SlideShare a Scribd company logo
1 of 10
Download to read offline
Applied Clay Science 212 (2021) 106212
Available online 22 July 2021
0169-1317/© 2021 Elsevier B.V. All rights reserved.
Molecular dynamics study on the zeta potential and shear plane of
montmorillonite in NaCl solutions
Huafu Pei , Siqi Zhang *
School of Civil Engineering, State Key Lab of Coastal and Offshore Engineering, Dalian University of Technology, Dalian 116026, China
A R T I C L E I N F O
Keywords:
Clay mineral
Zeta potential
Shear plane
Electrical double layer
Stern model
Molecular dynamics
A B S T R A C T
Zeta potential and the position of the shear plane are key physical properties characterizing the behavior of clay
minerals in the colloidal system, and have been applied in many fields such as electroosmotic consolidation and
electro-kinetic decontamination. In the past decades, numerous studies have been conducted on measuring and
calculating the zeta potential. Nevertheless, few researchers have been reported to achieve a systematic un­
derstanding and predicting the zeta potential and the shear plane's position, especially for clay particles. This
paper provided a molecular dynamics (MD)-based method to determine the zeta potential and shear layer
thickness simultaneously to fill the gap. In the paper, the structure of the electrical double layer (EDL) was
investigated for montmorillonite mesopore containing NaCl electrolyte in the concentration of 0.20–1.30 mol/L.
The density profiles of ion species were well predicted by the Stern model, combining the Stern potential's
determination. The calculated zeta potential based on the electroosmotic velocity profile in nonequilibrium
molecular dynamics (NEMD) simulation was improved by introducing the slip length and was found to be closely
comparable to the experimental values. Furthermore, the results confirmed that the shear plane cannot be
observed from the electroosmotic velocity profile and self-diffusion coefficient of species in the MD simulation.
The position of the zeta potential was determined by the Stern model and triple-layer model (TLM), showing a
certain distance from the Stern plane. The zeta position was found that has a linear relationship with the Debye
length and linearly depends on the ionic strength in the log scale, in agreement with previous investigations. The
findings provided a systematic insight into the electrical double layer structure, zeta potential and shear plane for
montmorillonite.
1. Introduction
As negatively charged particles, clay minerals are ubiquitous in the
natural world. The clay-water interface strongly attracts cation species
and forms the electrical double layer (EDL) structure (Lyklema, 1995).
The structure greatly influences not only the electrochemical properties
of clay minerals that characterize the efficiency of electroosmosis and
electro-kinetic decontamination of clay (Shang, 1997; Ou et al., 2015)
but also the behavior of aggregation, complexation and sedimentation
(Tombácz and Szekeres, 2004; Peng et al., 2015). Serval EDL models
including the Gouy-Chapman model, Stern-Gouy-Chapman model
(Stern model in this paper for short) (Shang et al., 1994) and triple-layer
model (TLM) (Leroy and Revil, 2004; Sverjensky, 2005) have been
developed to describe the ions and electric potential distribution in the
colloidal system. For the Stern model, the EDL consists of the Stern layer
that the electrostatic potential within decreases linearly with distance
and diffuse layer (Gouy layer) in which the potential follows the
Poisson-Boltzmann (PB) equation. The Stern layer contains two kinds of
adsorbate species including inner-sphere surface complexes (ISSC) and
outer-sphere surface complexes (OSSC). In the TLM, the electrostatic
potential drops linearly among the 0-, β- and d-planes controlled by two
capacitance factors, while the potential is modeled with the PB equation
beyond the d-planes. The detailed introductions of the Stern model and
TLM are shown in Supporting Information.
Zeta potential at the shear plane is a significant macroscopic mea­
surement used to estimate the stability of colloidal systems and
numerous electro-kinetic applications (Şans et al., 2017; Zhao et al.,
2017). The electric potential at the supposed shear plane that is the
boundary between an immobile layer strongly adsorbed by the particle
surface and the mobile fluid. When investigating the nanoparticle
agglomeration, zeta potential can evaluate the critical coagulation
concentration (Hunter, 1981). The potential at the shear plane is used to
* Corresponding author.
