Subscriber access provided by Gothenburg University Library
Interfaces: Adsorption, Reactions, Films, Forces, Measurement Techniques, Charge Transfer, Electrochemistry, Electrocatalysis, Energy Production and Storage
A mechanistic study of wettability changes on calcite by molecules containing a polar hydroxyl functional group and non-polar benzene rings Sooyeon Kim, Maria C. Marcano, and Udo Becker Langmuir, Just Accepted Manuscript • DOI: 10.1021/acs.langmuir.8b03666 • Publication Date (Web): 25 Jan 2019 Downloaded from http://pubs.acs.org on January 28, 2019
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 34 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
Langmuir
1
A mechanistic study of wettability changes on calcite by molecules containing
2
a polar hydroxyl functional group and non-polar benzene rings
3
Sooyeon Kim, Maria C. Marcano, and Udo Becker*
4 5
Department of Earth and Environmental Sciences, University of Michigan, 3021 North University Building, 1100 North University Avenue, Ann Arbor, MI 48109-1005, USA
6
*Corresponding author Udo Becker. Tel.: +1 734 615-6894; email:
[email protected] 7
Abstract
8
Oil extraction efficiency strongly depends on the wettability status (oil vs. water-wet) of reservoir
9
rocks during oil recovery. Aromatic compounds with polar functional groups in crude oil have a significant
10
influence on binding hydrophobic molecules to mineral surfaces. Most of these compounds are in the
11
asphaltene fraction of crude oil. This study focuses on the hydroxyl functional group, an identified
12
functional group in asphaltenes, to understand how the interactions between hydroxyl groups in asphaltenes
13
and mineral surfaces begin. Phenol and 1-naphthol are used as asphaltene surrogates to model the simplest
14
version of asphaltenes. Adsorption of oil molecules on the calcite {1014} surface is described using static
15
quantum-mechanical Density Functional Theory (DFT) calculations and classical Molecular Dynamics
16
(MD) simulations. DFT calculations indicate that adsorption of phenol and 1-naphthol occurs preferentially
17
between their hydroxyl group and calcite step edges. 1-naphthol adsorbs more strongly than phenol, with
18
different adsorption geometries due to the larger hydrophobic part of 1-naphthol. MD simulations show
19
that phenol can behave as an agent to separate oil from water phase, and to bind oil phase to the calcite
20
surface in water/oil mixture. In the presence of phenol, partial separation of water/oil with an incomplete
21
lining of phenol at the water/oil boundary is observed after 0.2 ns. After 1 ns, perfect separation of water/oil
22
with a complete lining of phenol at the water/oil boundary is observed, and the calcite surface become oil-
23
wet. Phenol molecules enclose decane molecules at the water/decane boundary preventing water from
24
repelling decane molecules from the calcite surface, and facilitate further accumulation of hydrocarbons
25
near the surface rendering the surface oil-wet. This study indicates phenol and 1-naphthol being good
1 ACS Paragon Plus Environment
Langmuir 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
26
proxies for polar components in oil, and they can be used in designing further experiment to test pH, salinity,
27
and temperature dependence to improve oil recovery.
28
Introduction
29
Oil production can be divided into three phases: primary, secondary, and tertiary. The primary oil
30
recovery refers to the natural rising of oil due to the pressure differences between the oil reservoir and the
31
wellbore or the use of lift systems such as rod pumps. Only around 10% of oil can be produced by these
32
primary methods. Water or gas injection can improve oil production, which is called secondary oil recovery.
33
Water or gas is injected into oil reservoirs to sweep oil to wellbores; then about 20 to 40 percent of the oil
34
can be produced using this recovery. Changing the properties of oil to make it more conducive to extraction
35
by injection of heat, gas, or chemicals is tertiary oil recovery, which is commonly known as enhanced oil
36
recovery (EOR). Oil production from the oil reservoir will be increased up to 70% by EOR.1
37
Lowering oil viscosity, surface tension, and capillary pressure are the main principles of EOR. Oil
38
flow is enhanced due to the reduced oil viscosity and surface tension, and this improves oil recovery.
39
Capillary forces play an important role in preventing micron-sized reservoir rock pores from releasing oil;
40
therefore, a lower capillary pressure also improves oil recovery. Wettability, which refers to a characteristic
41
of a solid surface showing a preference to be in contact with one fluid over another2, is one of the major
42
factors that control capillary pressure in pore spaces of rocks.3 Furthermore, secondary oil recovery by
43
waterflooding is directly related to the wettability of oil reservoir. Wettability also affects many crucial
44
parameters for evaluation of oil reservoirs, such as the relative permeability or the saturation profiles of the
45
oil reservoir.2, 4 Thus, understanding the wettability alteration of the reservoir rock surface is a key factor
46
for improving the efficiency of oil recovery.
47
While a mineral surface is often referred to as water-wet (hydrophilic) or oil-wet (lipophilic),
48
surface wettability is instead a continuum between an absolute preference for water (i.e., water-wet) and an
49
absolute preference for oil (i.e., oil-wet). Therefore, it is typically complicated to describe the actual wetting
2 ACS Paragon Plus Environment
Page 2 of 34
Page 3 of 34 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
Langmuir
50
scenarios in oil production because any reservoir will likely exhibit variable degrees or types of wettability.
51
Despite this complexity, a suitable model to predict the wettability alteration of mineral surfaces is needed
52
to increase our understanding of wettability change and ultimately to control the wettability for the purpose
53
of increasing the efficiency of oil recovery.
54
There are multiple factors that affect the wettability of mineral surfaces, including temperature,
55
pressure, pH, composition of crude oil, connate fluid compositions, and reservoir mineralogy.5 What makes
56
the oil molecules bind onto the rock surface are believed to be polar compounds, which are mostly found
57
in the asphaltene fraction of crude oil. Asphaltene is one of the four solubility classes obtained by the SARA
58
fractionation technique (i.e., Saturates, Aromatics, Resins, and Asphaltenes), which separates crude oil
59
based on solubility and polarity. Asphaltenes are the heaviest and the most polar among the four classes.
60
Although the chemical composition of asphaltenes is not well-defined, they generally consist of polycyclic
61
aromatic rings (usually 4 to 10) containing heteroatoms, such as S and N, and trace metal elements, with
62
peripheral alkane chains. The polar compounds in the asphaltene fraction are surface-active substances and
63
considered to be wettability modifiers in oil itself. Still, the mechanism of the wettability change of reservoir
64
rocks through these polar compounds is not fully understood due to the molecular complexity of crude oil.
65
Asphaltenes, as well as various polar compounds possibly existing in crude oil, have been used to
66
investigate the wettability alteration of mineral surfaces. Kumar et al.6 observed that surface wettability
67
change is controlled by asphaltene adsorption on mineral surfaces such as mica. In addition, they observed
68
similar adhesion forces for minerals aged with asphaltene fraction to that of the whole crude oil. Buckley7
69
has shown that the mechanisms of interaction between asphaltene and rock surfaces differ depending on
70
the presence of water. Ionic and colloidal interactions were suggested as the two main interaction
71
mechanisms that contribute to wetting alteration by asphaltene in a reservoir rock.
72
Polar compounds in crude oil contain various heteroatoms such as oxygen, nitrogen, and sulfur.8 O
73
atoms in asphaltenes mostly exist in compounds containing of carboxylic acid, hydroxyl, carbonyl groups,
74
or ethers.9 Previous studies have shown that carboxylic acids render the carbonate surfaces oil-wet.10-14 3 ACS Paragon Plus Environment
Langmuir 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
75
Hydroxyl groups also appear to be good wettability modifiers as shown in early studies when calcite
76
surfaces were exposed to short carbon chains of alcohols such as ethanol, propanol, and pentanol.15-19
77
Hydrogen bonding of ethanol to the calcite surface is stronger than that of water, so ethanol is firmly bonded
78
onto calcite {1014} terraces and forming an ordered and stable adsorption layer15-16, which can be disrupted
79
at steps17. Some previous works20-21 proved that diols (ethylene glycol and triethylene glycol) also work as
80
wettability modifiers decreasing water-wetting of calcite with a different orientation toward the calcite
81
surface compared to ethanol. Van Duin et al.22 calculated adsorption differences of polar and non-polar
82
compounds onto the calcite and quartz surfaces. They showed that phenol, as a polar compound, could
83
penetrate through thin water films and adsorb onto the mineral surfaces from a water-phenol mixture, and
84
even at low concentrations, phenol influences surface wettability.
