Sunteți pe pagina 1din 13

Fuel 216 (2018) 559–571

Contents lists available at ScienceDirect

Fuel
journal homepage: www.elsevier.com/locate/fuel

Full Length Article

Insight into the stability of hydrophilic silica nanoparticles in seawater for T


Enhanced oil recovery implications

Saeed Jafari Daghlian Sofla, Lesley Anne James , Yahui Zhang
Department of Process Engineering, Memorial University of Newfoundland, St. John’s, NL, A1B 3X9 Canada

G RA P H I C A L AB S T R A C T

A R T I C L E I N F O A B S T R A C T

Keywords: The stability of nanoparticles in the aqueous phase is a major challenge in the application of nanoparticles in
Enhanced oil recovery Enhanced Oil Recovery (Nano-EOR) processes. Previous studies evaluated the performance of nanoparticles for
Silica nanoparticles EOR purposes; either deionized water or water at very low ionic strength was used. Nanoparticles can be easily
Stability of nanoparticles dispersed in the deionized or low salinity water, whereas they are extremely unstable in high salinity seawater or
Electrical double layer
formation water. Typically, seawater or formation brine is injected for water-flooding and EOR purposes. If we
DLVO theory
want to change the fluid–fluid or fluid-rock properties by injecting nanoparticle enhanced water, then, the
stability of the nanoparticles in high salinity water is extremely important. In this work, a novel and simple
method to stabilize silica nanoparticles in seawater is proposed. First, the aggregation of silica nanoparticles in
the presence of different ions is investigated. The results show that the presence of positive multivalent ions in
the electrical double layer around nanoparticles can destabilize silica nanoparticles. In order to reduce the
concentration of positive multivalent ions around silica nanoparticles, a theory based on “H+ protection” is
proposed and its effectiveness is tested by particle size, turbidity, zeta-potential, and pH measurements. The
effect of the concentrations of nanoparticles and HCl on the stability of silica nanoparticles in seawater is
evaluated. Experimental results show that H+ protection, which can be obtained by adding HCl to the solution,
can effectively stabilize silica nanoparticles in seawater. The experiments show that the size of nanoparticles in
the seawater directly depends on the concentration of nanoparticles and inversely to the HCl concentration.


Corresponding author.
E-mail address: ljames@mun.ca (L.A. James).

https://doi.org/10.1016/j.fuel.2017.11.091
Received 13 June 2017; Received in revised form 11 October 2017; Accepted 22 November 2017
Available online 16 December 2017
0016-2361/ © 2017 Elsevier Ltd. All rights reserved.
S. Jafari Daghlian Sofla et al. Fuel 216 (2018) 559–571

1. Introduction particles should overcome attractive forces [27].


Derijaguin, Landau, Overbeek and Verwey [28,29] proposed the
Recently, there has been a growing interest in application of na- DLVO theory to explain the stability of colloids in the absence of any
noparticles in the Enhanced Oil Recovery (EOR) processes. Numerous polymer or surfactant. This theory combined two independent van der
experimental works has been published discussing the effect of nano- Waals attraction and electrostatic repulsion forces, explain dispersion
particles on increasing oil recovery [1–5]. It is reported that nano- mechanisms of colloids in the polar solution.
particles can adsorb at liquid-liquid interfaces and reduce interfacial
tension [6–8]. Whether nanoparticles adsorb at the interface or they 2.1. Electrical double layer
change the oil-water interfacial tension is still an ongoing debate. It is
accepted that this phenomenon occurs because the adsorption lowers Helmholtz [30] first introduced and termed the idea of the electrical
the total energy of the system [9]. Furthermore, nanoparticles can alter double layer, which was later extended by Gouy-Chapman and Stern
surface wettability from oil-wet to water-wet [10–13] which is favor- [31,32].The electrical double layer is a structure that appears on the
able for oil recovery. surface of a charged surface when it is exposed to a fluid. The first layer,
One of the most important challenges in the application of nano- the surface charge (either positive or negative), is comprised of ions
particles for EOR methods is their stability in an aqueous solution. adsorbed onto the surface due to chemical interactions. This layer,
Nanoparticle dispersion in the aqueous phase is not a thermo- which consists of a dense layer of ions of the opposite charge (counter-
dynamically stable. Dispersed nanoparticles are always subject to ions) that form around the nanoparticle, is known as the “Stern layer”.
Brownian motion with frequent collisions between them. The stability The second layer is composed of ions attracted to the surface charge via
of a dispersion is thus determined by the nature of the interactions the Coulomb forces, electrically screening the first layer. This second
between the particles during such collisions [14]. Although the poten- layer is loosely associated with the surface. It is made of free ions that
tial of silica nanoparticles in EOR processes is widely studied and their move in the fluid under the influence of electric attraction and thermal
effectiveness is well-documented [3,15,16] the applications of hydro- motion rather than being firmly anchored. It is thus called the “diffuse
philic silica nanoparticles are limited because the nanoparticles’ high layer” [33]. The high concentration of counter-ions within the diffuse
energetic hydrophilic surface, causes the silica nanoparticles to be ea- layer gradually decreases with increasing distance from the nano-
sily agglomerated [17]. particle until equilibrium is reached with the ion concentration in the
In most studies which evaluated performance of nanoparticles for bulk of the solvent [31]. The distribution of ions in the electrical double
EOR purposes with wettability, IFT measurement and core-flooding layer around negatively charged nanoparticles is illustrated in Fig. 1.
experiments, either deionized water or water at very low ionic strength The thickness of the double layer that forms at the charged surface
(especially NaCl brine) is used [18–25]. However, nanoparticles are is called “Debye Length”. Based on the electrolyte theories, interactions
extremely unstable in seawater, formation water and concentrated ionic in the low ionic strength solutions decrease exponentially with distance
solutions of multiple types and charges of ions. Water flooding and EOR or the Debye screening length. By increasing the ion concentration in
projects, especially for offshore reservoirs, use seawater. Even higher the solution, due to effective screening of charges over short distances,
salinity formation water is present in the reservoirs as well. The stabi- this length decreases monotonically [35]. Greater nanoparticle surface
lity of nanoparticles in these fluids is crucial for any successful nano- charge and longer Debye length leads to increasing nanoparticle sta-
EOR processes. In this paper, nanoparticle stability in mixed ionic so- bility in the aqueous solution [36,37]. The thickness of the double layer
lutions is systematically investigated and a new method to stabilize is a function of ionic strength. The ionic strength can be defined as [38]:
hydrophilic silica nanoparticles in seawater is proposed. We first ex-
amined the effect of the most common ions in seawater on the stability 1
I= ∑ z i2 ci
of silica nanoparticles. The purpose of this step was confirmation of 2 i (1)
previous findings (see [46,66]). Based on the results of these experi-
ments, a novel method, H+ protection, is proposed to stabilize silica where z and c are the charge number and molar concentration of ith
ion, respectively. The Debye length (k−1) in nanometer can be calcu-
nanoparticles in seawater. The effectiveness of proposed method in the
stability of silica nanoparticles in seawater is tested by particle size, lated as [39]:
turbidity, zeta-potential and pH measurements. Furthermore, in each εr ε ∘KB T
part, the results of experiments are compared with DLVO theory. k −1 =
e 2 ∑ ρ∞i Zi2 (2)
2. Theory of nanoparticle stability
where e is the elementary charge of an electron (C), T is the tempera-
ture (K), ε ∘ is (F/C), εr is absolute and solution relative dielectric con-
Colloidal systems consist of one or more dispersed phases and one
stant, KB is the Boltzmann’s constant, and ρ∞i is the number density of
continuous phase. On the nano-scale, due to an increase in the surface
ion i in the bulk solution. The electrostatic repulsion between two
area and possible changes in the structure and composition of the
equally sized spherical particles with ka > 5(where k is the reciprocal
surface, surface energy, and consequently the total energy of the
of Debye length (nm−1) and a is the radius of spherical nanoparticles in
system, increases. Nanoparticles tend to aggregate to reduce the surface
nanometer) can be calculated by [40]:
energy, thereby making a colloidal dispersion at the nano-scale non-
thermodynamically stable. Particles in the colloidal systems are always 32πKB Tεr ε ∘ρ∞ rγ 2 −kd
VEL = e
subject to Brownian motion and collisions frequently occur between k2 (3)
particles. The nature of interaction between the particles during these
collisions determines their stability in the solution. Van der Waals, where γ is the reduced surface potential and can be calculated as [40]:
electrical double layer, steric interaction, bridging, hydrophobic and zeE0 ⎞
hydration-solvation interaction are six main types of particle–particle γ = tanh ⎛ ⎜ ⎟