E-mail address: zhangsiqi9315@mail.dlut.edu.cn (S. Zhang).
Contents lists available at ScienceDirect
Applied Clay Science
journal homepage: www.elsevier.com/locate/clay
https://doi.org/10.1016/j.clay.2021.106212
Received 25 January 2021; Received in revised form 6 July 2021; Accepted 13 July 2021
Applied Clay Science 212 (2021) 106212
2
study the transport properties of clay minerals (Leroy et al., 2008). The
shear plane is also an essential parameter that characterizes the hy­
drodynamic properties of colloidal systems. The shear layer (the space
between the clay surface and shear plane) greatly impacts the hydro­
dynamic motion of pore fluid and suspended particles (Xu, 1998). As a
boundary condition, the shear plane is used to calculate the electroos­
motic flow velocity (Wang et al., 2006; Vasu and De, 2010). In soil
mechanics, water in the shear layer is regarded as tightly bound water
(Li, 2016). Compared to several measurement techniques such as elec­
trophoresis, electroosmosis and electroacoustics of the zeta potential
determination (Greenwood, 2003), to date, the measurement of the
shear layer thickness mainly follows two methods. The primary method
is based on the PB equation or modified PB equation. The potential
profile in the EDL is obtained, and then the shear plane's position is
calculated according to the experimental measurement of zeta potential
(Ding et al., 2015; Liu et al., 2017). In another way, only a few in­
vestigations directly measured the shear layer thickness according to its
hydrodynamic definition. In electrophoresis experiments, a surface of
shear beyond the particle interface can be observed due to the electro­
static interactions between the nanoparticle and the applied electric
field (Hunter, 1981). Besides, diffusion coefficients of the colloidal
particles, counterions or water were measured to determine the shear
plane thickness without applying the electric field (Xu, 1998; Jain et al.,
2021). Bourikas et al. (2001) and Panagiotou et al. (2008) referred to the
position by those two methods as the shear plane uniformly in common.
However, the obtained position in the first method is actually the po­
sition of zeta potential rather than that of the shear plane. Henceforth, to
distinguish the two positions, this study defines the zeta position
measured by the first method, which indicates the potential at the
calculated position corresponding to the zeta potential, and shear plane
position measured by the second method, at which the streaming ve­
locity is zero.
Although a number of observations have been achieved, based on the
mentioned methods, the accurate position of the zeta potential in the
EDL remains controversial. To simplify, many researchers have assumed
that the zeta position is close to the Stern plane or even locate at it;
namely, the capacity value between two planes is much large in the TLM
(Hiemstra and Van Riemsdijk, 2006). However, Li et al. (2003) and Liu
et al. (2017) demonstrated that the zeta position is much closer to the
Gouy plane than the Stern plane after comparing the zeta potential with
surface potential and Stern potential in montmorillonite-water systems.
Molecular dynamics (MD) is a computational method that models
the molecular structure and predicts the behaviors of materials at the
atomic scale (Zhang and Pei, 2020). MD was used to explore the micro-
mechanism of clay minerals (Moussa et al., 2017; Zhang et al., 2017; Liu
et al., 2021). Chang et al. (1998), Parsons and Ninham (2010), and
Bourg and Sposito (2011) successfully applied the MD technique to
investigate the structure of EDL for clay-water systems. MD simulations
can determine the profiles of species in the diffuse layer and bulk elec­
trolyte. Furthermore, MD can be utilized to simulate the electroosmotic
flows between two planar-charged surfaces. Based on the electroosmosis
simulation, scholars have also proposed methods to determine the zeta
potential of materials in different electrolyte solutions (Předota et al.,
2016; Biriukov et al., 2020). Therefore, MD is an appropriate tool to
study the zeta potential and shear plane position of clay.
In this paper, MD was performed to determine the zeta potential and
position of the shear plane in NaCl-montmorillonite systems with
different concentrations. The electric potential profiles, particularly for
the Stern potential and zeta potential, were calculated based on MD and
the EDL model. The shear layer thickness and zeta position were dis­
cussed based on several measurement methods in the paper. The ca­
pacitances in the TLM model were finally discussed.