85
In this study, we chose phenol and 1-naphthol as model surface-active substances to simulate the
86
binding of a polar hydroxyl functional groups to water-wet mineral surfaces and change them to an oil-wet
87
state by providing non-polar molecular terminations. Phenol and 1-naphthol carry both a hydroxyl
88
functional group and benzene rings, so they represent simplified asphaltenes which carry the core
89
characteristics of asphaltenes. The additional benzene ring of 1-naphthol is to determine the effect of
90
aromatic rings on the adsorption of asphaltenes by comparison with phenol adsorption reaction. A non-polar
91
component, decane, was chosen to model a general oil phase. We calculated oil molecule adsorption on the
92
calcite {1014} surface using both static quantum-mechanical Density Functional Theory (DFT)
93
calculations and Molecular Dynamics (MD) simulations using classical force-fields. The goal was to
94
understand the molecular mechanisms of the wettability changes on calcite surfaces by hydroxyl functional
95
group of polar components in crude oil. The calcite surface was selected because it is one of the most
96
abundant minerals in oil reservoirs and has stable and well-defined cleavage planes which are easy to model
97
and explore.
98
Using DFT computational method, adsorption energies and geometries of two asphaltene
99
surrogates, phenol and 1-naphthol, were investigated. Classical MD simulations were performed to 4 ACS Paragon Plus Environment
Page 4 of 34
Page 5 of 34 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
Langmuir
100
investigate the molecular sorption behaviors of phenol molecules on calcite surfaces in several different oil
101
(mimicked by decane)-water mixture scenarios. To determine how calcite surface wettability changes
102
between water-wet and oil-wet by phenol, simulation cells with and without phenol were calculated.
103
Throughout these calculations, we aim to determine whether phenol is a good wettability modifier of the
104
calcite surface (i.e., from water-wet to oil-wet) and how it changes surface wettability, which would further
105
our understanding of wettability-change mechanisms on calcite surfaces. Density profiles and radial
106
distribution functions (RDFs) were calculated to obtain the density changes of water and oil molecules on
107
the calcite surfaces to get molecular level information about dynamical and structural properties of the
108
adsorbates. A new set of force field parameters was developed using quantum-mechanical results to
109
describe decane-calcite interactions for classical MD simulations.
110
Methods
111
Quantum-mechanical calculations
112
The adsorption energies and optimized geometries of two different asphaltene surrogates (phenol,
113
1-naphthol) on the calcite {1014} surface were examined using DFT calculations implemented in the
114
software package DMol3 (local Density functional calculations on Molecules)23. The GGA-PBE exchange-
115
correlation potential24 was used, and core electrons were treated using effective core potentials. Double
116
Numerical plus d-functions (DND) were used as a basis set for atomic orbitals; d-functions were added to
117
improve the modeling of polarizability of atoms. The COnductor-like Screening MOdel (COSMO) was
118
used to approximate the effect of hydration by replacing the water solvent by a continuous dielectric fluid.25
119
The calcite {1014} surface was constructed by cleaving a bulk crystal to obtain a slab of three
120
molecular layers with a 20 Å thick vacuum layer, resulting in a supercell with dimensions of 25.70 Å ×
121
25.70 Å × 27.63 Å. All atoms in the top two molecular layers of the calcite substrate were allowed to relax,
122
and the bottom molecular layer was fixed in its bulk position during the adsorption reaction of the phenol
123
and 1-naphthol.
5 ACS Paragon Plus Environment
Langmuir 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
124
Page 6 of 34
Adsorption energies (Ead) were calculated as,
125
Ead = Emolecule-calcite – Ecalcite – Emolecule
(1)
126
where Ead is the adsorption energy of a molecule (adsorbate) on the calcite surface, Emolecule-calcite is
127
the total energy of the calcite slab with the adsorbate, Ecalcite is the energy of the calcite slab, and Emolecule is
128
the energy of the adsorbate.
129
Charge density differences were also calculated using CASTEP26, which visualize how the
130
adsorption of a molecule affects the electron distribution relative to the isolated molecule and the
131
unperturbed calcite surface. The charge density differences were given by
132
∆ρ = ρmolecule-calcite – (ρcalcite + ρmolecule)
(2)
133
where ρmolecule-calcite is the electron density of the total system (an adsorbate and the calcite surface),
134
ρmolecule and ρcalcite are the unperturbed electron densities of the adsorbate and the calcite surface, respectively.
135
Potential model for classical MD and the simulation details
136
Classical MD simulations were used to examine the dynamic behavior of the polar-nonpolar
137
character and its effects on the calcite surface wettability change. The TIP3P27 force field was used to
138
represent water molecules as a three-point model. For all organic molecules that represent our model oil
139
and asphaltene surrogate substances, the OPLS-AA (Optimized Potentials for Liquid Simulations-All
140
Atoms) force field28 was used. The force field for calcite is adopted from the potential by Pavese et al29.
141
Although this potential was fitted to the bulk properties of calcite, it is generally possible to use bulk fitted
142
potential sets to surface calculations in case of ionic materials.30 According to de Leeuw and Parker31, the
143
surface energy of {1014} face calculated using the potential set by Pavese et al. is 0.6 J/m2 in vacuum and
144
0.3 J/m2 when hydrated. To increase computational efficiency, polarization shells of the carbonate oxygen
145
were removed from the potential by Pavese et al., such that the O atoms have a charge of –1.045 e. After
146
removing polarization shells of the O atoms, the Buckingham potential parameters of both Ca-O and O-O
6 ACS Paragon Plus Environment
Page 7 of 34 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
Langmuir
147
interactions were modified to fit the lattice parameters of calcite to the experimental values32 (Table S1 and
148
Table S2). The surface energy of the {1014} face by this modified potential set is 0.56 J/m2.
149
While TIP3P, OPLS-AA, and the modified Pavese potential set are good representations of water,
150
organic molecules, and calcite, respectively, it is particularly important to have proper interatomic
151
potentials between mineral, water, and organic molecules. Reliable potential estimates are crucial in order
152
to properly describe immiscibility (e.g., of the water and oil phase) and competition of adsorption between
153
water and organic molecules on the calcite surface. Between water and organic molecules, the geometric
154
mixing rule was used to calculate interatomic forces. However, due to the differences of the bond character
155
(the organic molecules are described with a covalent model, while the calcite structure is described using
156
an ionic model), the method proposed by Freeman et al.33 was used to calculate interactions of calcite-water
157
and calcite-organics (Table S1). This method is generally in good agreement for polar organic molecules.
158
However, the interaction of calcite and non-polar molecules needs some modification to supplement Ca-
159
organic interactions. Thus, decane-calcite interatomic potentials were derived by reproducing our DFT
160
results (in Section 2.3).
161
The calcite and water-oil mixture model were prepared using Materials Studio 7.0 (Accelrys)34.
162
The calcite slab was cleaved along the {1014} face resulting in a 40.5 Å × 39.9 Å × 22.8 Å slab with a
163
60 Å thick vacuum layer in between the calcite slabs. A calcite slab of eight molecular layers was first
164
energy minimized and then relaxed using an NPT ensemble at 300 K and 1 atm. Then, the position of the
165
calcite slab was fixed in its relaxed geometry during subsequent MD calculations. A water-decane mixture
166
solvent (1:1 volume ratio) was added into the initial vacuum layer with an appropriate number of molecules
167
to obtain the density of the respective solvent at ambient conditions. In order to investigate the role of
168
hydroxyl functional group on the wettability alteration of the calcite surface, phenol molecules were added
169
into the solvent layer. The resulting solution had a 1:1 molecular number ratio of decane and phenol in its
170
oil phase.
7 ACS Paragon Plus Environment
Langmuir 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
171
Classical molecular-dynamics simulations were performed using the LAMMPS package35 for (i)
172
the water-decane random mixture and (ii) the water-decane-phenol random mixture, either on mixed-wet
173
or oil-wet calcite surfaces for both cases. For mixed-wet surfaces, a random mixture of oil/water was
174
adsorbed on calcite surfaces, showing more oil-wet conditions where phenol or decane is adsorbed and
175
more water-wet ones where no phenol is adsorbed. In contrast, only oil molecules (includes both decane
176
and phenol) were adsorbed for oil-wet calcite surfaces at the beginning. All calculations were carried out
177
at 300 K using a Nosé-Hoover thermostat in the NVT ensemble. The “sandwiching” a layer of the liquid
178
mixture between the two relaxed calcite surface is more of a computational necessity rather than the intent
179
to model oil behavior in nano-pore spaces. Since the actual pores from which oil can be extracted are
180
typically more than a few nm wide, ambient conditions reflect more natural conditions, unless one would
181
want to model wettability under different pressure conditions which is beyond the scope of this study. A
182
cutoff of 10.1 Å was applied for all short-range potentials, and, due to the periodicity of the cell, Ewald
183
summation was used for Coulomb interactions. The total simulation time was 1 ns with a time step of 1 fs
184
(e.g., one million time steps), and the results were visualized using VMD (Visual Molecular Dynamics)36.
