Polyacrylamide and Xanthan Gum solutions: Experimental measurements, mechanistic understanding, and model development Laura M. Corredor-Rojas, Abdolhossein Hemmati Sarapardeh, Maen M. Husein, Prof. Mingzhe Dong, and Brij B. Maini Energy Fuels, Just Accepted Manuscript • DOI: 10.1021/acs.energyfuels.8b02658 • Publication Date (Web): 20 Sep 2018 Downloaded from http://pubs.acs.org on September 22, 2018

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 30 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Energy & Fuels

Carboxylic acid

O R

Modified silica NP

Silica +

OH

+

+H

R

C

Si

-OH

R

C OH

OH

OH

OH

+

H

Silica

O

+

R

OH

R

R

OH2

O

Si

-H

OH

-H2 O

O

+

+

O Si

Si

Modified silica NP R1 CH2 CH

Silane

R1 Silica surface H

O Si HO

OH

O

H

H

H

O

O

O

Si

Si

Si OH

O

OH

OH

2

OR

+ R O Si OR H2C OH

R1

n

Si O HO

OH

O

O

ACS Paragon Plus Environment

Si

Si

Si OH

O

O

O

Si

n

+ 3ROH

n

Si

O

OH

O

OH

OH

O Si

Energy & Fuels 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

1

Rheological behavior of surface modified silica nanoparticles

2

dispersed in Partially Hydrolyzed Polyacrylamide and Xanthan

3

Gum solutions: Experimental measurements, mechanistic

4

understanding, and model development

5 6

Laura M. Corredor-Rojasa, Abdolhossein Hemmati-Sarapardehb,*, Maen M. Huseina,*, Mingzhe Donga, Brij B. Mainia

7 8

a

9

b

Department of Chemical and Petroleum Engineering, University of Calgary, 2500 University Drive Northwest, Calgary, Alberta T2N 1N4, Canada Department of Petroleum Engineering, Shahid Bahonar University of Kerman, Kerman, Iran

10

ABSTRACT

11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30

Polymer solutions are designed to develop a favorable mobility ratio between the injected polymer solution and the oil-water bank being displaced by the polymer. Subsequently, a more uniform volumetric sweep of the reservoir is produced. Chemical and mechanical degradation of the polymer solutions, on the other hand, reduce their viscosity which significantly affects their performance. The primary objective of this study is to investigate the effect of surface modification of silica nanoparticles (NPs) on the effective viscosity of partially hydrolyzed polyacrylamide (HPAM) and xanthan gum (XG) solutions at different NP concentrations and temperatures. The chemical functionalization of SiO2 NPs with carboxylic acids and silanes was confirmed by FTIR measurements. The experimental results showed that the addition of SiO2 NP increased the viscosity of XG solutions due to the formation of three-dimensional structures between the silica NPs and the polymeric chains. The thickening effect of HPAM was improved by the addition of silica NPs modified with 3-(methacryloyloxy)propyl] trimethoxy silane (MPS), octyl triethoxy silane (OTES), and oleic acid-method A (OAA). In addition, the HPAM and XG nanopolymer sols of modified silica NPs showed more temperature and brine tolerance than that of unmodified silica NPs. A model was developed based on multilayer perceptron (MLP) neural network for predicting viscosity of nanopolymer sols using 9900 data points. The MLP model was trained by Bayesian Regularization (BR), Levenberg-Marquardt (LM), Resilient Backpropagation (RB), and Scaled conjugate gradient (SCG) algorithms. The results revealed that the BR-MLP model outperformed the three other models and could predict all the viscosity data with an average absolute relative error of 2.46% and R2 of 0.999.

31 32

Keywords: surface modified, nanoparticle, HPAM, XG, viscosity, polymer flooding, MLP

33

*Corresponding authors: A. Hemmati-Sarapardeh ([email protected]) and M. Husein

34

([email protected]) 1 ACS Paragon Plus Environment

Page 2 of 30

Page 3 of 30 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Energy & Fuels

35

1. Introduction

36

Polymer flooding results from adding a high-molecular-weight water-soluble polymer to the

37

injected water in a waterflood to reduce the mobility ratio by increasing water viscosity and

38

reducing the formation permeability.1 The reduced mobility ratio, defined as the ratio of the

39

displacing fluid mobility to the displaced fluid mobility, results in an increase in the volumetric

40

sweep efficiency compared to water flooding. This effect is particularly significant in highly

41

heterogeneous reservoirs and reservoirs with poor vertical sweep efficiency due to gravity. The

42

viscosity-increasing feature of polymers is derived from the repulsion between polymer

43

molecules and between the segments of the same molecule.

44

The two types of polymers primarily used in polymer flooding are synthetic polymers and

45

biopolymers.2 HPAM is the most-used synthetic polymer to date. Its performance depends on its

46

molecular weight and its degree of hydrolysis. The typical molecular weight of HPAM used is

47

within the range of 2 - 20 × 106 Dalton and the degree of hydrolysis is within 25% - 35%. HPAM

48

solutions often suffers from viscosity loss due to conditions as high temperature, high pressure,

49

and presence of other chemical substances in the oil reservoirs.

50

XG is the most widely used biopolymer for polymer flooding, but its current use is low

51

compared to HPAM, in part due to its higher cost. The viscosity of the XG solutions is relatively

52

insensitive to the salinity of the brine solvent and mechanical shearing, but it is quite susceptible

53

to biological degradation.3

54

Recent studies have shown that the addition of silica NPs to polyacrylamide solutions increases

55

their mobility control ability, temperature tolerance, and salt-tolerance1,4,5 and consequently,

56

increases the oil recovery. Silica NPs have received more attention due to their well-defined

2 ACS Paragon Plus Environment

Energy & Fuels 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

57

ordered structure, high surface area, cost-effective production, and the ease of surface

58

modification.6

59

Maghzi et al.4 found that the addition of 0.1 wt.% of nanosilica in polymer solution increased

60

polyacrylamide (PAM) viscosity to 2-4 times, and the pseudoplasticity behavior of the solution

61

was improved only at low and medium shear rates. They concluded that it could be a reason for a

62

10% increasing oil recovery during flooding test by nanopolymer sols in comparison with

63

polymer flooding. Maurya and Mandal1 reported that PAM got adsorbed on the surface of silica

64

by virtue of hydrogen bonding and the silica particles acted as physical cross-linker between the

65

polymeric chains resulting in improvement of viscosity of the displacing fluid. The decrease in

66

viscosity in case of silica/PAM suspension was less than that in case of the PAM solution in

67

presence of salt. Also, the results indicated that the viscosity of both silica/PAM suspension and

68

PAM solution is strongly dependent on temperature. For both systems, the viscosity decreased as

69

the temperature increased. However, the viscosity of silica/PAM suspension was at least 1.5

70

times higher than that of PAM solution at all temperatures. Kennedy et al.7 studied the rheology

71

of XG, locust bean gum and mixed biopolymer gel with nanosilica. These results showed that

72

nanosilica changed the rheological behavior of XG, including enhanced viscosity and elasticity.