4
⎝ KB T ⎠ (4)
interaction forces that can exist in the dispersion medium [26]. The sum
of the attractive (van der Waals, bridging, and hydrophobic forces) and where E0 is the potential on the surface. For a surface charge (E0) below
repulsive (the electrical double layer force, steric effect, and hydration 30 mV or ka < 5, the electrostatic potential can be calculated by linear
force) forces between individual particles govern the stability and ag- Poisson-Boltzmann approximation [39]:
gregation of particle dispersions. In general, to prepare a stable dis-
persion or to kinetically slow the aggregation, repulsive forces between VEL ≈ 2πεr ε ∘rE02 ke−kd (5)

560
S. Jafari Daghlian Sofla et al. Fuel 216 (2018) 559–571

Fig. 1. Ion distribution in electrical double layer theory (after [34]).

2.2. Surface charge of hydrophilic silica nanoparticle

It is well documented that when a nanoparticle is immersed in an


aqueous solution, the protonation/deprotonation capacity of the par-
ticle surface is a key parameter for the charge transfer between solvent
and particle. The relative basicity or acidity of the solvent and the
particle indicates the direction of protonic transfer [41,42]. Two pro-
tonation reactions (reaction R1 and R2) are suggested for silica nano-
particles (see also Fig. 2) [43].

SiOH ⇆ SiO− + H+ (R1)

SiOH2+ ⇆ SiOH + H+ (R2)

Due to the presence of a functional group containing oxygen on the


surface of silica nanoparticles, the solution’s pH can significantly affect
the charge of silica nanoparticles. Because oxygen can be protonated or
deprotonated to become charged [44]. It is impossible to directly Fig. 3. Charge of silica nanoparticles in different pHs (after [45]).
measure the Stern potential. Instead, the zeta potential (E), the poten-
tial at the shear plane close to the Stern plane, can be experimentally
measured and is often used as a measure of the surface potential [26]. the surface charge of silica nanoparticles is virtually unchanged for a
The surface charge of the hydrophilic silica nanoparticles as a function pH greater than 6, and its point of zero charge (pzc) occurs when the pH
of pH is well documented in the literature [45–47]. As shown in Fig. 3, is between 2 and 3.

2.3. Van der Waals interaction

Van der Waals forces are general attraction forces that occurs be-
tween molecules and consists of dipole–dipole, dipole-induced dipole,
and London (instantaneous induced dipole-induced dipole) forces [48].
For two spherical nanoparticles with radius R and center to center
distance of d, the Van der Waals attraction interaction energy can be
calculated as [49]:

E wdv A 2R2 2R2 d 2−4R2 ⎞ ⎤


= − 131 ⎡ 2 + 2 + ln ⎛ ⎜ ⎟

kB T 6kB T ⎢
⎣ d − 4R2 d ⎝ d
2
⎠⎥⎦ (6)

where A131 is the Hamaker constant and accounts for the interaction
between two nanoparticles of the same material (component 1) through
a solvent medium (component 3). For silica nanoparticles in water, the
effective Hamaker constant is reported as 36.2 × 10−22 J [50], kB is the
Boltzmann constant, and T is absolute temperature (K).
Classical DLVO theory generally fails to predict particle stability in
Fig. 2. Schematic diagram of side-binding model for silica dioxide [41].
the cases of strongly hydrophilic or hydrophobic particle suspensions