2. Theoretical background
2.1. Stern model
In the classical EDL theory, the Poisson-Boltzmann equation de­
scribes the distributions of ions in electrolyte solutions (Lyklema, 1995).
d2
φ
dz2
= −
4π
ε0εr
∑
i
c0
i viFexp
(
− viFφ(z)
RT
)
(1)
where φ(z) is the potential distribution in the diffuse layer; ε0 is vacuum
dielectric constant; εr is the relative permittivity of bulk water; ci
0
is the
ion concentration within the bulk solution; vi is the valence of the ith
cation species. F is the Faraday constant; R is the gas constant. In the
Stern model, by solving Eq. (1), the distribution of electric potential
away from the clay surface in the Stern layer and diffuse layer can be
expressed as (Shang et al., 1994):
φ(z) =
⎧
⎪
⎪
⎪
⎪
⎪
⎨
⎪
⎪
⎪
⎪
⎪
⎩
φ0 −
φ0 − φs
ds
z, z ≤ ds
2RT
vF
ln
eκ(z− ds)
+ tanh
vFφs
4RT
eκ(z− ds)
− tanh
vFφs
4RT
, z > ds
(2)
φ0 =
σ
Cs
+ φs (3)
where Debye-Huckel parameter κ =
̅̅̅̅̅̅̅̅̅̅̅̅
2v2F2c0
ε0εrRT
√
(m− 1
) represents the
reciprocal of the thickness of the diffuse layer; ds is the thickness of the
Stern layer; φsrepresents the Stern layer potential; φ0is the surface po­
tential and can be calculated for a given surface charge density σ and
constant capacitance of the Stern layer Cs. Once ds and φs are deter­
mined, the potential at any distance from the clay surface in the EDL can
be calculated. In other words, given a zeta potential, the position of the
zeta potential can be located by the Stern model.
Previous investigations (Bourg and Sposito, 2011; Ricci et al., 2013;
Bourg et al., 2017; Hocine et al., 2016) have demonstrated that the Stern
plane is located at the position of the OSSC. Therefore, the Stern layer
thickness can be determined from the density profiles of cations in MD
simulation. Li et al. (2004) and Hou et al. (2009) have proposed a
method to calculate the potential at the top end of the diffuse layer based
on the Poisson-Boltzmann equation. The potential is only determined by
the mean concentration of counterions in the diffuse layer ci and the
concentration in the equilibrium solution (Hou et al., 2009):
Fφs = − 2RTln
1 − 0.5
[
− A
(
ci
)
+
̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅
A
(
ci
)2
− 10.873
√ ]
1 + 0.5
[
− A
(
ci
)
+
̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅
A
(
ci
)2
− 10.873
√ ] (4)
where
A
(
ci
)
= 3.7183 +
6.8731c0
i
ci − c0
i
(5)
The derivation process of Eq. (4) is detailed in the Supporting In­
formation. Liu et al. (2015, 2016) have applied the method to calculate
the Stern potential. For the first time, this method is combined with MD
simulations to solve the potential distribution in the EDL. The average
concentration of cations in the diffuse layer can be calculated from the
density profiles in MD simulation after locating the onset of the diffuse
layer. Thus, by means of Eqs. (2)-(5), the potential distribution is
determined using only the cations density profile near the clay surface.
H. Pei and S. Zhang
Applied Clay Science 212 (2021) 106212
3
2.2. Zeta potential
The electroosmotic velocity profile is governed by the Navier-Stokes
(NS) equation together with the PB equation (Lorenz and Travesset,
2007):
u(z) =
ε0εr
η
Ex[φ(z) − ζ ] (6)
where ζ is the zeta potential. When the fluid far away from the surface,
the electroosmotic velocity is linearly related to zeta potential, namely
the Helmholtz-Smoluchowski (HS) equation (Delgado et al., 2007):
ζ = -
ueoη
ε0εrEx
(7)
where ueo represents the streaming velocity of the bulk solution under
the electric field; η is the viscosity of the solution; Ex is the external
electric field along the x-direction. In Eq. (6), the electroosmotic flow is
defined as a plug-like flow under the no-slip condition. However, MD
simulations of electroosmotic flows showed a significant slippage phe­
nomenon in previous literature (Rezaei et al., 2018; Xie et al., 2020).