73

It is well known that the best performance of these nanopolymer sols can be obtained when the

74

silica NPs are uniformly dispersed in the polymer solution. Good extent dispersion can be

75

achieved by altering the silanols of the surface of silica NPs with organic functionality, either

76

physically or chemically by covalent bonding.6 The modification of silica particles leads to better

77

hydrophobic interactions with the polymer when compared to unmodified hydrophilic fillers.8

78

On the basis of chemical interaction, the silylation of hydroxyl groups using organoalkoxysilanes

79

is the most commonly used method for surface modification of silica particles .9–15 Other

3 ACS Paragon Plus Environment

Page 4 of 30

Page 5 of 30 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Energy & Fuels

80

modifiers have been proposed to avoid some drawbacks accompanying the silylation reaction as

81

some silanols groups of the coupling agent might remain in the product causing further

82

condensation reactions during the period of storage and usage of the formed nanopolymer sol.

83

Those modifiers include alcohols16, carboxylic acids17, and polymers. 6,18–26

84

All those surface modification techniques have been widely studied for producing

85

polymer/silica nanocomposites that have potential application as coatings, sensors, proton

86

exchange membranes, etc.27, but they have not been studied for enhanced oil recovery (EOR)

87

applications.

88

In this study, the silica NPs are treated by chemical grafting with carboxylic acids and silanes to

89

generate different interfacial interactions. To this end, two polymers are used and also eight NPs

90

are synthesized. The expected effects are: (i) hydrophobicity of the NPs is increased, facilitating

91

silica NPs/polymer miscibility and a more uniform dispersion of the NPs; (ii) the interaction

92

between the treated NPs and the polymer can be tailored by changing the species of the coupling

93

agents to improve the mobility and control the ability of polymer solutions for polymer flooding.

94

A model is developed based on multilayer perceptron (MLP) neural network for predicting

95

viscosity of nanopolymer sols (9900 data points). The MLP model is trained by Bayesian

96

Regularization (BR), Levenberg-Marquardt (LM), Resilient Backpropagation (RB), and Scaled

97

conjugate gradient (SCG) algorithms. Statistical and graphical error analyses are conducted to

98

evaluate the performance of the developed models.

99 100

101 102

1. Materials and Methods 2.1 Materials. Dodecanedioic acid (DDDA, 99%), stearic acid (C18H36O2, 95%), oleic acid (C18H34O2,

90%),

3-(Methacryloyloxy)propyl]

trimethoxy

4 ACS Paragon Plus Environment

silane

(MPS,

98%),

Energy & Fuels 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

103

Octyl(triethoxy)silane (OTES, ≥97.5%), sulfuric acid (H2SO4, 95-98%), ethanol (EtOH, 99%),

104

hydrochloric acid (ACS reagent, 37%), sodium dodecyl sulfate (SDS, ≥98.5%), formaldehyde

105

solution (37 wt.% in water, contains 10-15% of methanol as stabilizer), sodium chloride (NaCl ,

106

99%), fumed silica powder (0.007 µm) were all obtained from Sigma-Aldrich (USA). Acetic

107

acid (99.5%), cyclohexane (99.5%), and ammonium hydroxide (28-30 wt.% solutions of NH3 in

108

water) were obtained from Fisher Scientific. Xanthan gum (XG, MW>2.106 Dalton) was

109

obtained from MP Biomedicals and FloopamTM 3630s (HPAM, hydrolysis degree 30-35%, MW>

110

20.106 Dalton) was obtained from SNF Floerger, USA). All chemicals were used as received.

111

2.2

Synthesis. Synthesis of SiO2 NPs modified with dodecanedioic acid (DDDA). A mass of 4

112

g of SiO2 NPs was dispersed into 100 mL of acetic acid using an ultrasound bath for 1 h. Then, 1

113

g of DDDA and 0.1 mL of H2SO4 were added to the dispersion which was stirred at 150 rpm for

114

another hour at 70°C using a magnetic stirrer. Afterward, the acetic acid was recovered using a

115

rotary evaporator. The dispersion was neutralized with 5 mL of 28-30 wt.% aqueous ammonia

116

solution while agitation continued for 1 h. The mixture was centrifuged at 2500 rpm for 30 min.

117

The precipitate was washed three times with 20 mL of 1/1 EtOH/water (V/V) solution to remove

118

excess DDDA. The precipitate was dried in an oven at 90°C for 24 h to yield 2.02 g of modified

119

SiO2-DDDA powder.

120

Synthesis of SiO2 NPs modified with oleic acid. The surface functionalization of SiO2 NPs with

121

oleic acid was carried out by using two different procedures. For the first procedure, a mass of 3

122

g of SiO2 NPs was dispersed into 100 mL of cyclohexane using an ultrasound bath for 3 h. Then

123

1.5 mL of oleic acid was added to the dispersion which was stirred at 150 rpm for 1 h at 75°C.

124

The cyclohexane was recovered using a rotary evaporator. Then, 5 mL of 28-30 wt.% aqueous

125

ammonia solution was added to the solution. The solution was stirred for 1 h at 80 rpm and 45°C.

5 ACS Paragon Plus Environment

Page 6 of 30

Page 7 of 30 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Energy & Fuels

126

The mixture was centrifuged at 2500 rpm for 30 min. The precipitate was washed three times

127

with 25 mL of 1:1 EtOH/water (V/V) solution to remove the excess amount of oleic acid. The

128

precipitate was dried in an oven at 60°C for 24 h to yield 2.17 g of modified SiO2-OAA powder.

129

The second modification was performed following a procedure proposed by Mahdavian et al.19.

130

A mass of 3 g of SiO2 NPs was dispersed into 90 mL of deionized water (DI) using an ultrasound

131

bath for 3 h. Then 1.5 mL oleic acid was added into the dispersion which was stirred at 150 rpm

132

for 90 min at room temperature. Then, 5 mL of 25 wt.% aqueous ammonia solution was added

133

into the solution while agitation continued overnight. The dispersion was neutralized with 30

134

wt.% aqueous HCl solution. The mixture was centrifuged at 2500 rpm for 30 min. The

135

precipitate was washed three times with 15 mL of 1:1 EtOH/water (V/V) solution to remove the

136

excess amount of oleic acid. The precipitate was dried in an oven at 50°C for 24 h to yield 2.51 g

137

of modified SiO2-OAB.

138

Synthesis of SiO2 NPs modified with stearic acid. The surface functionalization of SiO2 NPs

139

with stearic acid was carried out using two different procedures. For the first procedure, 1 g of

140

stearic acid was dissolved in 100 mL of cyclohexane at 45°C under magnetic stirrer (80 rpm).