561
S. Jafari Daghlian Sofla et al. Fuel 216 (2018) 559–571

[51] and aqueous phases containing multivalent ions especially with nanoparticles was used without further modifications in part A and B
high ionic strength [52]. It is proposed that other interaction energies
(non-DLVO forces) [53,54], in addition to van der Waals and electro- 4.1.2. Aqueous solutions
static repulsion (DLVO forces), are the main reason for this short- In part A, solutions of different concentrations of inorganic salts in
coming. In the case of hydrophilic nanoparticles in an aqueous solution deionized water was used as an aqueous solution. Inorganic salts in-
with high ionic strength, hydration repulsion energy may be consider- cluding NaCl, Na2SO4, MgCl2 and MgSO4 with minimum purity of 99%
able [55,56]. The hydration force is a strong short-range repulsive force was purchased from Sigma-Aldrich.
that acts between polar surfaces separated by a thin layer of water, In part B, seawater was used as an aqueous solution. Seawater was
which decays quasi-exponentially with decay lengths of about 1 nm collected from the Grand Banks, offshore Newfoundland which its
[57]. Although there are some theoretical explanations about the origin composition is given in Table 1. Furthermore, different concentrations
and mechanism of hydration force (e.g. water-structuring models, of hydrochloric acid (HCl), a 6 M solution was used to form the H+
image-charge models, excluded-volume model, and dielectric-satura- protection layer around the nanoparticles.
tion models [58]), colloidal science researchers generally believe that
despite the proposed explanations, the origin of hydration repulsion 4.2. Methods
remains unclear [59]. The hydration force in the acidic solutions, due to
the penetration of protons into the surface, is not observed [60]. Based 4.2.1. Stability of hydrophilic silica nanoparticles in the presence of
on these studies, in the absence of any surfactant or polymer, the different ions
electrical double layer can be considered as a possible repulsion force In this part, the stability of silica nanoparticles in the presence of
between hydrophilic silica nanoparticles while van der Waals forces are different ions is evaluated. These experiments are designed to in-
responsible for the attraction of hydrophilic silica nanoparticles. Since vestigate the effect of type, and concentration of most common ions in
silica nanoparticles in the seawater are extremely unstable, van der seawater (Na+, Mg2+, Cl− and SO4−2) on the stability of silica nano-
Waals attraction outweighs the electrical double layer causing ag- particles. Hence, salt solutions with different concentrations of in-
glomeration of silica nanoparticles and thus instability. organic salts were prepared. Then, 0.05 wt% silica nanoparticles are
In this research, a new method to disperse hydrophilic silica nano- added to the prepared solution and ultra-sonicated for 15 min. To study
particles in seawater with the assistance of hydrochloric acid (HCl) is the effect of salt concentration on the stability of silica nanoparticles,
proposed and its effectiveness is experimentally evaluated. There is the salt concentration was changed from 0 to 0.5 wt%. This range of salt
numerous literature regarding the use of surfactants and polymers to concentration appears to be sufficient for Mg2+ and SO4-2 ions because
stabilize nanoparticles in high salinity solutions [61–63]. However, their concentrations in seawater are less than 0.5 wt%. However, for Cl-
using surfactants or polymers in EOR applications is challenging due to and Na+ ions, due to their higher concentrations in seawater, the sta-
factors such as the adsorption of surfactant to the rock surface during bility of silica nanoparticles in the presence of Cl- and Na+ ions is ex-
the injection and leading to possible formation damage in the case of amined up to the critical salt concentration of NaCl. Critical salt con-
polymers. Moreover, surfactants and polymers can lose their effec- centration is the maximum concentration of a salt in a nanoparticle
tiveness in the high salinity and high temperatures environments. solution in which the solution is still stable.
Conventional surfactants for EOR are also sensitive to hydrolysis The influence of cations on the stability of silica nanoparticles was
[64,65]. Surfactant and polymers are relatively expensive in compar- explored by selecting the salt solutions with the same anions but dif-
ison to using HCl. Using H+ protection method, which is obtained by ferent cations. For instance, in order to evaluate the effect of Na+ and
adding HCl in the solution, is a novel and simple solution to stabilize Mg2+ ions on the stability of silica nanoparticles, the stability of na-
silica nanoparticles in seawater and possibly other ionic solutions. It is noparticles in the solutions of NaCl and MgCl2, or Na2SO4 and MgSO4
known that by adding HCl, we are changing the pH which may have was compared. In these cases, the effect of anions on the stability of
further implications in EOR processes, i.e. possible mineral dissolution. nanoparticles in the both solutions is similar and the differences are due
These implications are not part of this work and are considered future to cations. Similarly, to examine the effect of anions, the salts with the
work. same cation but different anions were selected (NaCl− Na2SO4 and
MgCl2− MgSO4).
3. Experimental Section In all experiments, the procedure described in Supplementary ma-
terial is used to evaluate the stability of silica nanoparticles. According
The general description of experiments is shown in Fig. 4 where the to this procedure, in the first step, the samples are visually inspected.
experiments are divided into two parts: A and B. Part A investigates the The duration of this step depends on the required stability period for
effect of the most common ions in seawater on the stability of silica the practical application. In this study, the duration of the first step for
nanoparticles. Part A is performed to confirm previous findings (see all experiments was 48 h. The purpose of this step is to inspect and
Ref. [46,66]). Based on the results of these experiments, in part B, we record the cloudiness of solutions and possible sediments in the solu-
proposed and investigated a new method to stabilize silica nano- tions. At the end of this step, the solutions without obvious precipita-
particles in seawater using HCl. To the best of our ability, the use of HCl tion or severe cloudiness were selected for particle size measurement
as a means of stabilizing silica nanoparticles in seawater has not been and solutions with precipitation or severe cloudy solutions were con-
suggested previously. The effectiveness of HCl in stabilizing silica na- sidered unstable. It should be noted that the analysis of appearance
noparticles in seawater is examined by visual inspection, particle size, characteristics can be only used for screening of the obvious unstable
turbidity, and zeta potential measurements solutions with possible stable ones and the final decision about the
stability or instability of solutions cannot be made in this stage. In order
4. Material and methods to judge the stability or instability of solutions, first, based on the
purpose of experiments and its practical applications, the term of
4.1. Materials “stability” should be defined clearly. Then, the particle size measure-
ments should be conducted to see how the solutions meet the defined
4.1.1. Nanoparticles criteria.
The colloidal silica nanoparticle suspension (25 wt% silica nano- The “stability of nanoparticles” needs to be defined clearly for en-
particles in deionized water) with 5–35 nm diameter sizes was pur- hanced oil recovery purposes. To use nanoparticles in the EOR pro-
chased from US Research Nanomaterial, Inc. The purity of the 25 wt% cesses, the size of nanoparticles in the solution should be far less than
SiO2 nanoparticles is greater than 99.99%. The solution of silica the diameter of pore throats for nanoparticles to pass easily through the

562
S. Jafari Daghlian Sofla et al. Fuel 216 (2018) 559–571

Fig. 4. General description of experiments.

Table 1 4.2.2. Effect of hydrochloric acid on the stability of silica nanoparticles


Seawater composition. Several experiments were designed to assay the impact of hydro-
chloric acid on the stability of silica nanoparticles in seawater. The
Specification Concentration (ppm)
detail of experiments is provided in the Supplementary section. The
Sodium (Na ) +
1078 appearance characteristics of nanoparticle solution for the listed con-
Magnesium (Mg2+) 1284 centrations of HCl and silica nanoparticles were analyzed. Then, par-
Calcium (Ca2+) 412 ticle size, turbidity, zeta-potential and pH of solutions were measured.
Potassium (K+) 399
Chloride (Cl−) 19,353
In order to avoid any contamination during measurements, the so-
Sulfate (SO4−2) 2712 lutions after preparation were split into five samples. One sample was
Bicarbonate (HCO3−) 126 used for visual inspection and four of the samples were separated for
Other ions 114 the following experiments.
Total salinity 35,181

4.2.3. Particle size measurements


porous medium. Also, the nanoparticles must not agglomerate where After completing the visual inspection, the visually stable solutions
they may plug pore throats via log jamming. In the conventional oil were selected for particle size measurement. It should be noted that
reservoirs, the pore throat size is generally greater than 2 μm [67]. In DLS-based instruments like the Malvern zetasizer, measure the hydro-
coarse sandstone reservoirs the pore throat size is in the order of 4–7 μm dynamic size of the nanoparticles. This size is directly related to the
[67,68] and for carbonate reservoirs pore throats range from 3 to 10 μm diffusive speed of the particles in the solution. The diffusive speed can
[69,70]. Log-jamming or agglomeration is affected by particle hydro- be affected by altering the Debye length. Hence, the hydrodynamic
phobicity, particle retention, fluid ion strength, operating conditions diameter not only depends on the size of the particle “core”, but is also
and mineralogy of the formation [71]. The physics of log jamming is governed by the surface structure, the type of ions, and the ionic
not fully understood yet, hence, it is difficult to determine a definitive strength of the solution. At low ionic strengths, the thickness of the
size of nanoparticles to prevent log jamming. Based on the definition of electrical double layer or Debye length increases and consequently, the
nanoparticles, the particle that their size (at least in one dimension) is diffusion speed reduces, and a larger apparent dynamic diameter is
between 1 and 100 nm [72], we choose 100 nm, assuming that if the obtained.
size of nanoparticles in solution is less than 100 nm, the samples are still The size of nanoparticles in each sample was measured in three
“stable” and can be applied for EOR processes. However, if the size of separate runs at ambient temperature and pressure. For each run, the
nanoparticles is greater than 100 nm or the particle size measurement intensity-weighted mean diameter was recorded and the average of the
instrument did not pass the quality check, the solution is considered three runs is reported. The intensity-weighted average is very sensitive
unstable. The 100 nm criteria is widely used in literatures to define the to the presence of aggregates and contaminants.
stability of nanoparticles [78–80]. Normally, in the case of high poly-
dispersity or bad optical quality (severe cloudy samples) the instrument 4.2.4. pH measurement
does not pass the quality criteria. One of the four sets of samples was selected to measure the pH. The
In this research, the size of the nanoparticles is measured with a pH of the samples was measured with a Denver instrument up-5 pH-
Malvern Zetasizer Nano Series ZS instrument which calculates the size meter. In order to avoid any contamination, the pH probe was carefully
of the particles by measuring the Brownian motion of the particles in a cleaned with deionized water and calibrated using standard or “buffer”
sample using Dynamic Light Scattering (DLS). solutions before each measurement. For each nanoparticle solution,
three separate pH measurements were conducted for repeatability and
the average value is reported.