185
Derivation of calcite-decane interatomic potentials for classical MD
186
The potential parameters to describe calcite-decane interactions were calculated by fitting the
187
energy curve using LAMMPS to the result of DFT calculations. A calcite slab made of four molecular
188
layers was used for this calculation. The general method by Freeman et al.33 was used to estimate the
189
mineral-organic interatomic potential. To adjust the Ca(calcite)-C(decane) interactions, Buckingham
190
potentials plus Coulomb terms were added to the potential initially calculated by the Freeman et al. method.
191
Using this potential (Table S3), one decane molecule was put onto the calcite surface, and energy
192
minimization was performed. After energy minimization, adsorption energies were calculated at different
193
distances from the calcite surface. The adsorption energies were calculated using Equation (1).
194
All geometries at various distances from the calcite surface were applied for DFT energy
195
calculations in order to calibrate the potentials used for classical-mechanics force fields (as used with 8 ACS Paragon Plus Environment
Page 8 of 34
Page 9 of 34 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
Langmuir
196
LAMMPS). Single point energy calculations were performed for each distance using the GGA-PBE
197
exchange-correlation potential and the effective core potential for core electrons treatment. The resulting
198
energy curves of DFT and classical-mechanics force fields (FF) show some differences between the two
199
energy-well minima, but the Ca-C1 distance difference is only 0.11 Å (Figure S1). Dispersion correction
200
can, in general, be used to correct van der Waals interactions, which can improve the accuracy of adsorption
201
energy calculations of organic molecules. We have compared our energy versus distance curves against
202
conventional DFT, as described above, DFT-D using TS (Tkatchenko-Scheffler) scheme37, and finally the
203
Grimme approach38. Using those data, our energy vs. distance curve fall within the range of those
204
corrections schemes. The potential parameters derived here were used throughout this study to describe
205
calcite-decane interactions.
206 207 208 209
Results and Discussion Quantum-mechanical calculations to determine the adsorption energies and optimized structures of organic-mineral interactions
210
The adsorption energies of two asphaltene surrogates, phenol and 1-naphthol, were calculated on
211
the calcite {1014} surface, and their optimized geometries were analyzed. Since the pKa values of phenol
212
and 1-naphthol are 9.95 and 9.34, respectively, protonated forms of these molecules are expected in most
213
environments. Thus, throughout this study, the protonated form of phenol and 1-naphthol are used. Table 1
214
shows the adsorption energies of phenol and 1-naphthol in vacuum, water solvent, and n-decane solvent.
215
More negative values of adsorption energies indicate more favorable adsorption. As shown in Table 1(a),
216
for the calcite surface terrace sites, the adsorption energy of 1-naphthol (-1.72 eV (~ -66.9 kBT [J] with
217
T=298 K)) is stronger than that of phenol (-1.30 eV (~ -50.6 kBT [J] with T=298 K)) in vacuum. This trend
218
is consistent with the results in water and decane.
9 ACS Paragon Plus Environment
Langmuir 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 34
219
Table 1. Adsorption energies (in eV per molecule) of phenol and 1-naphthol to the calcite {1014} surface in vacuum, water, and decane: on flat terrace, and step edges. surface site (a) Terrace (b) Obtuse step (c) Acute step
substances phenol 1-naphthol phenol 1-naphthol phenol 1-naphthol
Ead (in vacuum) [eV] -1.30 -1.72 -1.40 -1.70 -1.62 -1.91
Ead (in water) [eV] -0.79 -1.04 -0.65 -0.82 -0.89 -0.93
Ead (in decane) [eV] -1.10 -1.46 -1.12 -1.41 -1.32 -1.53
Figure 1. Optimized geometries of two asphaltene surrogate substances with a hydroxyl functional group on the flat terrace of calcite {1014} surface. Two adsorbates are represented in ball-and-stick model, and the calcite slab are displayed using bond lines for clarity. (Ca-green, O-red, C- grey, H-white)
220 221
For both phenol and 1-naphthol, the distances from the calcite surface to their hydroxyl groups are
222
quite close (Figure 1, Ccalcite-Ohydroxyl is 2.6 Å for phenol, 2.7 Å for 1-naphthol). Hydroxyl groups are 10 ACS Paragon Plus Environment
Page 11 of 34 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
Langmuir
223
positioned like an anchor while the benzene ring part of the molecules floats away from the surface to be
224
available to bond with other oil components. The phenol molecule has a tilted geometry on the calcite
225
surface with an angle of 35° interacting with the calcite surface mainly through its hydroxyl functional
226
group (Figure 1(a)). 1-naphthol lies at a lower angle (~20°) with respect to the surface, thus increasing the
227
area of calcite covered per adsorbed molecule (Figure 1(b)). As the size of the molecule increases, more
228
van der Waals interactions occur. Since the side group of 1-naphthol is twice the size of that of phenol, with
229
two benzene rings rather than one, 1-naphthol experiences more van der Waals interactions with the calcite
230
surface resulting in a flatter geometry. This is reflected by the stronger adsorption of 1-naphthol to the
231
calcite surface compared to phenol. Similar adsorption geometries were observed by Torres et al.39 for the
232
adsorption of a prototypical asphaltene with four aromatic rings on the α-quartz surface using DFT
233
calculations. The model asphaltene molecule showed preferred adsorption in a parallel position to the
234
calcite surface due to its strong interaction with the α-quartz surface by London dispersion forces associated
235
with its four aromatic rings.
236
In addition, hydroxyl functional groups form hydrogen bonds with the carbonates of the calcite
237
surface. For example, ethanol more strongly adsorbs to the calcite surface than water via Ca-Ohydroxyl and
238
OHhydroxyl-Ocarbonate interactions15-16; thus, ethanol could replace water molecules at the calcite surface in
239
water/ethanol mixture solution.17 Our calculated distances between H of the hydroxyl group and adjacent
240
O of calcite carbonate are 1.7 Å and 1.6 Å for phenol and 1-naphthol, respectively, which are small enough
241
to be considered as H bonds.
242
Since hydrogen bonding has some non-negligible covalent character, it confers some charge-
243
transfer from an electron donor to an electron acceptor. Thus, hydrogen bonding causes some atomic charge
244
differences for both electron donor and acceptor.40-41 From the atomic charge analysis of phenol and
245
1-naphthol by Mulliken population analysis42, there are charge differences for both hydroxyl O and H when
246
we compare the charge values of the two species before and after adsorption reactions. This change in
247
charge for hydroxyl H (+0.1 e for both phenol and 1-naphthol) further confirms the existence of hydrogen 11 ACS Paragon Plus Environment
Langmuir 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 12 of 34
248
bonding with the calcite surface. Since both O and H of the hydroxyl group make strong bonds with calcite
249
through Ca-Ohydroxyl electrostatic interaction and OHhydroxyl-Ocarbonate hydrogen bonding, phenol and
250
1-naphthol can strongly bond onto the calcite surface.
251
Charge density differences were examined for the optimized adsorption geometries of both phenol
252
and 1-naphthol on the calcite {1014} surface. In Figure 2, the electron density increased (red surface) for
253
hydroxyl O and decreased (blue surface) for calcite Ca. Likewise, the electron density increased for
254
carbonate O, while it decreased for hydroxyl H. Therefore, Ca-Ohydroxyl and Hhydroxyl-Ocarbonate are the major
255
contributors to the adsorption of the both phenol and 1-naphthol onto the calcite surface. Previous work43
256
reported similar charge density differences for water molecules on the calcite surface, showing strong bonds
257
for Ca-Owater and Ocarbonate-Hwater.
(a)
(b)
Z
Z Y
X
X Y
Figure 2. Charge density differences for (a) phenol, and (b) 1-naphthol. The red and blue colors indicate increase and depletion of the charge density, respectively. (Ca-green, O-red, C- grey, H-white)
258 259
Moreover, the aromatic rings of 1-naphthol have strong interactions with the calcite surface as
260
shown in Figure 2(b). This confirms the stronger interactions of 1-naphthol with the calcite surface, which
261
were already suggested by the flatter placement of the 1-naphthol benzene rings compared to that of phenol.