141

Then, 4 g of SiO2 NPs and 0.2 mL of H2SO4 were added to the solution and agitation continued

142

for another hour. Afterward, the temperature was increased to 75°C while agitation continued for

143

2 h. The cyclohexane was recovered using a rotary evaporator. The precipitate was dispersed

144

with 100 mL of 1/1 EtOH/water (V/V) solution and 5 mL of 28-30 wt.% aqueous ammonia

145

solution. The mixture was centrifuged at 2500 rpm for 30 min. The precipitate was washed three

146

times with 25 mL of 1:1 EtOH/water (V/V) solution to remove excess stearic acid. The

147

precipitate was dried in an oven at 90°C for 24 h to yield 3.14 g of modified SiO2-SAA powder.

6 ACS Paragon Plus Environment

Energy & Fuels 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

148 149

For the second procedure, the same steps were followed except that H2SO4 was not added to the dispersion. A mass of 3.28 g of modified SiO2-SAB powder was obtained.

150

Synthesis of SiO2 NPs modified with MPS and OTES. A mass of 4 g of SiO2 NPs was dispersed

151

into 80 mL of cyclohexane by using an ultrasonic bath for 1 h. Then, 1.6 mL of MPS or 2.09 mL

152

of OTES was added to the dispersion. The dispersion was stirred for 12 h at room temperature.

153

Then, it was centrifugated at 2500 rpm for 30 min. The precipitate was washed three times with

154

EtOH. The precipitate was dried in an oven at 70°C for 24 h to yield 3.44 g of modified SiO2-

155

MPS powder and 3.29 g of SiO2-OTES.

156

2.3 Characterization of the NP. A Fourier transform infrared spectrometer, FTIR (model

157

IRaffinity-1s, Shimadzu, Japan) was used to analyze the NPs. Each spectrum was recorded over

158

the range of 4000–400 cm−1. Dry KBr was used for running the background spectrum. Scanning

159

electron microscopy coupled with energy dispersive X-ray, SEM-EDX (Phenom proX scanning

160

electron microscope, ThermoFisher scientific, Canada) was used to identify the elemental

161

composition of the surface modified silica NPs.

162

2.4

Nanopolymer sols preparation. Three different concentrations of SiO2 NPs (0.5, 1.0 and

163

2 wt.%) were used to prepare the nanopolymer sols. At first, SiO2 NPs were dispersed in DI

164

water and ultrasonicated for 1 h. A concentration of 0.1 wt.% of sodium dodecyl sulfate (SDS)

165

was added and the dispersions were stirred for 30 min. To prepare the HPAM nanopolymer sols,

166

0.4 wt.% of HPAM were added to the dispersions which were gently stirred for 48 h. To prepare

167

XG nanopolymer sols, 0.4 wt.% of XG were added to the dispersions which were stirred for 1 h

168

at 45°C. Finally, 1.0 wt.% of NaCl was added to each sample.

169

2.5 Viscosity of the nanopolymer sols. The viscosities were measured at 25°C and 70°C on a

170

Thermo Scientific™ viscometer (HAAKE RotoVisco 1, USA). The viscosity measurements

7 ACS Paragon Plus Environment

Page 8 of 30

Page 9 of 30 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Energy & Fuels

171

were conducted at shear rate from 5.0 to 100 s−1. Each measurement was repeated at least three

172

times to confirm the reproducibility, with the uncertainty found to be in the order of ± 1 to 7% of

173

the reported value.

174 175

3. Model Development

176

3.1. Multilayer perceptron neural network

177

One of computational intelligence systems is artificial neural networks (ANNs), which are

178

inspired form biological nervous systems. ANNs are widely used to find complex relationships

179

among outputs and inputs of a systems. Generally speaking, ANN has different elements

180

including weights, biases, nodes (neurons), transfer functions which help to find the unknown

181

relationships among the independent variables and the desirable output. Weights and biases are

182

interconnections which provide connections among processing elements, neurons

183

the proposed ANN models, Multilayer Perceptron (MLP) is recognized as the most commonly

184

used network in different areas such as petroleum engineering, chemical engineering, etc. This

185

type of network includes several layers in which the first layer is assigned to the inputs, while the

186

last layer is the desired output. The layers between the input and output layers are known as

187

hidden layers. Hidden layers establish the relationship between the output and inputs. 28–30

188

Each layer has a certain number of neurons; the number of neurons in the first layer is equal to

189

the number of input variables, where the last layer has only one neuron, which is the output of

190

the model. The number of neurons in the hidden layers are normally determined using a trial and

191

error procedure. Transfer functions are used in the neurons of hidden layers and output layer to

192

help the network providing more accurate results. Weights connect neurons in different layers to

8 ACS Paragon Plus Environment

28–30

. Among

Energy & Fuels 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 10 of 30

193

each other and each neuron in the hidden layers and output layer has also a bias. The value of a

194

neuron is passed through an activation/transfer function. In previous studies, we have completely

195

described mathematical formulation of the MLP networks. 31,32

196

The output of an MLP model having two hidden layers with logsig and purelin activation

197

functions for the two layers, respectively, and tansig for output layer is calculated as follows: Output = tan sig (w3 × ( purelin(w2 × (log sig ( x) + b1 )) + b2 ) + b3 )

(1)

198

In this formula, b1/b2/b2 are bias vectors of the first hidden layer/the second hidden layer/the

199

output layer. Furthermore, w1/w2/w3 are the weight matrices of the first hidden layer/the second

200

hidden layer/the output layer. Different methods can be used to train the MLP networks using

201

actual data. Normally, Levenberg-Marquardt (LM) is used for training such networks, while in

202

this study we used four different methods namely Scaled conjugate gradient (SCG), Bayesian

203

Regularization (BR), Resilient Backpropagation (RB), and Levenberg-Marquardt (LM) for

204

training neural networks, principle of which can be found in our previous publication32. The

205

proposed models based on these algorithms were called SCG-MLP, BR-MLP, RB-MLP, and

206

LM-MLP, respectively.

207

208

3.2 Preprocessing of the data for modeling

209

Each neural network model uses a set of inputs to calculate the desirable output. Selecting the

210

proper inputs for a model plays a key role in efficiency of the developed model. In fact, inputs

211

should cover all of affecting parameters on the output and also be independent. The aim of this

212

part is to find all of the affecting parameters/properties on the viscosity of nanopolymer sols.