563
S. Jafari Daghlian Sofla et al. Fuel 216 (2018) 559–571

4.2.5. Turbidity measurement Na2SO4 is twice that of NaCl. Second, the charge of SO42− is also twice
To detect very small changes in the solution clarity, the turbidity of that of Cl−. These differences cause a shorter Debye length for nano-
the samples was measured using a Turbidimeter 2100Q StablCal particles in the Na2SO4 and consequently less electrostatic repulsion
Standards instrument. The turbidity-meter measures the cloudiness or and less of an energy barrier. On the other hand, there is a secondary
haziness of the fluid sample and normally is reported in Nephelometric minimum in the Na2SO4 solution especially for the solution of 0.5 wt%,
Turbidity Units (NTU). The instrument was calibrated by using the which it is not observed in the same concentration of NaCl. The sec-
StablCal® turbidity standards and the calibration was checked regularly ondary minimum shows a weak aggregation structure which can be
during measurements. The turbidity of each nanoparticle solution was broken with modest shear stress. In general, DLVO theory predicts the
measured four times for repeatability and the average value was taken. stable solution for both cases which have a great harmony with ex-
periments.
4.2.6. Zeta-potential measurement The size of silica nanoparticles in the presence of different con-
The zeta-potential of the samples was measured with a Malvern centrations of NaCl-MgCl2 and Na2SO4-MgSO4 is illustrated in
Zetasizer nano series ZS. Before each zeta-potential measurement, the Fig. 6a and b, respectively. It was observed that for the concentrations
zeta cell was rinsed carefully with alcohol and deionized water. In order higher than 0.1 wt% of MgCl2 and MgSO4, the size of the nanoparticles
to ensure the accuracy of the measured data, we ensured the instru- grew rapidly and the silica nanoparticles were not stable. However,
ment’s quality check passes for each nanoparticle solution and the they were stable in the solution of NaCl and Na2SO4 salts in the tested
measurements were repeated three times. concentrations. Again, this stability behavior of silica nanoparticles
should be sought in the cations of the salts. The results show that Mg2+
4.2.7. DLVO calculation as a divalent ion can destabilize nanoparticles more effectively than
The results of the experiments were compared with the well-known Na+ as a monovalent one. These results are consistent with the li-
DLVO theory. To calculate the DLVO theory, Eqs. (1–5) were used to terature [46,66]. In general, it appears that divalent ions are screening
calculate the electrostatic repulsion (EEL ) and Eq. (6) were used to the charge of silica nanoparticles more effectively than monovalent ions
calculate the Van der Waals attraction forces (E vdw ). The total potential where divalent (and multivalent) ions in the electrical double layer of
was calculated by using the following equation: nanoparticles can destabilize the nanoparticles in the solution.
The stability of silica nanoparticles in the presence of different
Totalpotential = EEL−Evdw (7)
concentrations of MgSO4 and MgCl2 is modeled by DLVO theory and
the results are shown in Fig. 7. For MgCl2 salt in 0.025 wt%, a repulsive
5. Results and discussion energy barrier is predicted which it goes zero in 0.5 wt%. Therefore, the
critical salt concentration is predicted to be 0.5 wt%. This means that
5.1. Effect of different ions the silica nanoparticles in the solutions that contain less than 0.05 wt%
MgCl2 are stable. However, they are unstable in the solutions of higher
The sizes of silica nanoparticles in the solutions containing different concentrations. For MgSO4 salt, the DLVO theory predicts an unstable
salt concentrations were investigated and the results were compared solution for all modeled concentrations. In general, DLVO theory has a
with DLVO theory. As depicted in the Supplementary section, for the great consistency with experimental data in the high salt concentra-
nanoparticle solutions containing up to 0.5 wt% of NaCl and Na2SO4 tions. However, while the DLVO theory was predicted to be an unstable
salts, no remarkable change in the size of the nanoparticles was ob- solution for all concentrations of MgSO4, experimental results show that
served and the silica nanoparticles were stable in both solutions. The for less than 0.1 wt% of MgSO4, nanoparticles are stable.
NaCl and Na2SO4 have the same cation (Na+) but different anions (Cl−
and SO42−, respectively). Since silica nanoparticles are negatively 5.2. Stability of silica nanoparticles in seawater and H+ protection theory
charged in a wide pH range, anions are repelled with the nanoparticles
while the cations are attracted. Hence, it is safe to assume that the As Keller et al. [73] studied, metal oxide nanoparticles are not stable
concentration of anions in the electrical double layer of silica nano- in seawater. To confirm the instability of silica nanoparticles in sea-
particles is far less than that of cations. Therefore, the stability of silica water, two solutions of 0.15 wt% of hydrophilic silica nanoparticles in
nanoparticles is mostly governed by the type and concentration of ca- seawater were prepared. One solution was ultra-sonicated for 15 min
tions. Since the type of cation in both cases was the same (Na+), the and another one was mixed by a magnetic stirrer for one hour. After 48
stability of silica nanoparticles was only dependent on the concentra- h, their picture was taken and the results are illustrated in Fig. 8. It was
tion of the cation. The concentration of Na+ in Na2SO4 is twice that of observed that silica nanoparticles aggregate rapidly when they are in-
NaCl (for the same salt concentrations), however, tested concentrations troduced in seawater and they sediment in a short time (less than 1 h).
are far less than the critical salt concentrations of Na+ and because of The stability of silica nanoparticles in seawater was also modeled
that, the size of nanoparticles is virtually the same for both cases. It is with DLVO theory. As shown in Fig. 9, DLVO theory predicts a small
worth to mention that critical Na+ concentration was obtained as high energy barrier to aggregation of silica nanoparticles in seawater.
as 6 wt%. However, there is also a relatively large secondary minimum which
The stability of silica nanoparticles in the presence of Na2SO4 and nanoparticle can trap in that. If we consider that the nanoparticles are
NaCl is modeled with DLVO theory and the results are illustrated in trapped in the secondary minimum, then their aggregation most
Fig. 5. DLVO calculations show that there is a relatively large “repulsive probably is reversible and they should disperse again with a magnetic
energy barrier” or “energy barrier” for the aggregation of nanoparticles stirrer or by using ultra-sonication. However, as mentioned, the ex-
in both salts solutions. In other words, the electrostatic repulsion in perimental results show that the proposed methods are not effective in
both cases of the calculated concentrations is stronger than attractive stabilizing nanoparticles and breaking the aggregations.
Van der Waals force in a wide range of distances between particles. We propose a new method to stabilize silica nanoparticles in sea-
Hence, when the nanoparticles approach each other, the energy barrier water and call it “H+ protection”. As discussed in the part A of this
prevents them from sticking and aggregating. research, the stability of silica nanoparticles is mostly governed by the
The value of repulsive energy barrier for the NaCl solution in all presence of multivalent ions in the electrical double layer of nano-
concentrations is higher than Na2SO4. The repulsive energy barrier of particles. It was observed that even small concentrations of multivalent
NaCl is about 5× higher than Na2SO4 in 0.1 wt% and it increases to ions can destabilize silica nanoparticles. Hence, to stabilize silica na-
around 9× in 0.5 wt%. There are two main reasons for this trend. First, noparticles in a solution, the presence of multivalent ions in the elec-
in the same concentration of both salts, the concentration of Na+ in the trical double layer of nanoparticles should be limited. Tombácz, and

564
S. Jafari Daghlian Sofla et al. Fuel 216 (2018) 559–571

(a) Na2SO4 (b) NaCl


Fig. 5. DLVO calculation for silica nanoparticles in the presence of (a) Na2SO4 and (b) NaCl.