262
The charge density differences for a model asphaltene calculation by Alvim et al.44 also showed increased
12 ACS Paragon Plus Environment
Page 13 of 34 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
Langmuir
263
and decreased electron densities mainly at the aromatic rings. Their asphaltene model had eleven aromatic
264
rings and charge density changes were observed mostly between the aromatic rings and the calcite surface.
265
The influence of the calcite surface geometry on the adsorption of asphaltene surrogates were also
266
investigated. Adsorption of phenol and 1-naphthol to calcite step edges were energetically more favorable
267
than to flat terraces (Table 1). Calcite step edges were constructed by deleting the middle two rows of the
268
top molecular layer along [481] direction from the calcite model. Two layers from the top of the calcite
269
structure including the step features were free to relax and the bottom layer was fixed at its bulk state during
270
the calculation. When phenol was adsorbed to the calcite acute step edge (with an angle of 78° at its step
271
edge) in vacuum, the adsorption energy was 0.32 eV more negative (stronger) than the adsorption to the
272
terrace surface. 1-naphthol also showed a 0.19 eV more negative adsorption energy to the calcite acute step
273
edge than to the calcite flat terrace in vacuum. Comparing the two different step edges of the calcite surface,
274
the adsorption energy of phenol to the acute step edge is 0.22 eV more favorable than the one to the obtuse
275
step edge. 1-naphthol also shows 0.21 eV of favorable adsorption reaction to the acute step edge than the
276
obtuse one.
277
The optimized adsorption geometries of both species at the calcite acute step edges are similar. In
278
both cases, hydroxyl O is located near the Ca2+ at the step edges, and hydroxyl H forms a hydrogen bond
279
with carbonate O at the step edges as shown in Figure 3. Adsorption to the acute step edge is more favorable
280
than to the obtuse step edge due to the shorter distance (less than 3.0 Å) between the two adjacent Ca2+
281
atoms: one is at the step edge and the other one is in the 2nd layer of the calcite slab right underneath the
282
hydroxyl O (shown in Figure 3 with arrows). Van der Waals interactions between the calcite surface and
283
the benzene rings of both species also stabilize adsorbed structures of these molecules. Likewise, the
284
adsorption to the flat terrace, 1-naphthol lies flatter than phenol due to the stronger van der Waals
285
interactions. Since the contribution of Ca-Ohydroxyl electrostatic interaction is significant for the adsorption
286
of both substances, acute step edges with two adjacent Ca2+ result in stronger interactions compared to
287
obtuse step edges where there is only one Ca2+. 13 ACS Paragon Plus Environment
Langmuir 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
288
Considering the solvent effect for both surface terrace and step edge sites, adsorption reactions in
289
hydrophobic decane solvent were much more favorable for both phenol and 1-naphthol compared to the
290
adsorption reactions in hydrophilic water solvent (Table 1). Similar to the vacuum case, phenol and
291
naphthol adsorption to acute step edges were favored over obtuse step edges in water and decane solvents.
292
For all cases, the optimized geometries show that both phenol and 1-naphthol molecules interact with the
293
calcite slab mainly through their hydroxyl functional groups, leaving the hydrophobic parts toward solution
294
to interact with surrounding hydrophobic oil molecules.
295
Figure 3. Optimized geometries of phenol and 1-naphthol simulating the role of asphaltenes for subsequent bonding of oil to calcite {1014} surface with step edges along the [481] direction. (a) Phenol and (b) 1-naphthol at the acute step edge of the calcite surface. Two adsorbates and the first molecular layer of the calcite slab are represented by balland-stick model, and the second and the third layers of the calcite slab are displayed using bond lines for clarity. (Cagreen, O-red, C- grey, H-white)
14 ACS Paragon Plus Environment
Page 14 of 34
Page 15 of 34 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
296 297
Langmuir
Classical-mechanics molecular-dynamics simulations to study dynamical and structural properties of the organic-mineral interactions
298
Water-decane mixture
299
Water and decane were randomly mixed and inserted into the gap between the calcite slabs of {101
300
4} face with 1:1 volume ratio (Simulation 1). This was done to determine if immiscibility can be observed
301
at the time and length scale of MD calculations. In addition, we wanted to observe whether there is a
302
preference for water (“water wet”) or decane (“oil wet”) to adsorb to the calcite surface. As shown in Figure
303
4(A), most of the decane molecules had separated from irregular regions of water after 0.1 ns. After 0.3 ns,
304
approximately regular water layers had occupied the upper (calcite surface above the liquid phase due to
305
the periodic boundary conditions) and lower (calcite surface under the liquid phase) surface of the calcite
306
slab, and a decane layer became isolated as in Figure 4(A)-(iii). After separation between water and decane,
307
no significant change of the overall shape of the liquid fractions was observed. Thus, the calcite surface
308
became water-wet at 0.5 ns. Although both water and decane had a chance to adsorb onto the calcite surface,
309
the calcite surface preferred water rather than decane because of much stronger interactions between calcite
310
and water O atoms.
311
The same calculation as in Figure 4(A), but with an initial “oil (decane)-wet” calcite surface was
312
performed (Simulation 2) to test whether initially adsorbed hydrophobic decane molecules can stay
313
adsorbed on the calcite surface when they compete with water molecules. At the beginning of this
314
simulation, each side of surfaces is fully covered by one molecular layer of decane (~26 molecules each,
315
shown in black arrows in Figure 4(B)-(i)). The thickness of decane layer would affect the behavior of decane
316
and water on the calcite surface. Further testing could be performed using thicker decane layers, but for this
317
study, we used one monolayer of decane. After 0.2 ns, some of the water molecules have passed through
318
the decane layer and adsorbed onto the calcite surface, while the water/decane mixture became separated.
319
The decane layer initially adsorbed to the upper surface expands as the decane molecules agglomerate, but
320
still some water molecules are adsorbed onto the upper calcite surface. Gradually, more water molecules 15 ACS Paragon Plus Environment
Langmuir 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
321
break through the decane layer on the upper surface and get adsorbed onto the calcite. After 0.6 ns, about
322
two thirds of the upper calcite surface was covered by water molecules, with a completely water-wetted
323
calcite surface at the lower calcite surface. After 1 ns, the fluid is divided in three discrete layers as in
324
Simulation 1 (water-decane-water), and both sides of the calcite slab become water-wet surfaces. However,
325
the two simulation cells of water/decane mixture without any surface-active molecules are an incomplete
326
oil model, because most carbonate reservoirs are oil-wet.
16 ACS Paragon Plus Environment
Page 16 of 34
Page 17 of 34 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
Langmuir
Figure 4. (A) Simulation 1 snapshots of water and decane mixture on the calcite {1014} surface. (i) At time 0 ns, decane and water were thoroughly mixed. (ii) After 0.1 ns, water and decane were separated in most areas with some smaller mixed regions remaining. (iii) After 0.5 ns, the whole liquid splits up into three parts; water-decane-water. (B) Simulation 2 snapshots of water and decane mixture on the oil-wet calcite {1014} surface. (i) At time 0 ns, the simulation starts with a monolayer of decane covering the calcite surface and a random mixture solution of water and decane filling the gap; (ii) after 0.2 ns, water covers a large portion of the calcite surface and significant immiscibility has taken place. (iii) After 1 ns, complete water wetting of calcite and complete separation of the fluids has taken place. The two black arrows indicate the oil-wet surface layers at both sides of the calcite surface. Teal represents water molecules, red molecules are decane, and grey molecules are calcite.
327
17 ACS Paragon Plus Environment
Langmuir 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
328 329
Water-decane-phenol mixture with pre-adsorbed phenol molecules on the calcite surface
330
Calculations with phenol as an asphaltene surrogate with a hydroxyl functional group were
331
performed on the calcite surface. The phenol in this calculation is to mimic the polar-nonpolar character of
332
asphaltenes in crude oil. Since DFT calculations already compared the effect of additional benzene ring in
333
1-naphthol, only phenol was added for MD simulations. The goal is to investigate the dynamic behavior of
334
the polar-nonpolar character and its effects on the calcite surface wettability change. The research questions
335
to be answered here include: Does phenol function as a bridge between the calcite surface and the oil-
336
mimicking decane? Can phenol link non-polar decane droplets and polar water ones? Based on this, two
337
simulations were performed, water-decane-phenol mixtures with pre-adsorbed phenol on the (i) basic
338
calcite surface (mixed-wet surface; Simulation 3), and (ii) oil-wet (with both phenol and decane) calcite
339
surface (Simulation 4).