Dufreche et al. (2005) have investigated the electro-osmosis in Na-
montmorillonite by MD simulations, which shows a slip length should
be taken into account to describe the profile of electroosmotic flow. An
overestimated value of the zeta potential may be obtained when
considering the no-slip condition. A Navier-type slip condition is intro­
duced to describe the slip velocity of electroosmotic flow at the interface
(Celebi and Beskok, 2018):
u‖(z0) = b
du
dz
⃒
⃒z=z0
=
ε0εr
η
Exb
dφ
dz
⃒
⃒z=z0
(8)
where u‖(z0) is the slip velocity at the slip wall of z = z0; b is the slip
length. Thus, zeta potential can be calculated as (Celebi and Beskok,
2018):
ζ =
η
ε0εrEx
(
− ueo + b
dφ
dz
⃒
⃒z=z0
)
(9)
Nonequilibrium molecular dynamics (NEMD) can be applied to
simulate the electroosmotic flow in montmorillonite nanopore, and the
zeta potential can be calculated by Eq. (9). Finally, the zeta position can
be determined by the zeta potential together with the potential profile
from the Stern model.
2.3. Triple-layer model
The triple-layer model is a popular model proposed to characterize
the electrochemical properties of colloidal systems, especially for clay-
solution systems. Taking the example of montmorillonite in a NaCl
electrolyte, the main equations of the TLM are expressed as follows
(Tournassat et al., 2009; Leroy et al., 2015):
Q0 + Qβ + Qd = 0 (10)
Qd = −
̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅
8ε0εrkbTci0NA
√
× sinh
(
Fφd
2RT
)
(11)
φβ = φd −
Qd
C2
(12)
φ0 = φβ +
Q0
C1
(13)
KNa = -
Q0 + Qβ
Qβ
aNaexp
(
−
Fφβ
RT
)
(14)
where Q0, Qβ and Qd are the surface charge densities at the 0, β-plane
and d-plane respectively and Q0is equal to the surface density of the
basal surface of montmorillonite minus that of ISCC. φ0, φβ and φd are
the corresponding potential. Notably, φd corresponds to zeta potential in
TLM. kb is the Boltzmann constant.NA is Avogadro number. KNa is the
equilibrium constant at β-plane. C1 and C2 are the capacitances within
β-plane and between β-plane and d-plane. The TLM can be simplified as
the basic Stern model by assuming φβ equal to φd.
2.4. Simulation methodology
In this paper, the zeta potential and the shear plane of montmoril­
lonite in NaCl solution were investigated. The molecular model of
montmorillonite sheets was constructed with the unit cell formula of
Si8(Al3.25Mg0.75)O20(OH)4(Greathouse et al., 2015). Isomorphous sub­
stitution occurs in the octahedral sheet, that is, Mg2+
substitution for
Al3+
. The clay layer used in the simulations consists of 4 × 4 unit cells
and has the size 20.66 Å × 35.85 Å × 6.54 Å and a surface charge density
of − 0.13C/m2
due to 12 Mg for Al substitutions. The layer was divided
in two halves located at the top and bottom of the simulation box to form
a slit-like pore with a width of 5.0 nm, which prevents overlapping of the
EDL. Water molecules with the number of 1187 were added in the pore
to reach a bulk density of 0.997 g/cm3
. Na+
was chosen as the coun­
terion. Counterions were randomly distributed in the pore to compen­
sate for the charge deficit of the clay layer. Besides, Na+
and Cl−
atoms
with the number of 2, 5, 11 and 22 were inserted in the water box,
leading to concentrations in the center of 0.20, 0.37, 0.70 and 1.30 mol/
L. The initial configuration of montmorillonite in 0.20 mol/L NaCl is
shown in Fig. 1. Periodical boundaries were set in all directions. The
interactions of minerals-minerals and minerals-water were described by
the ClayFF force field in this study (Cygan et al., 2004). The extended
simple point charge (SPC/E) water model was utilized to calculate the
water-water interactions. LAMMPS was used to perform all the MD
simulations (Plimpton, 1995).