Fig. 6. Silica nanoparticle size in the presence of (a) NaCl and MgCl2 ions, and (b) Na2SO4 and MgSO4 ions.

Szekeres reported that H+ has a very high affinity to neutralize nega- We now demonstrate that H+ ions can stabilize silica nanoparticles
tively charged surfaces [74]. Hence, H+ and multivalent ions in sea- in seawater. The positively charged cations including H+, Na+, Ca2+,
water (mostly Ca2+ and Mg2+) compete to be present in the electrical and Mg2+, H+ ions are attracted and want to concentrate in the elec-
double layer of silica nanoparticles. More recently, it is reported that trical double layer of silica nanoparticles. We believe that with the
HCl can influence the surface chemistry iron oxide nanoparticles [75]. sufficient H+ ions in solution, the H+ ions can form a protective layer

Fig. 7. DLVO calculation for silica nanoparticles in the presence of (a) MgCl2 and (b) MgSO4.

565
S. Jafari Daghlian Sofla et al. Fuel 216 (2018) 559–571

around silica nanoparticles to reduce the concentration of multivalent


ions in the electrical double layer of silica nanoparticles. This is re-
presented in Fig. 10. First, we examine solution preparation of the H+
protection and its effect on the stability of silica nanoparticles in sea-
water, after which we examine the effect of concentration of the silica
nanoparticles and HCl used to H+ protect and stabilize the silica na-
noparticles in seawater.

5.2.1. Preparation procedure


To determine whether or not the preparation procedure influences
the stability of silica nanoparticles in seawater, solutions are prepared
based on two different procedures. As shown in Fig. 11, in the first
procedure, hydrochloric acid (less than 10 µL) is initially diluted with a
small amount of seawater (about 1000 µL) to form an acidic solution.
Then, silica nanoparticles were added into the solution. In the last step,
Silica nanoparticles the prepared nanoparticle dispersion was further diluted with the de-
Fig. 8. Silica nanoparticles in seawater (without HCl) after mixing with magnetic stirrer sired amount of seawater to achieve the desired concentration. In the
(A) and ultra-sonication (B). second procedure, the silica nanoparticles were added to the seawater
and then, hydrochloric acid was added to the solution. All solutions,
after preparation, are mixed with a magnetic stirrer for 15 min. Two
samples with described procedures were prepared. Table 2 provides
detailed information about the prepared samples. The results show that
the size of the silica nanoparticles in the sample A after 48 h was still
around 20 nm showing that the silica nanoparticles in this solution are
stable. However, the average size of silica nanoparticles in the sample B
after 48 h was 436.5 ± 4 nm showing that the nanoparticles are ag-
gregating.
In the sample A, by initially diluting the HCl with a small volume of
seawater, an extremely acidic solution was created. By adding the na-
noparticles to this solution, there were sufficient free H+ ions in the
solution to accumulate in the electrical double layer of nanoparticles.
Hence, the H+ protection layer formed quickly and with a further ad-
dition of seawater, the protection layer was still stable. Preparation
method B is when the nanoparticles were directly added to the seawater
(neutral pH) and then the HCl was added. The multivalent ions were
first to occupy the nanoparticle electrical double layer and when the
HCl was added, the free H+ ions could not invade the electric double
Fig. 9. DLVO calculation of silica nanoparticles in seawater.
layer to form an H+ protection layer. Hence, multivalent ions that ac-
cumulated in the electrical double layer could destabilize the silica
nanoparticles. Adding the H+ ion to this solution could not remove the
attracted multivalent ions from the electrical double layer of silica
nanoparticles and the nanoparticles aggregated in preparation method
B.
The pH of solution is another indication that there is formation of
the H+ protection layer. The pH of preparation method A is higher than
preparation method B. This implies that in sample A, some of the free
H+ ions were attracted in the electrical double layer around nano-
particles and there was less free H+ for the pH meter to detect.

5.3. Effect of HCl on the stability of silica nanoparticles in seawater

The effect of HCl on the stability of silica nanoparticles in the so-


lutions that contain different concentrations of nanoparticles and HCl
was investigated. All samples were prepared according to sample pre-
paration method A and the stability of nanoparticles was tested based
on the procedure described in the Supplementary section. The samples
were visually inspected after 48 h. Figs. 12–15 show the visual status of
different concentrations of silica nanoparticles in the solutions of 0.025,
0.0076, 0.003 and 0.002 wt% HCl after 48 h, respectively. As shown in
the figures, at low concentrations of silica nanoparticles, the solutions
are clear and there are no sediments or cloudiness. However, with in-
creasing concentrations of nanoparticles, the solutions become cloudy.
Fig. 10. H+ protection layer around silica nanoparticles.
The change in clarity also depends on the concentration of HCl. De-
creasing the concentration of HCl shows that the samples become

566
S. Jafari Daghlian Sofla et al. Fuel 216 (2018) 559–571

Fig. 11. Two different preparation procedures


(A and B) to prepare H+ nanoparticle seawater
solutions.

Table 2
The composition of two prepared samples and their stability.

Preparation method Nanoparticle concentration Seawater volume HCl(aq)-6 M HCl Concentration (wt Measured pH Visual Particle size after 48 h
(wt%) (ml) volume (ml) %) stability? (nm)

A 0.15 100 0.01 0.0022 6.70 ± 0.05 Stable 20.02 ± 0.8


B 0.15 100 0.01 0.0022 6.58 ± 0.05 Unstable 436.5 ± 4.0

cloudy in lower concentrations of silica nanoparticles. For instance, in does not change significantly by increasing the concentration of silica
Fig. 12, for samples containing 0.025 wt% HCl, a little cloudiness is nanoparticles. Furthermore, it is observed that for samples containing
observed for samples with nanoparticle concentrations of 0.4, 0.45 and 0.0076, 0.003 and 0.002 wt% HCl, the cloudiness is starting from
0.5 wt%. However, in Fig. 13, for samples with 0.0076 wt% HCl, the samples 6, 6 and 5, respectively. The turbidity of these samples is
obvious cloudy condition begins at sample 7 (sample with 0.3 wt% si- pointed in Fig. 16. For the samples containing 0.003 and 0.002 wt%
lica nanoparticles) and increases in Samples 8–11. HCl, the observations are confirmed with the turbidity measurements.
Turbidity is a better measure of cloudiness than visual inspection. However, in the sample with 0.0076 wt%, the cloudiness is starting
The turbidity of the prepared samples was measured and the results are from sample number 6 which is not distinguishable with visual ob-
reported in Fig. 16. As expected, the turbidity of samples is directly servations. Hence, in order to reduce the possible human errors, re-
related to the concentration of nanoparticles and reversely to the HCl placing the visual observations with the turbidity measurements can be
concentration. For the samples containing 0.025 wt% HCl, the turbidity considered as an option. It is worth to mention that the accuracy of

Fig. 12. The visual status of silica nanoparticles with different concentrations (the concentrations are written on the top of each sample) in the solution of seawater with 0.025 wt% HCl
after 48 h.