213

Therefore, viscosity is regarded as the output of our proposed models. The affecting parameters 9 ACS Paragon Plus Environment

Page 11 of 30 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Energy & Fuels

214

on the viscosity include shear rate, type of NP, type of polymer, temperature, and concentration

215

of NPs. In this study, we only used a certain amount of SDS and NaCl in all experiments;

216

therefore, these parameters should not be selected as inputs as they are the same in all systems. It

217

should be noted the input of the model should quantitative rather than qualitative. To This end,

218

we assigned a number for the type of NP and polymer. For polymers, instead of using the name

219

of polymers as inputs, we used 1 for HPAM and 2 for XG. For NPs, we used 0 for SiO2

220

(unmodified), 3 for SiO2-SAB, 4 for SiO2-SAA, 5 for SiO2-OAB, 6 for SiO2-OAA, 7 for SiO2-

221

DDDA, 8 for SiO2-MPS, 9 for SiO2-OTES, and 10 for the systems without any NPs. In

222

summary, viscosity is regarded as a function of type of polymer, type of NP, concentration of NP

223

(%), temperature (Celsius), and shear rate (s-1).  =      ,    , . %   ,    , ℎ  2

224

4. Results and Discussion

225

3.1 FTIR and SEM-EDX results. Figures 1 and 2 show the FTIR spectra of unmodified and

226

modified silica NPs after normalization of the peak area. Unmodified silica shows peaks at 3745

227

and 3367 cm−1 which are assignable to (Si-OH) isolated groups and (Si–OH) stretching

228

vibration.33 The peaks at 1627 and 1130 cm−1 correspond to (O–H) bending vibration of

229

adsorbed molecular water and asymmetric stretching vibration of (Si-O-Si band). The peaks at

230

806 and 469 cm−1 are assigned to (Si-O-Si) symmetric bending vibration and (Si–O) rocking

231

vibration33, respectively.

232

The characteristic peaks of MPS are 2951 and 2481 cm−1 which are assigned to the asymmetric

233

and symmetric (C-H) stretching vibrations, and the peaks at 1703, 1456 and 1406 cm−1

234

correspond to the stretching vibration of (C=O), methylene (C–H) and vinyl (C–H) bending

235

vibration of MPS34, respectively. The identification of these peaks and the peaks at 970 and 1274 10 ACS Paragon Plus Environment

Energy & Fuels 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

236

cm-1 ascribed to Si-C bond35 in the IR spectrum of the modified SiO2 (Figure 1), proves the

237

bonding of MPS in the silica surface.

238

The characteristics peaks of OTES are 2926 and 2858 cm-1 which show very intense

239

asymmetric and symmetric (C-H) stretching vibrations. These bands are more intense than those

240

shown by MPS due to the large number of -CH2 groups in the octyl group36. These peaks are

241

identified in the IR spectrum of the SiO2-OTES NPs. Also, two more peaks are identified in this

242

spectrum. The peak at 1392 cm−1 which is assigned to the asymmetric deformation vibration of

243

(C-H) shown a slight substitution of the octyl groups in place of the -OH groups37 and the peak at

244

960 cm-1 which is ascribed to (Si-C) bond. The identification of all these peaks in the IR

245

spectrum of the modified SiO2 (Figure 1), proves the bonding of OTES in the silica surface.

246 247

Figure 1. IR spectrum of SiO2, SiO2-MPS, and SiO2-OTES.

248 249

The characteristic peaks of the carboxylic acids (DDDA, SA and OA) are ~2927 and ~2851

250

cm−1 which are assigned to the asymmetric and symmetric (C-H) stretching vibrations of -CH3 or

251

-CH2 groups, ~1690 cm-1 which correspond to the stretching vibration of (C=O), 1446 cm-1

252

which is ascribed to the symmetric stretching of (COO-)38, and 1400 cm-1 which is ascribed to 11 ACS Paragon Plus Environment

Page 12 of 30

Page 13 of 30 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Energy & Fuels

253

the (C–H) bending vibration. The identification of these peaks in the IR spectrums of SiO2-OA,

254

SiO2SA, and SiO2-DDDA (Figure 2), proves the bonding of the carboxylic acids in the silica

255

surface.

256 257 258

Figure 2. Infrared spectroscopy of (a) SiO2-DDDA, (b) SiO2-OAA and SiO2-OAB, and (c) SiO2-SAA and SiO2-SAB. 12 ACS Paragon Plus Environment

Energy & Fuels 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

259 260

The chemical modification of the silica surface with all these compounds is confirmed through SEM-EDX results presented in Table 1. Table 1. Elemental analysis per EDX results.

261

Sample SiO2 SiO2-SAA SiO2-SAB SiO2-OAA SiO2-OAB SiO2-DDDA SiO2-MPS SiO2-OTES

Concentration, wt.% O Si C 58.72 36.39 4.89 60.46 26.47 13.07 46.91 46.3 6.79 54.82 38.07 7.11 33.49 43.98 22.53 58.52 34.41 7.07 48.23 43.99 7.79 40.66 54.23 5.11

262

263 264 265 266

3.2 Effect of NP concentration and temperature on the viscosity of the HPAM nanopolymer sols. The results of HPAM solutions and HPAM nanopolymer sols are presented in Figure 3 and 4, respectively.

267

It is known that electrolytes tend to suppress the pseudo-plasticity of HPAM solutions.26,39–41

268

When HPAM is dissolved in water, the groups COO- of the molecular chain repel each other

269

causing its structure to remain extended. This feature gives the molecular chain a higher

270

hydrodynamic volume, increasing the viscosity of the polymer solution. When salt is added, the

271

cations of the salt neutralize the intrinsic electrical charges on the polymer particles. The charge

272

shielding effect causes the chains of the polymer to coil up into roughly spherical coils exposing

273

the minimum surface to the water which decreases the interactions between polymer chains.

274

Also, the polymer chains hydrate poorly so that the effective size of the swollen polymer

275

molecules decreases. Both effects cause the solution viscosity and its dependence on the shear

13 ACS Paragon Plus Environment

Page 14 of 30

Page 15 of 30 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Energy & Fuels

276

rate to decrease (Figure 3). On the other hand, the apparent viscosity of the HPAM solutions did

277

not significantly decrease with temperature.

278 279 280 281

Figure 3. Effect of salt and temperature on the viscosity of the HPAM solutions at constant polymer (0.4 wt.%), SDS (0.1 wt.%) and salt concentration (1 wt.%), whenever applicable.

282

It is observed from Figure 4 that the addition of untreated silica to the HPAM solutions has a

283

positive effect on the viscosity values at all concentrations. The same effect was observed for

284

silica modified with both silanes and OAA. The viscosity increase is attributed to adsorption of

285

the polymer molecules at the silica surface via hydrogen bonding between the oxygen or nitrogen

286

from HPAM and the hydrogen from the SiO2 surface (SiO-H ּ◌···N-H or SiO-H ּ◌···O-CNH2) or

287

between the hydrogen from the HPAM and the oxygen of the SiO2 surface (SiO···HNH-CO-C)42–

288

44

289

modifier molecules. In both cases, the silica particles act as physical cross-linker between the

290

polymeric chains.

or the hydrophobic interaction between the hydrocarbon backbone of the polymer and the

291

It has been demonstrated that despite the unfavorable electrostatic interactions between the

292

anionic polymer and the negatively charged silica particles, polymer adsorption does occur. 14 ACS Paragon Plus Environment

Energy & Fuels 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

293

Otsubo and Watanabe45 reported that during the adsorption process, polymeric chains can be

294

attached to more than one particle at a time and a number of polymeric chains can be attached to

295

the surface of one particle. Upon adsorption, only a portion of the polymer chain is in contact

296

with the NP surface (at single or several points), while the rest extend away from the silica

297

particle. This free chain can be adsorbed onto different silica particles, leading to the formation

298

of a three-dimensional network of flocs.46 The network structure is controlled by five parameters

299

that include polymer size, NP size, NP surface coverage, thickness of the adsorbed polymer

300

layer, and the interparticle repulsion range.47

301

When the silica particles concentration increases the viscosity of the nanopolymer sol increases

302

because of insufficient coverage of the particle surface with polymeric chains. This in turn

303

promotes the bridging of the polymer among more silica particles. This network is stable and not

304

easily broken by the irreversible adsorption of the polymer. Polymer adsorption is irreversible

305

because the polymeric chains may not be able to desorb simultaneously from all attached sites.