340
Simulation 3 (Figure 5(A)) started with 10/12 (upper/lower) phenol molecules pre-adsorbed on the
341
calcite surface. The number of pre-adsorbed phenol molecules was carefully tested and chosen to effectively
342
show the effect of phenol on wettability changes. The pre-adsorbed phenol molecules slowly drifted along
343
the surface, but they preferred to stay on the calcite surface. These phenol molecules attracted more phenol
344
molecules to the surface. The decane and phenol molecules away from the surfaces clustered and got
345
separated from the water phase. The liquid part eventually separated into two phases, an oil one (including
346
both phenol and decane) and a water one. The oil phase extended from the upper calcite layer to the lower
347
calcite layer until ~ 0.9 ns, and then it separated from the lower calcite surface. The number of surface
348
adsorbed phenol as a function of time was measured and shown in Figure S2, and there were 21/10
349
(upper/lower) phenol molecules adsorbed on the calcite surface after 1 ns. At this point, about 75% of the
350
initially pre-adsorbed phenol molecules were still attached to the lower calcite surface. Near the upper
351
calcite surface, ~ 60% of the calcite surface was covered with the oil phase and the remainder of the surface
352
was water-wet at 1 ns. From the comparison between the two surfaces with 10 vs. 12 pre-adsorbed phenol, 18 ACS Paragon Plus Environment
Page 18 of 34
Page 19 of 34 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
Langmuir
353
phenol and decane distribution (whether they are distributed close each other or away from each other) in
354
the solution is a more relevant condition that determines the wettability changes than the number of pre-
355
adsorbed phenol, at this coverage. Most of the phenol molecules that were not adsorbed on the calcite
356
surfaces were located at the water-decane boundary.
357
In Simulation 4, the same number of phenol molecules were pre-adsorbed at the beginning (10/12
358
at upper/lower calcite surface), but this time the calcite surfaces were “oil-wet” (Figure 5(B)). At the start
359
of the simulation, there were twelve molecules of phenol with 20 decane molecules and no water molecules
360
in the first adsorbed layer (shown in black arrows) on the lower calcite surface and ten phenol molecules
361
with 21 decane molecules and no water in the upper calcite surface first adsorbed layer. After 0.1 ns, decane
362
molecules show aggregation, and the pre-adsorbed phenol molecules held onto the calcite surface and
363
attracted more decane and phenol molecules toward the surface. After 0.2 ns, decane molecules
364
agglomerated into two groups and separated from water with phenols at the interfaces. Unlike Simulation 2,
365
the anchored phenols on the calcite surfaces cause the oil-phase layers (black arrows) to remain attached to
366
the surfaces in Simulation 4. After 1 ns, the number of phenol molecules adsorbed on the calcite surface
367
increased from 10/12 (upper/lower) to 21/24 (upper/lower) (Figure S2).
19 ACS Paragon Plus Environment
Langmuir 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
Figure 5. (A) Simulation 3 snapshots of water, decane, and phenol mixture on the calcite {1014} surface with 10/12 (upper/lower) molecules of pre-adsorbed phenol molecules on the calcite surface, (i) at time 0 ns, (ii) after 0.5 ns, and (iii) after 1 ns. (B) Simulation 4 snapshots of a water-decane-phenol mixture on an oil-wet calcite {1014} surface with 10/12 (upper/lower) molecules of pre-adsorbed phenols on each side of the calcite surface; (i) at time 0 ns, (ii) after 0.2 ns, and (iii) after 1 ns. The two black arrows indicate the oil-wet surface layers on both sides of the calcite surface consisting of phenol and decane. Teal represents water molecules, red molecules are decane, light green molecules are phenol, and grey molecules are calcite.
368 369
Analysis of molecular distributions
370
Radial distribution functions (RDFs, g(r)) yield the number density of particles at a given distance
371
interval from a reference atom relative to the average number density of those particles. Therefore, 20 ACS Paragon Plus Environment
Page 20 of 34
Page 21 of 34 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
Langmuir
372
calculating RDFs for pairs of atoms shows how the density of a particle varies as a function of distance
373
from a reference atom. To investigate differences between water-calcite and phenol-calcite interactions,
374
RDFs were calculated for both cases. The procedure consists in creating a histogram of distances between
375
all the particle pairs and normalizing that histogram with respect to the volume of each spherical shell. As
376
the distance increases toward infinity, g(r) functions converge to one, which means that the number density
377
of the atoms around the reference atom is homogeneous and equal to the average number density. The first
378
peak of the RDFs corresponds to the average distance between the two atoms in a pair.
379
The RDFs of water (Figure S3(a)) calculated from the final result of Simulation 1 show the average
380
distance between water and the calcite surface because the calcite surface became completely water-wet at
381
the end of the simulation. Cacalcite-Owater shows its first peak at 2.5 Å, and Ocalcite-Hwater has its first peak at
382
1.9 Å, which agrees well with previous studies15-16. For phenol adsorption onto the calcite surface, RDF of
383
Ca-Ophenol has its first peak at 2.6 Å, and Ocalcite-Hphenol (hydroxyl group) has its first peak at 1.7 Å (Figure S3(c)).
384
These distances are almost the same as the ones in our DFT calculations, meaning that our interatomic
385
potentials between phenol and calcite are consistent with our quantum-mechanical calculations.
386
According to the RDF analyses, the average distance of Cacalcite-Owater is shorter than Cacalcite-Ophenol
387
distance. However, the adsorption energy calculation using the force fields showed a 0.08 eV more
388
favorable adsorption of phenol than that of water on the calcite {1014} surface. This explains why 75% of
389
the pre-adsorbed phenol molecules in Simulation 3 remained on the calcite surface after 1 ns, while decane
390
molecules were driven away from the surface by water molecules. Putting together the Ca-O distances and
391
the adsorption energy values of water and phenol adsorption reactions, we can infer that phenol has
392
attractive interactions with calcite not only through its hydroxyl functional group but also through its
393
benzene ring.
394
However, the interaction between calcite and the benzene ring of phenol is considered to be
395
somewhat weaker than the Ca-Ophenol interaction. This is obviously observed when we compare the
396
optimized geometries of phenol at high vs. low coverage. The energy optimized geometry of one phenol 21 ACS Paragon Plus Environment
Langmuir 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
397
molecule on the calcite surface using our force field was almost the same as our DFT calculation result
398
(2.6 Å of Ca-Ophenol distance and ~35° angle relative to the calcite surface). However, when the surface
399
coverage increases, as shown in Figures 5 simulation images, phenol interacts with the calcite surface
400
mainly through its hydroxyl group and shows a more upright position with respect to the surface.
401
Interactions between the benzene rings of neighboring phenol molecules could also play a role to their
402
upright position at high coverage. Similar results were observed in calculations for benzoate, which has one
403
benzene ring with a carboxyl functional group, adsorption onto the calcite surface by Chun et al45. With a
404
high coverage of benzoate, benzoate molecules stood perpendicular to the surface. In contrast, at low
405
coverage, half of the benzoate molecules lay at a low angle due to the interactions between the phenyl group
406
of benzoate and the calcite surface. At high surface coverage, adsorbates predominantly interact with
407
mineral surfaces through their functional groups, while at low surface coverage, the hydrophobic
408
hydrocarbons can play a more relevant role.
409
Comparison between the density profiles of Simulation 2 and that of Simulation 4 graphically
410
demonstrates the effect of phenol on the wettability change of the calcite surface. Simulation 2 contains
411
only decane in its oil phase, and simulation 4 contains both decane and phenol in its oil phase. The initial
412
density profiles of the two cases are almost the same (dotted lines in Figure 6). In both scenarios, only the
413
oil phase (black lines) without any water is located near the calcite surfaces, and water molecules were
414
initially distributed in between the two surface-adsorbed oil layers with a density of ~0.53 g/cm3 (as the
415
liquid phase is half water and half oil).
416
After 1 ns, the calcite surfaces conditioned with only decane (Figure 6(a)) were covered by water
417
molecules (red solid line), and oil molecules (black solid line) concentrated in the middle between the two
418
calcite slabs. In contrast, the calcite surfaces conditioned with both decane and phenol (Figure 6(b)) show
419
two aggregates of oil phases near the calcite surfaces. Water molecules were mostly located in between the
420
two oil phases, although there were still some water molecules near the calcite surfaces. As observed in the
421
snapshots of each simulation, phenol molecules were strongly bonded to the calcite surface through its 22 ACS Paragon Plus Environment
Page 22 of 34
Page 23 of 34
422
hydroxyl functional group and attracted oil-mimicking decane molecules toward the calcite surfaces. Since
423
there was no polar components in the oil phase of the Figure 6(a), decane molecules were driven away from
424
the calcite surface by water molecules and eventually, water molecules covered the whole calcite surfaces.