567
S. Jafari Daghlian Sofla et al. Fuel 216 (2018) 559–571

Fig. 13. The visual status of silica nanoparticles with different concentrations (the concentrations are written on the top of each sample) in the solution of seawater with 0.0076 wt% HCl
after 48 h.

measured turbidity is ± 0.05 NTU. respectively. These points are illustrated with arrows in Fig. 17. For the
In the second step of stability analysis, the average size of nano- solutions containing 0.025 wt% HCl, the “break point concentration” is
particles is measured. The change in the average size of nanoparticles higher than 0.5 wt% of silica nanoparticles. The stability analysis ex-
by altering the concentration of nanoparticles in the solutions of four periments are ended at the “break point concentrations”. After this
distinct HCl concentrations including 0.025, 0.0076, 0.003 and concentration, the samples not only are not stable, but also the nano-
0.002 wt% HCl is shown in Fig. 17. The error in the measurement of particles may sediment during operation.
particle size is ± 0.8 nm. As illustrated in the figure, for all four HCl Based on the stability definition, as shown in Fig. 18, all samples
concentrations, with increasing the concentration of silica nano- with 0.025 wt% HCl, samples 1–7 for 0.0076 wt% HCl and samples 1–6
particles from 0.03 wt%, the average size of nanoparticles exponentially for solutions with 0.003 and 0.002 wt% HCl are considered as a stable
increases. However, the rate of increasing is different for the samples sample. It is worth to mention that maximum oil recovery can be ob-
with different HCl concentrations. For instance, in the solution of tained using solutions in which the concentration of nanoparticles
0.025 wt% HCl, increasing in the size of nanoparticles is very gentle, varies between 0.05 and 0.1 wt% [13,76,77]. These concentrations are
however, for 0.003 and 0.002 wt% of HCl, the relatively sharp in- obtained using core displacement experiments.
creasing in the size of nanoparticles is observed. Generally, with de- In the solutions with constant HCl concentration, at low con-
creasing the HCl concentrations, the rate of increasing in the size of centrations of silica nanoparticles, there are sufficient free H+ ions in
silica nanoparticles, increases. The exponentially increasing in the size the solution to accumulate in the electrical double layer of each na-
of nanoparticles is ending at the “break point concentration” where noparticle. Hence, the H+ ions form an effective protection layer and
with further increasing of the concentration of nanoparticles, the size of prevent nanoparticles to aggregate. However, by increasing the con-
nanoparticles is jumping to very high values. The “break point con- centration of silica nanoparticles, constant number of H+ ions should
centration” for the solutions of 0.0076, 0.003 and 0.002 wt% HCl are be divided into more numbers of silica nanoparticles. Hence, the con-
obtained as 0.45, 0.25 and 0.25 wt% of silica nanoparticles, centration of H+ ions in the electrical double layer of silica

Fig. 14. The visual status of silica nanoparticles with different concentrations (the concentrations are written on the top of each sample) in the solution of seawater with 0.003 wt% HCl
after 48 h.

568
S. Jafari Daghlian Sofla et al. Fuel 216 (2018) 559–571

Fig. 15. The visual status of silica nanoparticles with different concentrations (the concentrations are written on the top of each sample) in the solution of seawater with 0.002 wt%.

Fig. 18. The samples with less than 100 nm size.

Fig. 16. The turbidity of samples as a function of the concentration of nanoparticles and
HCl.

Fig. 19. Change in the pH of samples with altering the concentration of silica nano-
Fig. 17. The size of silica nanoparticles in the solutions with different HCl and silica particles in different concentrations of HCl.
nanoparticles – The arrows are showing the “break point concentrations”.

for low concentrations and it decreases with increasing the concentra-


nanoparticles, with increasing the concentration of nanoparticles, de- tion of silica nanoparticles till reach plateau at high concentrations of
creases. In this case, the multivalent ions had an opportunity to present silica nanoparticles. This trend shows that with introducing the silica
in the electrical double layer of nanoparticles. Therefore, the H+ ions nanoparticles in a solution, the free H+ ions are accumulating around
could not completely prevent the nanoparticles from aggregation and silica nanoparticles and the concentration of free H+ ions in the solu-
the size of nanoparticles (aggregation of nanoparticles) increases. This tion decreases. With further increasing the concentration of silica na-
is also can be observed in the change of pH in the solution. In fact, the noparticles, the concentration of H+ ions in the solution is not enough
pH-meter is measuring the number (concentration) of free H+ ions in to form a complete H+ protection layer around all nanoparticles. In this
the solution. As illustrated in Fig. 19, for four samples of HCl, without point, nanoparticles have adsorbed almost all added H+ ions and the
silica nanoparticles, the pH is less than when the silica nanoparticles are pH of solution moves toward the neutral pH and do not change any-
introduced into the solution. With increasing the concentration of silica more.
nanoparticles, the pH is raising. The rate of increasing in the pH is sharp Zeta-potential measurement is another useful method to understand

569
S. Jafari Daghlian Sofla et al. Fuel 216 (2018) 559–571

Fig. 20. (a) Zeta-potential of silica nanoparticles in the samples with different HCl and nanoparticle concentrations, and (b) Zeta-potential of nanoparticles as a function of pH.