306

Desorption is determined by several factors, such as the charge density of the polyion, the nature

307

and charge of the surface, the concentration and molecular weight of the polymer, the salt

308

concentration in the solution, and the non-electrostatic interactions of the macromolecules with

309

each other and with the surface.43

310

15 ACS Paragon Plus Environment

Page 16 of 30

Page 17 of 30 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Energy & Fuels

T = 25°C

T = 70°C

a

b

c

d

e

f

311 312 313 314

Figure 4. Viscosity of the HPAM nanopolymer sols with (a-b) 0.5 wt.%, (c-d) 1.0 wt.%, and (ef) 2.0 wt.% NP concentration at 25°C and 70°C.

16 ACS Paragon Plus Environment

Energy & Fuels 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

315

The reduction in viscosity values caused by the addition of SAA, SAB, and OAB modified

316

silica NPs into the HPAM solutions suggests that the amount of polymer adsorbed at the silica

317

surface is high. It appears that the polymer molecule attaches to the particle at multiple points,

318

which promotes coiling of the polymer chain. Such adsorption at the silica surface reduces the

319

bridging of polymer with different silica particles and reduces the hydrodynamic size of the

320

polymer molecule which results in a lower viscosity than the HPAM solution. Sedeva et al.48

321

studied the adsorption of modified PAM on gold substrates with different degrees of

322

hydrophobicity and found that both the adsorbed amount of polymer and the adsorption rate

323

increase with the hydrophobicity of the gold substrates. The EDX analysis (Table 1) shows that

324

the silica NPs modified with SAA, SAB, and OAB are more hydrophobic than the silica NPs

325

modified with OAA and silanes. This suggests that polymer adsorption occurs via hydrophobic

326

interaction between the R-CH3 chain on the silica surface and the polymer backbone (-CH2-CH2-

327

).

328

Adsorption of the polymer molecules at the silica surface in case of the HPAM-DDDA

329

nanopolymer sols can occur via hydrogen bonding between carboxyl groups of the HPAM and

330

the COOH of the silica surface. The reduction in viscosity can be attributed to the low interaction

331

between NPs and polymer chains caused by the formation of the hydrogen bonding between the

332

carboxyl groups of the NPs. The increase of the concentration of the silica NPs modified with

333

SAA, SAB, DDDA, and OAB in HPAM solutions did not dramatically affect the viscosity,

334

contrary to what was expected. It suggests that the polymer adsorption mechanism is

335

predominant in the viscosity reduction.

336 337

17 ACS Paragon Plus Environment

Page 18 of 30

Page 19 of 30 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

338 339 340 341

Energy & Fuels

3.3 Effect of NP concentration and temperature on the viscosity of the XG nanopolymer sols. The results of XG solutions and XG nanopolymer sols are presented in Figures 5 and 6, respectively.

342 343 344 345

Figure 5. Effect of salt and temperature on the viscosity of the XG solutions at constant polymer (0.4 wt.%), SDS (0.1 wt.%) and salt concentration (1 wt.%), whenever applicable.

346

The XG polymer and nanopolymer sols exhibited non-Newtonian flow and shear-thinning

347

behavior due to the uncoiling and partial alignment of the polymer chains in the high shear rate

348

region (Figure 5). The addition of 1 wt.% salt made the viscosity higher than that of a salt-free

349

solution. This effect was previously studied by Wyatt and Liberatore.49 They reported that the

350

effect of salt concentration on the viscosity change is highly dependent on the polymer

351

concentration. For polymer concentrations below the critical concentration (CC~ 2000 ppm), the

352

viscosity dramatically decreases upon addition of salt. The addition of salt neutralizes the

353

charges on the XG molecules, therefore the elongated molecules (disordered conformation state)

354

are transformed into helix molecular conformation (ordered conformation state), which occupy 18 ACS Paragon Plus Environment

Energy & Fuels 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

355

smaller hydrodynamic volume. When the concentration is > CC, the interaction among the

356

intermolecular bonds increases, and its effect on solution viscosity is relatively more important

357

than the effect caused by the change in the hydrodynamic volume of the molecules.

358

As observed in Figure 5, XG solutions viscosity decreased at 70°C. The change in viscosity is

359

attributed to a conformation transition of the backbone from helix (ordered) to elongated

360

molecules (disordered) with progressive decrease of the rigidity of the (1-4)-β-D-glucan chain as

361

temperature increases.50

362

Addition of unmodified silica cannot mitigate the viscosity reduction when the temperature

363

increases from 25 to 70°C. However, the nanopolymer viscosity is higher than that of the XG

364

solution at low shear rates. The significant viscosity reduction can be attributed to the decrease in

365

the number of cross-linkers in the network formed by the NPs and the polymeric chains caused

366

by the reduction in the adsorbed amount of polymer onto the SiO2 surface. Also, the increase in

367

viscosity with the addition of silanes, SAA, OAA, OAB, and DDDA-modified silica can be

368

attributed to the crosslinking between the NPs and the polymeric chains through hydrogen

369

bonding or hydrophobic interaction. For silica modified with silanes, SAA, OAA, and OAB, the

370

crosslinking occurs via hydrophobic interaction between the R-CH3 chain on the silica surface

371

and the XG backbone. For XG-DDDA nanopolymer sols, the crosslinking can occur via

372

hydrogen bonding between the carboxylate groups on the trisaccharide side chains of the XG and

373

the COOH groups of the DDDA molecules attached to the silica surface. The viscosity reduction

374

caused by the increment of temperature from 25 to 70 °C is attributed to the weakening of the

375

hydrogen bonds and the hydrophobic interactions between the modified NPs and the XG chains.

376

19 ACS Paragon Plus Environment

Page 20 of 30

Page 21 of 30 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Energy & Fuels

T = 25°C

T = 70°C

a

b

c

d

e

f

377 378 379 380

Figure 6. Viscosity of the XG nanopolymer sols with a-b) 0.5 wt.%, c-d) 1.0 wt.%, and e-f) 2.0 wt.% of NP concentration at 25°C and 70°C.