2.0
2.0 0s
1 ns
0s
oil (decane) water
1.5
Density (g/cm3)
Density (g/cm3)
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
Langmuir
1.0
0.5
0.0
0
20
40
Distance from the calcite surface (Å)
oil (decane+phenol) water
1.5
1.0
0.5
0.0
60
1 ns
0
20
40
60
Distance from the calcite surface (Å)
Figure 6. Mass density profiles of water-oil random mixtures on the oil-wet calcite surface with (a) only decane (Simulation 2) and (b) both phenol and decane (Simulation 4) in their oil phase. The distance from the calcite surface shows the perpendicular distance from the calcite {1014} surface. At both distances of 0 Å and 60 Å, the liquid phase is in contact with calcite surfaces due to the periodic boundary conditions. Red lines represent the water density profiles and black lines represent the oil phase (includes both phenol and decane) density profiles. Dashed lines show the density profiles at time 0 s, and solid lines show the density profiles at 1 ns. Data were collected with a bin width of 1 Å.
425
In order to compare the interactions of a decane molecule with a monolayer of water vs. a
426
monolayer of phenol, the adsorption strength of the decane molecule on each monolayer was calculated.
427
First, a monolayer of phenol or water previously adsorbed to a calcite slab was relaxed to its lowest energy
428
level. Then, the optimized positions of a decane molecule on both monolayers were calculated (Figure 7(a)
429
and (b)). Finally, with fixed geometries of phenol or water monolayer, the adsorption energies of a decane
430
molecule were calculated at several different distances between the decane molecule and the calcite surface
431
using the equation (1) (Figure 7(c)). Results show that decane is more strongly bonded onto the
432
phenol-wetted calcite surface than onto the water-wetted surface (by 0.3 eV). Considering the enthalpy of
433
vaporization for decane molecules (51.5 kJ/mol or 0.53 eV)46, the interaction with other decane molecules
434
is stronger than the interaction with a water monolayer (0.25 eV) and is comparable to the adsorption energy 23 ACS Paragon Plus Environment
Langmuir
435
to a phenol monolayer (0.54 eV). Thus, decane would prefer to stick to itself rather than adsorb onto water
436
layer, while decane self-aggregation and adsorption onto the phenol layer would compete with each other.
437
This explains why the oil phase accumulates on the calcite surfaces only when phenol molecules were pre-
438
adsorbed on the surfaces.
(a)
decane
(c) phenol
(b) water
Adsorption energy (eV)
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 34
Water layer Phenol layer
0.5
0.0
-0.5 0
2
4
6
Distance (Å)
Figure 7. Adsorption energy calculations of a decane molecule on either a phenol layer or a water layer adsorbed on a calcite surface. Optimized geometry of a decane molecule adsorbed on (a) a phenol-wetted calcite surface, and (b) a water-wetted calcite surface (Ca-green, O-red, C- grey, H-pink). (c) Adsorption energy curves as a function of the distance from the calcite surface to the decane carbon, where the distance 0 Å is set to be the optimized height for each case; 7.6 Å for phenol and 5.2 Å for water.
439
Although there are polar components like phenol in the oil phase, the non-polar oil components
440
will not easily stick onto calcite surfaces unless the surfaces are already oil-wet. In the density profiles from
441
results of Simulation 3 and Simulation 4 (Figure 8), both simulations had the same number of surface-
442
adsorbed phenol molecules at the beginning. However, only the simulation that started with an oil-wet
443
surface (Simulation 4) showed adsorbed decane molecules at its lower calcite surface at time 1 ns (Figure
444
8(a), black lines). In the density profiles from the final result of Simulation 4 (solid lines), decane molecules
445
are attracted to the calcite surface adsorption layers (shaded areas) where there is a high concentration of
446
phenol with consequently a lower water density. In contrast, the density profiles obtained from the final
447
result of Simulation 3 (dotted lines) show no decane in the lower surface adsorption layer and few decane
24 ACS Paragon Plus Environment
Page 25 of 34
448
in the upper surface adsorption layer. The water density is ~1 g/cm3 near the lower calcite surface for
449
Simulation 3, which means that the surface is almost exclusively covered by water molecules at t = 1 ns.
450
When phenol molecules were adsorbed onto the calcite surface sporadically, this did not necessarily induce
451
oil accumulation on the surface. However, forming even one monolayer of oil film, which contains phenol
452
molecules, facilitates further accumulation of hydrocarbons close to the surface.
(a)
(b) 1.2
1.2
decane water phenol non oil-wet surface oil-wet surface
1.0
Density (g/cm3)
Density (g/cm3)
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
Langmuir
0.8 0.6 0.4
1.0
decane water phenol non oil-wet surface oil-wet surface
0.8 0.6 0.4 0.2
0.2
0.0
0.0 0
5
10
15
40
20
45
50
55
60
Distance from the calcite surface (Å)
Distance from the calcite surface (Å)
Figure 8. Mass density profiles of water-phenol-decane random mixture on the mixed-wet calcite surface (dotted lines, open symbols, Simulation 3) and oil-wet calcite surface (solid lines, closed symbols, Simulation 4) at time 1 ns. Mixed-wet and oil-wet surfaces describe the initial surface configurations at time t=0 s. The distance from the calcite surface shows the perpendicular distance from the calcite {1014} surface. At both distances of 0 Å and 60 Å, the liquid is in contact with the calcite surfaces due to the periodic boundary conditions. (a) Density profiles near the lower calcite surface with a distance range from 0 Å to 20 Å, and (b) density profiles near the upper calcite surface with a distance range from 40 Å to 60 Å. Shaded areas indicate the surface adsorption layers. Black lines – decane, blue lines –water, red lines – phenol.
453
Our calculations show that decane molecules hardly stay on the calcite surface when they compete
454
with water molecules. However, when they co-exist with polar components of oil like phenol, which bind
455
strongly on the calcite surface through their polar functional groups, decane molecules can stay on the
456
calcite surface linked by phenol molecules. At the interface of water-decane, phenol molecules enclose
457
clusters of decane molecules (mimicking “oil droplets”) and prevent water from pushing away decane
458
molecules from the surface. Phenol also attract other phenol and decane molecules toward the calcite
459
surface which results in a bigger aggregation of oil molecules.
25 ACS Paragon Plus Environment
Langmuir 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
460
To investigate the influence of temperature on the behavior of oil molecules on calcite surfaces,
461
additional simulations were performed at 390 K (Simulation 5), which is a plausible temperature range in
462
an oil reservoir. Simulation 5 (Figure S4) started with the same configuration as in Simulation 4 (Figure
463
5(B)). At this high temperature, water penetrated faster into the oil-wetted layer to reach the calcite surface
464
compared to the results at 300 K. Even at 0.1 ns, some water molecules adsorbed onto the upper calcite
465
surface, starting to make the upper calcite surface mixed-wet. Density profiles of the final configuration (at
466
t=1 ns) are shown in Figure S5. More water molecules gathered near the upper calcite surface compared to
467
the 300 K result. But the overall density profile shape is similar; oil molecules agglomerate into two groups
468
near the calcite surface, and water phase mostly centered in between the two oil phases. In this higher
469
temperature condition, phenol defines and decorates the phase boundary between decane and water less
470
well than at 300 K.
471
Temperature dependence of wettability changes depends on the reservoir conditions. In the case of
472
low salinity water (LSW), wettability change to water-wet state was more activated at high temperature in
473
several studies. Zhang et al.47 have proven that lowering salinity but increasing sulfate concentration
474
significantly improves oil release at higher temperature when comparing the oil recovery of the core
475
samples at 70 °C and 120 °C. Heidari et al.48 showed obtaining the intermediate wetting condition from
476
aged core samples took less time at higher temperature. In combination of high temperature with
477
appropriate salinity condition, we could increase the extent of water wetting in the oil reservoir.
478
At higher temperature, dissolution of calcite could takes place. Calcite dissolution is also one of
479
the possible mechanisms that releases oil from the reservoir pore spaces, which increases oil recovery.49-50
480
However, due to the limitation of simulation time scale and the size of the system, modeling the calcite
481
dissolution contribution in our system is left for further study. Other conditions such as lower pH, different
482
water compositions, and even salinity changes can also cause calcite dissolution.