The MD simulation was conducted in two parts. First, an equilibrium
molecular dynamics (EMD) simulation was performed under NVT
ensemble at 300 K. After a relaxation time of 10 ns, the trajectories of
atoms were recorded every 1 ns for another 200 ns. To prevent un­
wanted movements of the clay particles, Al and Mg atoms in the clay
layer were fixed during the simulations. The density profiles and mean
square displacement of ion species and water molecules were obtained
in EMD. An electric field of 5 × 108
V/m was then applied parallel to the
clay surface acting on water molecules, cations and mobile surface
atoms (surface O atoms) in the NEMD simulation. Notably, the intensity
of the electric field is higher than the strength used in experiments, while
it is in the common range in MD studies. The purpose of the high electric
Fig. 1. Snapshot of MD model for a 50 Å mesopore constructed by the parallel
montmorillonite surfaces containing 0.20 mol/L NaCl. The width of the pore is
the distance between siloxane oxygens on each side of the clay layer.
H. Pei and S. Zhang
Applied Clay Science 212 (2021) 106212
4
field is to avoid excessive noise and obtain a reasonable accuracy. The
influence of electric field intensity on NEMD was discussed in Sup­
porting Information. The system was relaxed by 20 ns under NVT
ensemble at 300 K to reach the equilibrium state. The streaming velocity
of water along the clay surface was recorded for 200 ns. The relative
permittivity and viscosity of the SPC/E water model (72.4 and 7.29 ×
10− 4
Pa⋅s) were used in Eq. (9) to calculate the zeta potential (Balasu­
bramanian et al., 1996; Gereben and Pusztai, 2011). In this study, a
standard MD thermostat (Nosé-Hoover thermostat) was used in both
EMD and NEMD simulations. However, the thermostat used in NEMD
simulations is coupled only to the degrees of freedoms perpendicular to
the flow direction. The Lennard Jones (LJ) interactions were truncated
to 12 Å, and the Lorentz-Bertholet mixing rule was adopted for the LJ
interactions between different atoms. The Coulomb interactions were
calculated by the PPPM method (Hockney and Eastwood, 1988) with
99.99% accuracy. The time step of 1 fs was set to integrate Newton's
motion equations in the simulations.
3. Results and discussion
3.1. Ion distributions and stern potential
The density distributions of water along the aperture direction in the
montmorillonite mesopore containing different NaCl solutions present
similar tendencies, as shown in Fig. S2. It is noted that the location of
surface O atoms in the montmorillonite layer is the origin of the z-axis.
Pronounced water layers occur near the clay surface and dissipate
beyond 10 Å, which is in agreement with water on the clay surfaces from
previous investigations (Park and Sposito, 2002). The first peak of water
density is located at 2.7 Å, close to the value for other MD simulations
(2.7 ± 0.5 Å) (Bourg and Sposito, 2011) and X-ray experiments of mica
(2.5 ± 0.2 Å) (Cheng et al., 2001). For each NaCl-montmorillonite sys­
tem, the density profiles of Na+
and Cl−
as a function of distance from
the clay surface are presented in Fig. 2. The convergence and error
analysis of ions density profile were discussed in Supporting Informa­
tion. The density profile of Na+
close to the surface shows the first two
peaks corresponding to the plane of ISSC and OSSC, respectively. The
stern plane positions for 0.20, 0.37, 0.70 and 1.30 mol/L are determined
as the distance of 4.3 Å away from the clay surface. The thickness of
Stern layer in MD simulations of montmorillonite in mixed NaCl-CaCl2
solutions (4.35–4.55 Å) (Bourg and Sposito, 2011) and measurement
values of divalent cations OSSCs on mica (4.52 ± 0.24 Å) by X-ray (Park
et al., 2006) are close to the results in this study. The simulations show
that the z-coordination of OSSC is independent of the ionic
concentration.