20 ACS Paragon Plus Environment

Energy & Fuels 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

381

3.4. Evaluation of the viscosity models of the nanopolymer sols

382

An efficient and reliable model for accurate predicting the viscosity of the nanopolymer sols is

383

developed based on the experimental data. To this end, MLP neural network was selected which

384

has been widely used in many industries. The data points were randomly divided into two sets

385

including training (80%) and testing sets (20%). A four layer MLP model was designed to

386

accurately predict the large number of viscosity data. The first layer corresponds to inputs which

387

have five neurons and the last layer is assigned to output (viscosity) which has one neuron, as

388

stated earlier. The optimum architecture for the hidden layers was found to be 10 neurons for

389

both hidden layers, and tansig and logsig activation functions for the first and second hidden

390

layers, respectively. Purelin was selected as the activation function of the output layer.

391

For finding the optimum structure, different number of neurons in the first and second hidden

392

layers were considered and the networks were trained with Bayesian Regularization (BR),

393

Levenberg-Marquardt (LM), Resilient Backpropagation (RB), and Scaled Conjugate Gradient

394

(SCG) algorithms. The performance of each network depends, not only the number of neurons in

395

hidden layers, but also on the initial weights and biases which were randomly assigned to the

396

network. To find an appropriate number of neurons in the hidden layers, different number of

397

neurons in the hidden layers were tried and 10 neurons for both hidden layers were found a good

398

structure using all training algorithms after running the networks 20 times. Afterward, 10

399

neurons in each hidden layer were considered, and each model was trained a 100 times using

400

random initial weights and biases and the best model for each of the four algorithms, based on

401

statistical error analysis, were selected. Instruction for using the models is reported in the

402

Supplementary File. To evaluate the performance of each model, some statistical parameters;

403

including average absolute relative error (AARE%), average relative error (ARE%), standard

21 ACS Paragon Plus Environment

Page 22 of 30

Page 23 of 30 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Energy & Fuels

404

deviation (SD), root mean square error (RMSE), and R2 were used. Formulations and definitions

405

of these parameters can be found elsewhere32. The aforementioned statistical parameters for the

406

developed models are reported in Table 2 for training, testing, and whole data sets. As can be

407

seen in Table 2, all of the developed models have acceptable performance and could predict

408

viscosity of the nanopolymer sols with satisfactory accuracy in all sub sets. Nevertheless, BR-

409

MLP model has higher accuracy followed by LM-MLP, SCG-MLP, and RP-MLP. BR-MLP

410

model predicts all the data with an average absolute relative error of 2.47% and R2 of 0.9996. It

411

is interesting to note that the statistical parameters in training and testing sets are close to each

412

other, suggesting that overfitting was not an issue during model training. In addition, the average

413

relative error in all models is close to zero, illustrating no overestimation or underestimation in

414

the models.

415 416

Table 2. Statistical errors of the proposed models for predicting viscosity of nanopolymer sols Model

BR-MLP

LM-MLP

SCG-MLP

Set

ARE%

AARE%

RMSE

SD

R2

Train

-0.12

2.45

0.0029

0.0014

0.9996

Test

-0.20

2.52

0.0036

0.0015

0.9994

Total

-0.14

2.46

0.0031

0.0014

0.9996

Train

-0.22

3.99

0.0047

0.0046

0.9991

Test

-0.25

3.88

0.0051

0.0040

0.9987

Total

-0.23

3.96

0.0048

0.0045

0.9990

Train

-1.24

7.78

0.0147

0.0134

0.9916

Test

-0.98

7.78

0.0143

0.0141

0.9910

Total

-1.19

7.78

0.0146

0.0135

0.9915

22 ACS Paragon Plus Environment

Energy & Fuels

RP-MLP

Train

-1.28

10.37

0.0183

0.0237

0.9874

Test

-1.38

10.64

0.0177

0.0274

0.9844

Total

-1.30

10.43

0.0182

0.0244

0.9869

417

2.5

2.5

2

2 Predicted Viscosity, Pa.s

Predicted Viscosity, Pa.s

418

1.5

1 BR-MLP 0.5

1.5

1 LM-MLP 0.5

Y=X

0

Y=X

0 0

0.5 1 1.5 2 Experimental Viscosity, Pa.s

2.5

0

0.5 1 1.5 2 Experimental Viscosity, Pa.s

2.5

419 2.5

2.5

2

2

1.5

1 SCG-MLP 0.5

Predicted Viscosity, Pa.s

Predicted Viscosity, Pa.s

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 24 of 30

1 RP-MLP 0.5

Y=X

0

Y=X

0 0

420 421 422 423

1.5

0.5 1 1.5 2 Experimental Viscosity, Pa.s

2.5

0

0.5 1 1.5 2 Experimental Viscosity, Pa.s

2.5

Figure 7. Crossplot of the proposed MLP models for predicting the viscosity of the nanopolymer sols.

23 ACS Paragon Plus Environment

Page 25 of 30

424

The performance of the developed models was further checked graphically. For this purpose,

425

crossplot of all the models is depicted in Figure 7. In this figure, the predicted valued by the

426

proposed models were plotted against their corresponding experimental data. Figure 7 clearly

427

shows that the data points in all the models are close to the unit slope line, suggesting good

428

performance of the developed models. Nevertheless, the dispersion of the data around the unit

429

slope line is less in BR-MLP model followed by LM-MLP model. These graphical results are in

430

good agreement with statistical parameters, which attests to the superiority of the BR-MLP

431

model. It is noticeable that the RP-MLP model underestimates the viscosity data at high values.

432

To further graphically compare the performance of the developed models, the cumulative

433

frequency of the data versus their absolute relative errors was plotted in Figure 8. The proximity

434

of the model prediction to the vertical axis shows a higher model accuracy. Figure 8 again

435

confirms that the proposed BR-MLP model has the best performance among the developed

436

models and could predict 50% of the data with an absolute relative error of less than 1.35% and

437

80% of the data with an absolute relative error of less than 4.4%.

438 1

Cumulative Frequency

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Energy & Fuels

0.8

0.6 BR-MLP

0.4

LM-MLP 0.2

SCGMLP

0 0

5

10

15

20

Absolute Relative Error,%

439 440 441

Figure 8. Cumulative frequency versus absolute relative error of the proposed MLP models for predicting viscosity of nanopolymer sols 24 ACS Paragon Plus Environment

Energy & Fuels 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

442 443

Conclusions

444

The steady shear viscosity of HPAM nanopolymer sols exhibited shear-thinning behavior at

445

every NP concentration and temperature tested in this study. It was observed that the addition of

446

unmodified silica NPs and modified silica NPs (with both silanes and OAA) improved the

447

thickening behavior of the HPAM solution. However, silica NPs modified with DDDA, SAA,

448

SAB, and OAB resulted in reduction in the viscosity of the HPAM solution. The XG

449

nanopolymers sols, on the other hand, exhibited shear-thinning behavior over the entire range of

450

shear rate tested as well as enhanced viscosity with increased modified/unmodified NP

451

concentration. The increase in viscosity is most likely due to the formation of three-dimensional

452

structures between the silica NPs and the XG polymeric chains. Both HPAM and XG

453

nanopolymer sols of the modified silica NPs showed more temperature tolerance than the

454

unmodified silica NPs. As expected, HPAM solutions and nanopolymer sols were more sensitive

455

to salinity and temperature changes than the XG sols. The results of neural network modeling

456

confirmed that the developed BR-MLP model outperformed the other models. BR-MLP model

457

prediction displayed an average absolute relative error of 2.46% and R2 of 0.9996. The proposed

458

model can predict the viscosity of the nanopolymer sols at different temperatures, shear rates,

459

types of polymers, and type and concentration of the modified/unmodified NPs with high

460

accuracy.