483
During oil migration in a reservoir, the water film on the water-wet calcite surface can be
484
destabilized by the high disjoining pressure51-53. At this point, crude oil can break through the water film 26 ACS Paragon Plus Environment
Page 26 of 34
Page 27 of 34 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
Langmuir
485
and reach calcite surfaces. After the water-film breakage, if the oil phase covers the calcite surface and
486
forms a locally oil-wet surface (similarly in our Simulation 4), the surface-active molecules in oil will
487
aggregate and stay on the calcite surface changing the mineral surface wettability to oil-wet. In addition,
488
π-π stacking interactions between asphaltenes facilitate the formation of asphaltene nanoaggregates44. Thus,
489
when polar components of crude oil such as asphaltenes break through the water film and adsorb onto the
490
calcite surface through their polar functional groups, such as hydroxyl, they will form aggregates and
491
clusters that can stick onto the calcite surface making the calcite surface oil-wet.
492
Conclusions
493
The effect of phenol and 1-naphthol on the wettability alteration of the calcite surface were studied
494
using DFT and classical MD calculations. DFT results show that the adsorption of phenol and 1-naphthol
495
occur preferentially between their hydroxyl group and step edges of the calcite surface. However, their
496
adsorption geometries were different due to the larger van der Waals interactions associated with the
497
additional benzene ring in 1-naphthol. In the classical MD simulations, separation of oil and water was
498
clearly observed. Non-polar oil components (decane) were separated from the water phase, while phenol
499
locates at the oil/water interface. When oil contains only non-polar oil components, the oil phase is repelled
500
from the calcite surface by water molecules. In contrast, the oil phase can reside on the calcite surface when
501
it contains phenol. The results show that phenol is an appropriate representative of asphaltenes in crude oil
502
with a hydroxyl functional group, and this hydroxyl group has a relevant role in adsorption processes of oil
503
onto mineral surfaces that will result in surface wettability changes. However, although the oil phase is a
504
mixture of polar and non-polar components, non-polar components were found to reside on the calcite
505
surface only when the simulations began with an oil-wet calcite surface. Initially mixed-wet calcite surface
506
resulted in one side of the surface water-wet while the other side showing mixed-wet after 1 ns of simulation.
507
Having model substances that mimic the role of wettability modifiers in oil helps designing
508
experiments and computational approaches to better understand and control how oil ad- and desorbs from
509
specific mineral surfaces. This study elucidates to what degree phenol and 1-naphthol can be used as such 27 ACS Paragon Plus Environment
Langmuir 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
510
model substances. Overall, we find that phenol and 1-naphthol have a similar adsorption behavior to
511
compounds, mainly in the asphaltene fraction of oil, that are responsible for creating the link between the
512
non-polar fraction of the oil and the typically polar mineral surfaces. One difference though is the greater
513
size of molecules in asphaltene which also causes some differences in the physicochemical adsorption
514
properties, e.g., obscuring small surface steps by asphaltene rather than pronouncing them. However, there
515
is good enough of an agreement to experimentally and computationally test the influence of, e.g., pH on
516
the wettability changes caused by phenol/1-naphthol. Since the pH of ocean water, the pKa of phenol, and
517
the point of zero charge (pzc) of calcite are all in the slightly basic range, small changes in pH can have a
518
significant influence on the strength of interaction between these. Given that phenol and 1-naphthol are
519
good proxies for wettability modifiers, they can also be used in future experimental and computational
520
studies to test the influence of temperature (to some degrees done in this study), pressure, salinity, and
521
competition with other surfactants that may be used in future enhance oil recovery procedures.
522
Acknowledgments
523
This work was supported by ACS-PRF grant #ACS PRF 55659-ND5. The authors are grateful for the
524
support from the American Chemical Society Petroleum Research Fund.
525
Supporting Information
526
The Supporting Information is available free of charge on the ACS Publications website.
527
Interatomic potential sets parameters (Table S1, Table S3); calculated calcite lattice parameters using our
528
potential sets (Table S2); energy curve for validity of our potential sets (Figure S1); number of surface
529
adsorbed phenol (Figure S2); RDF graphs (Figure S3); high temperature simulation (Figure S4) and its
530
density profile (Figure S5).
531
28 ACS Paragon Plus Environment
Page 28 of 34
Page 29 of 34 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
Langmuir
532
References
533
(1) Donaldson, E. C.; Chilingarian, G. V.; Yen, T. F., Enhanced oil recovery. Dev. Pet. Sci. 1989, 17 (Part
534
B).
535
(2) Abdallah, W.; Buckley, J. S.; Carnegie, A.; Herold, J. E. B.; Fordham, E.; Graue, A.; Signer, T. H. N.
536
S. C.; Hussain, H.; Montaron, B.; Ziauddin, M., Fundamentals of wettability. Oilfield Rev. 2007, 19,
537
44-63.
538 539 540 541 542 543 544 545 546 547 548 549 550 551 552
(3) Falode, O.; Manuel, E., Wettability effects on capillary pressure, relative permeability, and irredcucible saturation using porous plate. J. Pet. Eng. 2014, 2014 (465418). (4) Anderson, W. G., Wettability literature survey. Part 5: The effects of wettability on relative permeability. J. Pet. Technol. 1987, 39 (11), 1453-1468. (5) Anderson, W. G., Wettability literature survey. Part 1: Rock-oil-brine interactions and the effects of core handling on wettability. J. Pet. Technol. 1986, 38 (11), 1125-1144. (6) Kumar, K.; Dao, E. K.; Mohanty, K. K., Atomic force microscopy study of wettability alteration by surfactants. Soc. Pet. Eng. J. 2008, 13 (2), 137-145. (7) Buckley, J. S., Wetting alteration of solid surfaces by crude oils and their asphaltenes. Rev. Inst. Fr. Pet. 1998, 53 (3), 303-312. (8) Denekas, M. O.; Mattax, C. C.; Davis, G. T., Effects of crude oil components on rock wettability. Soc. Pet. Eng. J. 1959, 216, 330-333. (9) Yen, T. F.; Chilingarian, G. V., Asphaltenes and asphalts, 2. 1st ed.; Elsevier Science: Amsterdam ; New York, 2000. (10) Ghatee, M. H.; Koleini, M. M.; Ayatollahi, S., Molecular dynamics simulation investigation of
553
hexanoic acid adsorption onto calcite (1014) surface. Fluid Phase Equilib. 2015, 387, 24-31.
554
(11) Sedghi, M.; Piri, M.; Goual, L., Atomistic molecular dynamics simulations of crude oil/brine
555
displacement in calcite mesopores. Langmuir 2016, 32 (14), 3375-3384.
29 ACS Paragon Plus Environment
Langmuir 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
556 557
(12) Legens, C.; Palermo, T.; Toulhoat, H.; Fafet, A.; Koutsoukos, P., Carbonate rock wettability changes induced by organic compound adsorption. J. Pet. Sci. Eng. 1998, 20 (3-4), 277-282.
558
(13) Karoussi, O.; Skovbjerg, L. L.; Hassenkam, T.; Stipp, S. L. S.; Hamouda, A. A., AFM study of
559
calcite surface exposed to stearic and heptanoic acids. Colloid Surf., A 2008, 325 (3), 107-114.
560 561 562 563
(14) Standnes, D. C.; Austad, T., Wettability alteration in chalk 1. Preparation of core material and oil properties. J. Pet. Sci. Eng. 2000, 28 (3), 111-121. (15) Cooke, D. J.; Gray, R. J.; Sand, K. K.; Stipp, S. L.; Elliott, J. A., Interaction of ethanol and water with the {1014} surface of calcite. Langmuir 2010, 26 (18), 14520-14529.
564
(16) Sand, K. K.; Yang, M.; Makovicky, E.; Cooke, D. J.; Hassenkam, T.; Bechgaard, K.; Stipp, S. L. S.,
565
Binding of ethanol on calcite: The role of the oh bond and its relevance to biomineralization. Langmuir
566
2010, 26 (19), 15239-15247.
567
(17) Keller, K. S.; Olsson, M. H. M.; Yang, M.; Stipp, S. L. S., Adsorption of ethanol and water on
568
calcite: Dependence on surface geometry and effect on surface behavior. Langmuir 2015, 31 (13),
569
3847-3853.
570 571 572
(18) Bovet, N.; Yang, M.; Javadi, M. S.; Stipp, S. L. S., Interaction of alcohols with the calcite surface. Phys. Chem. Chem. Phys. 2015, 17 (5), 3490-3496. (19) Hakim, S. S.; Olsson, M. H. M.; Sørensen, H. O.; Bovet, N.; Bohr, J.; Feidenhans'l, R.; Stipp, S. L.