the mechanism of dispersing silica nanoparticles in seawater with the of nanoparticles in seawater is increasing by raising the concentration
assist of HCl. As shown in Fig. 20a, with increasing the concentration of of nanoparticles and decreasing the HCl concentration.
HCl in the solutions, the zeta-potential of silica nanoparticles from
negative values is moving toward zero, and then positive values. On the Acknowledgment
other hand, except for the samples with 0.025 wt% HCl, the zeta-po-
tential of samples has a decreasing trend. Regardless of HCl or silica The authors appreciate the financial support from the Hibernia
nanoparticle concentration, the absolute value of the zeta-potential in Management and Development Corporation (HMDC), Chevron Canada,
the most samples is less than 10 mV and for some samples is even less InnovateNL, and the Natural Sciences and Engineering Research
than 4 mV. At this ranges of zeta-potentials, due to the weaker elec- Council of Canada (NSERC), without which this work could not have
trostatic repulsion force, it is expected that the nanoparticles aggregate been performed.
rapidly. However, as discussed before, the nanoparticles are stable in
these ranges after 48 h. It appears that some H+ ions are screening the Appendix A. Supplementary data
charge of silica nanoparticles with firmly sticking on the surface of si-
lica nanoparticles between the surface of nanoparticles and the shear Supplementary data associated with this article can be found, in the
plane which are also the part of H+ protection layer. Furthermore, as online version, at http://dx.doi.org/10.1016/j.fuel.2017.11.091.
shown in Fig. 20b, the zeta-potential of silica nanoparticles is not
completely the function of pH. For instance, for the pH range 6–7, the References
zeta-potential of silica nanoparticles can be any value between −11 mV
and −4 mV. It appears that in the presence of sufficient free H+ ions in [1] Bell AT. The impact of nanoscience on heterogeneous catalysis. Science
the solution, nanoparticles are adsorbing H+ ions and the zeta-potential 2003;299(5613):1688–91.
[2] Perez JM. Iron oxide nanoparticles: hidden talent. Nat Nanotechnol
of nanoparticles goes to near zero or positive values. However, when 2007;2(9):535–6.
the concentration of H+ ions in the solution is not enough to screen all [3] Jafari S, et al. Experimental investigation of heavy oil recovery by continuous/WAG
charge of nanoparticles, only a portion of the charge of nanoparticles is injection of CO2 saturated with silica nanoparticles. Int J Oil Gas Coal Technol
2015;9(2):169–79.
neutralized with the H+ ions and the final zeta-potential of nano- [4] Munshi A, et al. Effect of nanoparticle size on sessile droplet contact angle. J Appl
particles remain negative. It is worth to mention that the zeta-potential Phys 2008;103(8):084315.
of silica nanoparticles in DI-water and at neutral pH was [5] Chávez-Miyauchi TsE, Firoozabadi A, Fuller GG, et al. Nonmonotonic elasticity of
the crude oil–brine interface in relation to improved oil recovery. Langmuir
−22 ± 0.5 mV. 2016;32(9):2192–8.
[6] Davies R, et al. Induced seismicity and hydraulic fracturing for the recovery of
hydrocarbons. Mar Pet Geol 2013;45:171–85.
6. Conclusions [7] Rosen MJ, et al. Ultralow interfacial tension for enhanced oil recovery at very low
surfactant concentrations. Langmuir 2005;21(9):3749–56.
[8] Ahmadi MA, Shadizadeh SR. Adsorption of novel nonionic surfactant and particles
In this paper, we proposed a new method to stabilize silica nano- mixture in carbonates: enhanced oil recovery implication. Energy Fuels
particles in seawater. Silica nanoparticles are extremely unstable in 2012;26(8):4655–63.
seawater and they aggregate as soon as are introduced in seawater. [9] Fan H, Striolo A. Nanoparticle effects on the water-oil interfacial tension. Phys Rev
E 2012;86(5):051610.
Ultra-sonicating or mixing the solution is not effective in stabilizing
[10] Maghzi A, et al. Monitoring wettability alteration by silica nanoparticles during
nanoparticles and breaking the aggregations. Based on the stability of water flooding to heavy oils in five-spot systems: a pore-level investigation. Exp
silica nanoparticles in the presence of various ions in the aqueous so- Therm Fluid Sci 2012;40:168–76.
[11] Giraldo J, et al. Wettability alteration of sandstone cores by alumina-based nano-
lution, using hydrochloric acid to stabilize silica nanoparticles in sea-
fluids. Energy Fuels 2013;27(7):3659–65.
water is proposed and its effectiveness is tested. It was observed that [12] Karimi A, et al. Wettability alteration in carbonates using zirconium oxide nano-
hydrochloric acid can effectively stabilize silica nanoparticles in the fluids: EOR implications. Energy Fuels 2012;26(2):1028–36.
seawater. [13] Al-Anssari S, et al. Effect of temperature and SiO2 nanoparticle size on wettability
alteration of oil-wet calcite. Fuel 2017;206:34–42.
The “H+ Protection” is the possible mechanism for stabilizing of [14] Somasundaran P. Encyclopedia of surface and colloid science. CRC press; 2006.
silica nanoparticles in seawater by using HCl. H+ ions limit the pre- vol. 2.
sence of multivalent ions in the electrical double layer of nanoparticles [15] Zhang H, Nikolov A, Wasan D. Enhanced oil recovery (EOR) using nanoparticle
dispersions: underlying mechanism and imbibition experiments. Energy Fuels
and prevent them from aggregation. We observed that the average size