461

462 463

ACKNOWLEDGMENT

25 ACS Paragon Plus Environment

Page 26 of 30

Page 27 of 30 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Energy & Fuels

464

Laura M Corredor-Rojas expresses her gratitude to Ecopetrol S.A for the awarded scholarship to

465

pursue her graduate studies. Special thanks to Amitabha Majumdar for his great help in the

466

experimental work.

467 468 469

NOMENCLATURE SiO2-DDDA

Silica modified with 1-12 dodecanedioic acid

SiO2-MPS

Silica modified with 3-(methacryloyloxy)propyl] trimethoxy silane

SiO2-OAA

Silica modified with oleic acid-Method A

SiO2-OAB

Silica modified with oleic acid-Method B

SiO2-OTES

Silica modified with octyl triethoxy silane

SiO2-SAA

Silica modified with stearic acid-Method A

SiO2-SAB

Silica modified with stearic acid-Method B

470 471 472

REFERENCES

473 474 475 476 477 478 479 480 481 482 483 484 485 486 487 488 489 490 491 492

1.

2. 3. 4.

5.

6.

7.

Maurya, N. K. & Mandal, A. Studies on behavior of suspension of silica nanoparticle in aqueous polyacrylamide solution for application in enhanced oil recovery. Pet. Sci. Technol. 34, 429–436 (2016). Sheng, J. J., Leonhardt, B. & Gmbh, W. H. Status of Polymer-Flooding Technology. (2015). Wei, B. Flow characteristics of three enhanced oil recovery polymers in porous media. J. Appl. Polym. Sci. 132, 1–7 (2015). Maghzi, A., Mohebbi, A., Kharrat, R. & Ghazanfari, M. H. An experimental investigation of silica nanoparticles effect on the rheological behavior of polyacrylamide solution to enhance heavy oil recovery. Pet. Sci. Technol. 31, 500–508 (2013). Maghzi, A., Kharrat, R., Mohebbi, A. & Ghazanfari, M. H. The impact of silica nanoparticles on the performance of polymer solution in presence of salts in polymer flooding for heavy oil recovery. Fuel 123, 123–132 (2014). El-Hoshoudy, A. N., Desouky, S. E. M., Betiha, M. A. & Alsabagh, A. M. Use of 1-vinyl imidazole based surfmers for preparation of polyacrylamide-SiO2 nanocomposite through aza-Michael addition copolymerization reaction for rock wettability alteration. Fuel 170, 161–175 (2016). Kennedy, J., Kent,K., and Brown, J. Rheology of dispersions of xanthan gum, locust bean gum and mixed biopolymer gel with silicon dioxide nanoparticles. Mater. Sci. Eng. C 48, 347–353 (2015). 26 ACS Paragon Plus Environment

Energy & Fuels 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

493 494 495 496 497 498 499 500 501 502 503 504 505 506 507 508 509 510 511 512 513 514 515 516 517 518 519 520 521 522 523 524 525 526 527 528 529 530 531 532 533 534 535 536 537 538

8. 9. 10. 11. 12.

13.

14. 15.

16.

17.

18.

19.

20.

21.

22.

23.

Althues, H, Henle, J., Kaskel, S. Functional inorganic nanofillers for transparent polymers. Chem. Soc. Rev. 36, 1454–1465 (2007). Jesionowski, T. & Krysztafkiewicz, A. Influence of silane coupling agents on surface properties of precipitated silicas. Appl. Surf. Sci. 172, 18–32 (2001). Shang, X. Y., Zhu, Z. K., Yin, J. & Ma, X. D. Compatibility of soluble polyimide/silica hybrids induced by a coupling agent. Chem. Mater. 14, 71–77 (2002). Feng, Y. & Gordon, L. N. PMMA/silica nanocomposite studies: Synthesis and properties. J. Appl. Polym. Sci. 91, 3844–3850 (2004). Wang, H., Xu, P., Zhong, W., Shen, L. & Du, Q. Transparent poly(methyl methacrylate)/silica/zirconia nanocomposites with excellent thermal stabilities. Polym. Degrad. Stab. 87, 319–327 (2005). Guyard, A., Persello, J., Boisvert, J.P., Cabame, B. Relationship Between the Polymer/Silica Interaction and Properties of Silica Composite Materials. J. Polym. Sci. Part B Polym. Phys. 1134–1146 (2006). doi:10.1002/polb Wei, L., Hu, N. & Zhang, Y. Synthesis of polymer-mesoporous silica nanocomposites. Materials (Basel). 3, 4066–4079 (2010). Yoon, D. H., Jang, J. W. & Cheong, I. W. Synthesis of cationic polyacrylamide/silica nanocomposites from inverse emulsion polymerization and their flocculation property for papermaking. Colloids Surfaces A Physicochem. Eng. Asp. 411, 18–23 (2012). Fuji, M., Takei, T., Watanabe, T. & Chikazawa, M. Wettability of fine silica powder surfaces modified with several normal alcohols. Colloids Surfaces A Physicochem. Eng. Asp. 154, 13–24 (1999). Mallakpour, S. & Marefatpour, F. An effective and environmentally friendly method for surface modification of amorphous silica nanoparticles by biodegradable diacids derived from different amino acids. Synth. React. Inorganic, Met. Nano-Metal Chem. 45, 376–380 (2013). Ding, X., Zhao, J., Liu, Y., Zhang, H. & Wang, Z. Silica nanoparticles encapsulated by polystyrene via surface grafting and in situ emulsion polymerization. Mater. Lett. 58, 3126–3130 (2004). Mahdavian, A. R., Ashjari, M. & Makoo, A. B. Preparation of poly (styrene-methyl methacrylate)/SiO2 composite nanoparticles via emulsion polymerization. An investigation into the compatiblization. Eur. Polym. J. 43, 336–344 (2007). Yang-Chuan, K., Guang-Yao, W. & Yi, W. Preparation, morphology and properties of nanocomposites of polyacrylamide copolymers with monodisperse silica. Eur. Polym. J. 44, 2448–2457 (2008). Ponnapati, R. et al. Polymer-functionalized nanoparticles for improving waterflood sweep efficiency: Characterization and transport properties. Ind. Eng. Chem. Res. 50, 13030– 13036 (2011). Nguyen, Q., Bao-Phuong Huu Do1, SPE Student, Duy-Khanh Pham1, H.-A. N. & DuyQuang Pham Dao1, and B.-D. N. P. Evaluation on the EOR Potential Capacity of the Synthetized Composite Silica-Core / Polymer-Shell Nanoparticles Blended with Surfactant Systems Viscous for the HPHT Offshire Reservoir Conditions. SPE Pap. Present. SPE Int. Oilf. Nanotechnol. Conf. Noordwijk, Netherlands, 12–14 June 2012. 1– 10 (2012). Yin, X. Z., Tan, Y. Q., Song, Y. H. & Zheng, Q. Dispersion stability and rheological behavior of suspensions of polystyrene coated fumed silica particles in polystyrene 27 ACS Paragon Plus Environment