573
S., Interactions of the calcite {10.4} surface with organic compounds: Structure and behaviour at
574
mineral - organic interfaces. Sci. Rep. 2017, 7.
575
(20) Olsen, R.; Leirvik, K. N.; Kvamme, B.; Kuznetsova, T., Adsorption properties of triethylene glycol
576
on a hydrated {1014} calcite surface and its effect on adsorbed water. Langmuir 2015, 31 (31), 8606-
577
8617.
578 579 580 581
(21) Tomson, M. B.; Kan, A. T.; Fu, G.; Al-Thubaiti, M.; Shen, D.; Shipley, H. J., Scale formation and prevention in the presence of hydrate inhibitors. Soc. Pet. Eng. J. 2006, 11 (02), 248-258. (22) van Duin, A. C.; Larter, S. R., A computational chemical study of penetration and displacement of water films near mineral surfaces. Geochem. Trans. 2001, 2 (1), 35. 30 ACS Paragon Plus Environment
Page 30 of 34
Page 31 of 34 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
582 583 584 585 586
Langmuir
(23) Delley, B., DMol, a standard tool for density functional calculations: Review and advances. Theor. Comput. Chem. 1995, 2, 221-254. (24) Perdew, J. P.; Burke, K.; Wang, Y., Generalized gradient approximation for the exchange-correlation hole of a many-electron system. Phys Rev B Condens Matter 1996, 54 (23), 16533-16539. (25) Klamt, A.; Schuurmann, G., COSMO: a new approach to dielectric screening in solvents with
587
explicit expressions for the screening energy and its gradient. J. Chem. Soc., Perkin Trans. 2 1993, (5),
588
799-805.
589
(26) Segall, M. D.; Lindan, P. J. D.; Probert, M. J.; Pickard, C. J.; Hasnip, P. J.; Clark, S. J.; Payne, M. C.,
590
First-principles simulation: ideas, illustrations and the CASTEP code. J. Phys.: Condens. Matter 2002,
591
14 (11), 2717-2744.
592 593 594
(27) Price, D. J.; Brooks III, C. L., A modified TIP3P water potential for simulation with Ewald summation. J. Chem. Phys. 2004, 121 (20), 10096-10103. (28) Jorgensen, W. L.; Maxwell, D. S.; TiradoRives, J., Development and testing of the OPLS All-Atom
595
force field on conformational energetics and properties of organic liquids. J. Am. Chem. Soc. 1996, 118
596
(45), 11225-11236.
597 598 599 600
(29) Pavese, A.; Catti, M.; Parker, S. C.; Wall, A., Modelling of the thermal dependence of structural and elastic properties of calcite, CaCO3. Phys. Chem. Miner. 1996, 23 (2), 89-93. (30) de Leeuw, N. H.; Parker, S. C.; Harding, J. H., Molecular dynamics simulation of crystal dissolution from calcite. Phys. Rev. B 1999, 60 (19), 13792-13799.
601
(31) deLeeuw, N. H.; Parker, S. C., Atomistic simulation of the effect of molecular adsorption of water on
602
the surface structure and energies of calcite surfaces. J. Chem. Soc. Faraday T 1997, 93 (3), 467-475.
603
(32) Graf, D. L., Crystallographic tables for the rhombohedral carbonates. Am. Mineral. 1961, 46, 1283-
604 605 606 607
1316. (33) Freeman, C. L.; Harding, J. H.; Cooke, D. J.; Elliott, J. A.; Lardge, J. S.; Duffy, D. M., New forcefields for modeling biomineralization processes. J. Phys. Chem. C 2007, 111 (32), 11943-11951. (34) Dassault Systèmes BIOVIA, Materials Studion 7.0. San Diego: Dassault Systèmes 2018. 31 ACS Paragon Plus Environment
Langmuir 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
608 609 610 611 612 613 614 615 616
(35) Plimpton, S., Fast parallel algorithms for short-range molecular dynamics. J. Compt. Phys. 1995, 117, 1-19. (36) Humphrey, W.; Dalke, A.; Schulten, K., VMD - Visual Molecular Dynamics. J. Mol. Graphics 1996, 14, 33-38. (37) Tkatchenko, A.; Scheffler, M., Accurate molecular van der waals interactions from ground-state electron density and free-atom reference data. Phys. Rev. Lett. 2009, 102 (7). (38) Grimme, S., Semiempirical GGA-type density functional constructed with a long-range dispersion correction. J. Comput. Chem. 2006, 27 (15), 1787-1799. (39) Torres, A.; Suarez, J. A.; Remesal, E. R.; Marquez, A. M.; Sanz, J. F.; Canibano, C. R., Adsorption
617
of prototypical asphaltenes on silica: First-principles DFT simulations including dispersion corrections.
618
J. Phys. Chem. B 2018, 122 (2), 618-624.
619 620 621
(40) Ronca, E.; Belpassi, L.; Tarantelli, F., A quantitative view of charge transfer in the hydrogen bond: the water dimer case. ChemPhysChem 2014, 15 (13), 2682-2687. (41) Ratajczak, H.; Orville-Thomas, W. J., Charge-transfer properties of hydrogen-bonds. 3. Charge-
622
transfer theory and relation between energy and enhancement of dipole-moment of hydrogen-bonded
623
complexes. J. Chem. Phys. 1973, 58 (3), 911-919.
624 625 626 627 628
(42) Mulliken, R. S., Electronic population analysis on LCAO-MO molecular wave functions. 1. J. Chem. Phys. 1955, 23 (10), 1833-1840. (43) Lu, J. Y.; Ge, Q.; Li, H.; Raza, A.; Zhang, T., Direct prediction of calcite surface wettability with first-principles quantum simulation. J. Phys. Chem. Lett. 2017, 8 (21), 5309-5316. (44) Alvim, R. S.; Lima, F. C. D. A.; Sanchez, V. M.; Headen, T. F.; Boek, E. S.; Miranda, C. R.,
629
Adsorption of asphaltenes on the calcite (10.4) surface by first-principles calculations. Rsc Adv. 2016, 6
630
(97), 95328-95336.
631 632
(45) Chun, B. J.; Lee, S. G.; Choi, J. I.; Jang, S. S., Adsorption of carboxylate on calcium carbonate (101 4) surface: Molecular simulation approach. Colloid Surf., A 2015, 474, 9-17.
32 ACS Paragon Plus Environment
Page 32 of 34
Page 33 of 34 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
633
Langmuir
(46) Viton, C.; Chavret, M.; Jose, J., Enthalpies of vaporization of normal alkanes from nonane to
634
pentadecane at temperatures from 298 to 359 K. ELDATA: Int. Electron. J. Phys.-Chem. Data 1996, 2
635
(3), 103.
636
(47) Zhang, Y.; Sarma, H., Improving waterflood recovery efficiency in carbonate reservoirs through
637
salinity variations and ionic exchanges: A promising low-cost "smart-waterflood" approach. Soc. Pet.
638
Eng. 2012.
639
(48) Heidari, M. A.; Habibi, A.; Ayatollahi, S.; Masihi, M.; Ashoorian, S., Effect of time and temperature
640
on crude oil aging to do a right surfactant flooding with a new approach. Proceedings of the Offshore
641
Technology Conference-Asia; Kuala Lumpur, Malaysia 2014.
642 643 644 645 646 647 648 649 650
(49) Morrow, N.; Buckley, J. S., Improved oil recovery by low-salinity waterflooding. J. Pet. Technol. 2011, 63 (5), 106-112. (50) Zahid, A.; Stenby, E. H.; Shapiro, A. A., Smart waterflooding (high sal/low sal) in carbonate reservoirs. J. Pet. Eng. 2012. (51) Jackson, M. D.; Al-Mahrouqi, D.; Vinogradov, J., Zeta potential in oil-water-carbonate systems and its impact on oil recovery during controlled salinity water-flooding. Sci. Rep. 2016, 6. (52) Sohal, M. A.; Thyne, G.; Søgaard, E. G., Review of recovery mechanisms of ionically modified waterflood in carbonate reservoirs. Energy Fuels 2016, 30 (3), 1904-1914. (53) Hirasaki, G. J., Wettability: Fundamentals and surface forces. SPE Form. Eval. 1991, 6, 217-226.
651
33 ACS Paragon Plus Environment
Langmuir 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
652
For Table of Contents Only
653
654 655 656
34 ACS Paragon Plus Environment
Page 34 of 34