570
S. Jafari Daghlian Sofla et al. Fuel 216 (2018) 559–571

2014;28(5):3002–9. 2011.
[16] Taborda EA, et al. Effect of nanoparticles/nanofluids on the rheology of heavy [49] Saunders S, et al. Total interaction energy model to predict nanoparticle dis-
crude oil and its mobility on porous media at reservoir conditions. Fuel persability in CO2-expanded solvents. Comput Aided Chem Eng 2010;28:1651–6.
2016;184:222–32. [50] Ueno K, et al. Colloidal stability of bare and polymer-grafted silica nanoparticles in
[17] Ma X, et al. Preparation and characterization of silica/polyamide-imide nano- ionic liquids. Langmuir 2008;24(10):5253–9.
composite thin films. Nanoscale Res Lett 2010;5(11):1846–51. [51] Molina-Bolivar J, Ortega-Vinuesa J. How proteins stabilize colloidal particles by
[18] Torsater O, Li S, Hendraningrat L.. A coreflood investigation of nanofluid enhanced means of hydration forces. Langmuir 1999;15(8):2644–53.
oil recovery in low-medium permeability Berea sandstone. In: SPE international [52] Szilagyi I, Sadeghpour A, Borkovec M. Destabilization of colloidal suspensions by
symposium on oilfield chemistry: Society of Petroleum Engineers; 2013. multivalent ions and polyelectrolytes: from screening to overcharging. Langmuir
[19] Li S., Hendraningrat L, Torsaeter O. Improved Oil Recovery by Hydrophilic Silica 2012;28(15):6211–5.
Nanoparticles Suspension: 2 Phase Flow Experimental Studies. In: IPTC 2013: [53] Bizmark N, Ioannidis MA. Effects of ionic strength on the colloidal stability and
International Petroleum Technology Conference; 2013. interfacial assembly of hydrophobic ethyl cellulose nanoparticles. Langmuir
[20] Hendraningrat L, Torsæter O. Metal oxide-based nanoparticles: revealing their po- 2015;31(34):9282–9.
tential to enhance oil recovery in different wettability systems. Appl Nanosci [54] Bizmark N, Ioannidis MA, Henneke DE. Irreversible adsorption-driven assembly of
2015;5(2):181–99. nanoparticles at fluid interfaces revealed by a dynamic surface tension probe.
[21] Mohajeri M, Hemmati M, Shekarabi AS. An experimental study on using a nano- Langmuir 2014;30(3):710–7.
surfactant in an EOR process of heavy oil in a fractured micromodel. J Petrol Sci [55] Grasso D, et al. A review of non-DLVO interactions in environmental colloidal
Eng 2015;126:162–73. systems. Rev Environ Sci Biotechnol 2002;1(1):17–38.
[22] Gahrooei HRE, Ghazanfari MH. Application of a water based nanofluid for wett- [56] Liu HH, et al. Effect of hydration repulsion on nanoparticle agglomeration eval-
ability alteration of sandstone reservoir rocks to preferentially gas wetting condi- uated via a constant number Monte-Carlo simulation. Nanotechnology
tion. J Mol Liq 2017;232:351–60. 2015;26(4):045708.
[23] Guo F, Aryana S. An experimental investigation of nanoparticle-stabilized CO2 foam [57] Ruths M, Israelachvili JN. Surface forces and nanorheology of molecularly thin
used in enhanced oil recovery. Fuel 2016;186:430–42. films. Nanotribology and nanomechanics. Springer; 2008. p. 417–515.
[24] Maghzi A, et al. The impact of silica nanoparticles on the performance of polymer [58] Kralchevsky PA, Danov KD, Basheva ES. Hydration force due to the reduced
solution in presence of salts in polymer flooding for heavy oil recovery. Fuel screening of the electrostatic repulsion in few-nanometer-thick films. Curr Opin
2014;123:123–32. Colloid Interface Sci 2011;16(6):517–24.
[25] Zargartalebi M, Kharrat R, Barati N. Enhancement of surfactant flooding perfor- [59] Brzozowska A, et al. On the stability of the polymer brushes formed by adsorption
mance by the use of silica nanoparticles. Fuel 2015;143:21–7. of ionomer complexes on hydrophilic and hydrophobic surfaces. J Colloid Interface
[26] Somasundaran P. Colloid systems and interfaces stability of dispersions through Sci 2011;353(2):380–91.
polymer and surfactant adsorption. Boca Raton, FL: CRC Press; 2009. [60] Donaldson EC, Alam W. Wettability. Elsevier; 2013.
[27] Vasiliev PO, et al. Colloidal aspects relating to direct incorporation of TiO 2 na- [61] de Lara LS, Rigo VA, Miranda CR. The stability and interfacial properties of func-
noparticles into mesoporous spheres by an aerosol-assisted process. J Colloid tionalized silica nanoparticles dispersed in brine studied by molecular dynamics.
Interface Sci 2008;319(1):144–51. Eur Phys J B 2015;88:261.
[28] Derjaguin B, Landau L. Theory of the stability of strongly charged lyophobic sols [62] Bagaria HG, et al. Stabilization of iron oxide nanoparticles in high sodium and
and of the adhesion of strongly charged particles in solutions of electrolytes. Acta calcium brine at high temperatures with adsorbed sulfonated copolymers. Langmuir
physicochim URSS 1941;14(6):633–62. 2013;29(10):3195–206.
[29] Verwey EJW, Overbeek JTG, Overbeek JTG, Theory of the stability of lyophobic [63] Nap RJ, Park SH, Szleifer I. On the stability of nanoparticles coated with poly-
colloids: Courier Corporation; 1999. electrolytes in high salinity solutions. J Polym Sci Part B Polym Phys
[30] von Helmholtz H. Ueber einige Gesetze der Vertheilung elektrischer Ströme in 2014;52(24):1689–99.
körperlichen Leitern mit Anwendung auf die thierisch-elektrischen Versuche. Ann [64] Olajire AA. Review of ASP EOR (alkaline surfactant polymer enhanced oil recovery)
Phys Chem 1853;89:211–33. technology in the petroleum industry: prospects and challenges. Energy
[31] Gipson K, et al. The influence of synthesis parameters on particle size and photo- 2014;77:963–82.
luminescence characteristics of ligand capped Tb3+: LaF3. Polymers [65] Ehtesabi H, et al. Enhanced heavy oil recovery in sandstone cores using TiO2 na-
2011;3(4):2039–52. nofluids. Energy Fuels 2013;28(1):423–30.
[32] Gouy G. Constitution of the electric charge at the surface of an electrolyte. J phys [66] Wang D, et al. Bridging interactions and selective nanoparticle aggregation medi-
1910;9(4):457–67. ated by monovalent cations. ACS Nano 2010;5(1):530–6.
[33] Napper DH. Polymeric stabilization of colloidal dispersions. Academic Press; 1983. [67] Nelson PH. Pore-throat sizes in sandstones, tight sandstones, and shales. AAPG Bull
vol. 3. 2009;93(3):329–40.
[34] Lin K-W, et al., Micro/Nano Lithography Sub-20nm node photomask cleaning en- [68] Lindquist WB, et al. Pore and throat size distributions measured from synchrotron
hanced by controlling zeta potential. X-ray tomographic images of Fontainebleau sandstones. J Geophys Res Solid Earth
[35] Smith AM, Lee AA, Perkin S. The electrostatic screening length in concentrated 2000;105(B9):21509–27.
electrolytes increases with concentration. J Phys Chem Lett 2016;7(12):2157–63. [69] Dominguez G. Carbonate reservoir characterization: a geologic-engineering ana-
[36] Bukar N, et al. Influence of the Debye length on the interaction of a small molecule- lysis. Elsevier; 1992. vol. 30.
modified Au nanoparticle with a surface-bound bioreceptor. Chem Commun [70] Ausbrooks R., et al. Pore-size distributions in vuggy carbonates from core images,
2014;50(38):4947–50. NMR, and capillary pressure. In: SPE annual technical conference and exhibition:
[37] French RA, et al. Influence of ionic strength, pH, and cation valence on aggregation Society of Petroleum Engineers; 1999.
kinetics of titanium dioxide nanoparticles. Environ Sci Technol 2009;43(5):1354–9. [71] Peng B, et al. A review of nanomaterials for nanofluid enhanced oil recovery. RSC
[38] Hiemenz P, Rajagopalan R. Electrophoresis and other electrokinetic phenomena. Adv 2017;7(51):32246–54.
New York.: Marcel Dekker Inc.; 1977. p. 452–487. [72] Horikoshi S, Serpone N. Introduction to nanoparticles. Microwaves in nanoparticle
[39] Zeng Y. Colloidal dispersions under slit-pore confinement. Springer Science; 2012. synthesis: fundamentals and applications 2013:1–24.
[40] Elimelech M, Gregory J, Jia X. Particle deposition and aggregation: measurement, [73] Keller AA, et al. Stability and aggregation of metal oxide nanoparticles in natural
modelling and simulation. Butterworth-Heinemann; 2013. aqueous matrices. Environ Sci Technol 2010;44(6):1962–7.
[41] Birdi K. Handbook of surface and colloid chemistry. CRC Press; 2015. [74] Tombácz E, Szekeres M. Surface charge heterogeneity of kaolinite in aqueous sus-
[42] Hunter RJ. Zeta potential in colloid science. principles and applications. Academic pension in comparison with montmorillonite. Appl Clay Sci 2006;34(1):105–24.
press; 2013. vol. 2. [75] Korpany KV, et al. Iron oxide surface chemistry: effect of chemical structure on
[43] Barisik M, et al. Size dependent surface charge properties of silica nanoparticles. J binding in benzoic acid and catechol derivatives. Langmuir 2017;33(12):3000–13.
Phys Chem C 2014;118(4):1836–42. [76] Hendraningrat L, Li S, Torsæter O. A coreflood investigation of nanofluid enhanced
[44] Butt H-J, Graf K, Kappl M. Physics and chemistry of interfaces. John Wiley & Sons; oil recovery. J Petrol Sci Eng 2013;111:128–38.
2006. [77] Bai M, et al. Studies of injection parameters for chemical flooding in carbonate
[45] Mandel K, et al. Surfactant free superparamagnetic iron oxide nanoparticles for reservoirs. Renew Sustain Energy Rev 2017;75:1464–71.
stable ferrofluids in physiological solutions. Chem Commun 2015;51(14):2863–6. [78] Royal Society and Royal Academy of Engineering. Nanoscience and nanotechnol-
[46] Metin CO, et al. Stability of aqueous silica nanoparticle dispersions. J Nanopart Res ogies: opportunities and uncertainties; 2004.
2011;13(2):839–50. [79] De Jong M, Wim H, Borm Paul JA. Drug delivery and nanoparticles: applications
[47] Choi H-H, Park J, Singh R. Nanosized CuO encapsulated silica particles using an and hazards. Int J Nanomed 2008;3(2).
electrochemical deposition coating. Electrochem Solid-State Lett 2004;7(1):C10–2. [80] Ding Yaobo, Riediker Michael. A system to assess the stability of airborne nano-
[48] Israelachvili JN. Intermolecular and surface forces. Revised 3rd ed. Academic press; particle agglomerates under aerodynamic shear. J Aerosol Sci 2015;88:98–108.

571

S-ar putea să vă placă și