Page 28 of 30

Page 29 of 30 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

539 540 541 542 543 544 545 546 547 548 549 550 551 552 553 554 555 556 557 558 559 560 561 562 563 564 565 566 567 568 569 570 571 572 573 574 575 576 577 578 579 580 581 582 583 584

Energy & Fuels

24. 25.

26.

27. 28. 29.

30.

31.

32.

33.

34.

35. 36.

37. 38. 39. 40.

41.

solutions. Chinese J. Polym. Sci. (English Ed. 30, 26–35 (2012). Behzadi, A. & Mohammadi, A. Environmentally responsive surface-modified silica nanoparticles for enhanced oil recovery. J. Nanoparticle Res. 18, 266–285 (2016). Choi, S. K., Son, H. A., Kim, H. T. & Kim, J. W. Nanofluid Enhanced Oil Recovery Using Hydrophobically Associative Zwitterionic Polymer-Coated Silica Nanoparticles. Energy and Fuels 31, 7777–7782 (2017). 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. 70, 319–330 (2017). Mallakpour, S. & Madani, M. A review of current coupling agents for modification of metal oxide nanoparticles. Prog. Org. Coatings 86, 194–207 (2015). Mohaghegh, S. Virtual-Intelligence Applications in Petroleum Engineering: Part 3— Fuzzy Logic. J. Pet. Technol. 52, 82–87 (2000). Lashkarbolooki, M., Hezave, A. Z. & Ayatollahi, S. Artificial neural network as an applicable tool to predict the binary heat capacity of mixtures containing ionic liquids. Fluid Phase Equilib. 324, 102–107 (2012). Mohagheghian, E., Zafarian-Rigaki, H., Motamedi-Ghahfarrokhi, Y. & HemmatiSarapardeh, A. Using an artificial neural network to predict carbon dioxide compressibility factor at high pressure and temperature. Korean J. Chem. Eng. 32, 2087– 2096 (2015). Hemmati-Sarapardeh, A., Dabir, B., Ahmadi, M., Mohammadi, A. H. & Husein, M. M. Modelling asphaltene precipitation titration data: A committee of machines and a group method of data handling. Can. J. Chem. Eng. 9999, 1–11 (2018). Hemmati-Sarapardeh, A., Varamesh, A., Husein, M. M. & Karan, K. On the evaluation of the viscosity of nanofluid systems: Modeling and data assessment. Renew. Sustain. Energy Rev. 81, 313–329 (2018). Rahman, I. A., Vejayakumaran, P., Sipaut, C. S., Ismail, J. & Chee, C. K. Size-dependent physicochemical and optical properties of silica nanoparticles. Mater. Chem. Phys. 114, 328–332 (2009). Sondi, I., Fedynyshyn, T. H., Sinta, R. & Matijević, E. Encapsulation of nanosized silica by in situ polymerization of tert-butyl acrylate monomer. Langmuir 16, 9031–9034 (2000). Dean, J. A. Lange’s Handbook of Chemistry. (1999). Al-Oweini, R. & El-Rassy, H. Synthesis and characterization by FTIR spectroscopy of silica aerogels prepared using several Si(OR)4 and R’’Si(OR’)3 precursors. J. Mol. Struct. 919, 140–145 (2009). Moral, L. M. TESIS DOCTORAL : Development of functionalized aerogels for applications as catalyst and hydrides matrix. Ph.D.Thesis. Univ. Valladolid (2016). Oomens, J. & Steill, J. D. Free carboxylate stretching modes. J. Phys. Chem. A Lett. 112, 3281–3283 (2008). Mungan, N. Improved waterflooding through mobility control. Can. J. Chem. Eng. 49, 32–37 (1971). Zeyghami, M., Kharrat, R. & Ghazanfari, M. H. Investigation of the Applicability of Nano Silica Particles as a Thickening Additive for Polymer Solution Applied in EOR Processes. Energy Sources Part A 36, 1315–1324 (2014). Zhang, R., He, X., Cai, S. & Kun Liu. Rheology of diluted and semi-diluted partially 28 ACS Paragon Plus Environment

Energy & Fuels 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

585 586 587 588 589 590 591 592 593 594 595 596 597 598 599 600 601 602 603 604 605 606 607 608

42. 43.

44.

45. 46. 47. 48. 49. 50.

hydrolyzed polyacrylamide solutions under shear: Experimental studies. Petroleum 1–10 (2016). doi:10.1016/j.petlm.2016.08.001 Lafuma, F., Wong, K. & Cabane, B. Bridging of colloidal particles through adsorbed polymers. J. Colloids Sci. 143, 9–21 (1991). Samoshina, Y., Diaz, A., Becker, Y., Nylander, T. & Lindman, B. Adsorption of cationic, anionic and hydrophobically modified polyacrylamides on silica surfaces. Colloids Surfaces A Physicochem. Eng. Asp. 231, 195–205 (2003). Hu, Z., Haruna, M., Gao, H., Nourafkan, E. & Wen, D. Rheological Properties of Partially Hydrolyzed Polyacrylamide Seeded by Nanoparticles. Ind. Eng. Chem. Res. 56, 3456– 3463 (2017). Otsubo, Y. & Watanabe, K. Rheological studies on bridging flocculation. Colloids and Surfaces 50, 341–352 (1990). Otsubo, Y. Effect of particle size on the bridging structure and elastic properties of flocculated suspensions. J. Colloid Interface Sci. 153, 584–586 (1992). Kawaguchi, M. Effects of polymer adsorption and desorption on rheological properties of silica suspensions. Colloids Surfaces A Physicochem. Eng. Asp. 86, 283–289 (1994). Sedeva, I. G., Fornasiero, D., Ralston, J. & Beattie, D. A. The influence of surface hydrophobicity on polyacrylamide adsorption. Langmuir 25, 4514–4521 (2009). Wyatt, N. B. & Liberatore, M. W. The effect of counterion size and valency on the increase in viscosity in polyelectrolyte solutions. Soft Matter 6, 3346–3352 (2010). Milas, M. & Rinaudo, M. Conformational investigation on the bacterial polysaccharide xanthan. Carbohydr. Res. 76, 189–196 (1979).

29 ACS Paragon Plus Environment

Page 30 of 30