The diffuse layer region is determined based on the Debye length
after locating the Stern plane. The mean concentration of cations in the
diffuse layer can be calculated from its density profiles, which are listed
in Table 1. The Stern potential of 0.20–1.30 mol/L NaCl solutions are
calculated as − 80.2, − 67.0, − 52.0 and − 39.7 mV by Eq. (4), respec­
tively. The Stern potential decreases with increasing ion concentration,
which is consistent with the EDL theory. Miller and Low (1990) used
several methods combined with experiments and theories to determine
the Stern potential of montmorillonite in distilled water. The Stern po­
tentials of Na-montmorillonite with the cation exchange capacity of 90
meq/100 g were measured in the range of − 55.5 to − 59.1 mV. The Stern
potential of montmorillonite with − 0.1 to − 0.15C/m2
surface charge
density in 0.1 mol/L NaCl solution was calculated in the range of − 79.5
to − 96.8 mV in theory (Liu et al., 2016). Compared with their results,
the calculated Stern potential is within a reasonable range.
Once the Stern potential is determined, the electric potential distri­
bution in the diffuse layer can be calculated by Eq. (2). The density
profiles of ion species beyond the Stern plane predicted by the Stern
model are consistent with MD results, as shown in Fig. 2. The Stern
model with a simple way of the Stern potential determination can well
match the molar density of diffuse swarm (DS) species, however, slightly
underestimates the density of the OSSC. The difference at the Stern
plane may be attributed to the PB equation neglecting some effects such
as ion polarizability. However, this study still focuses on the classical
EDL theory in this paper. A popular model named the modified Gouy-
Chapman (MGC) model has been proposed to describe quantitatively
the ion concentration profiles in the EDL (Tournassat et al., 2009; Le
Crom et al., 2020). The main equations are expressed as:
⎧
⎪
⎪
⎪
⎨
⎪
⎪
⎪
⎩
Fφ(a)
RT
= − 2 × sinh− 1
(
F|σ|
2κε0εrRT
)
Fφ(z)
RT
= 4 × tanh− 1
[
tanh
(
Fφ(a)
4RT
)
× exp( − κ(z − a) )
] (15)
where a is a distance from the surface. σ is the surface charge minus the
ions charge within a. For montmorillonite in 0.20 M NaCl solution, the
comparison of the GC, MGC and Stern model is shown in Fig. 3. The Na+
density profile predicted by the GC model has the maximum molar
density at the clay surface, which shows a distinct difference from MD
simulation. Although the profiles by the MGC model with different a are
close to the MD result beyond 10 Å, the values of calculated profiles are
in disagreement with each other near the clay surface. In addition, the
predicted profiles of the MGC model at the Stern plane are less accurate
than the Stern model. The density profile of the Stern model is slightly
larger than that of the GC and MGC model beyond 10 Å. The fitted
0 5 10 15 20 25
0
1
2
3
4
5
6
7
8
Na+
by MD
Na+
by Stern model
Na
+
concentration
(mol/L)
Distance from the clay surface (Å)
0
1
2
Cl- by MD
Cl- by Stern model
Cl
-
concentration
(mol/L)
Stern plane
(OSSC)
1.30 mol·L-1
Fig. 2. Molar density profiles of Na, Cl at the clay surface as obtained from the
MD model of montmorillonite in 1.30 mol/L NaCl electrolyte. The light dash
lines are the density profiles of Na and Cl predicted by the Stern model. The
density profiles of 0.70, 0.37 and 0.20 mol/L NaCl electrolytes are shown
in Fig. S3.
Table 1
Calculated parameters in the Stern model for montmorillonite in NaCl solutions.
The capacitance in the Stern layer is 1.0 F m− 2
.
0.20
mol/L
0.37
mol/L
0.70
mol/L
1.30
mol/L
Mean concentration of Na+
in
diffuse layer, c(mol/L)
1.43 1.94 2.55 3.50
Electroosmotic mobility of bulk
water, μeo/Ex (m2
/V⋅s)
11.02 10.13 8.77 7.90
Thickness of Stern plane, ds (Å) 4.3 4.3 4.3 4.3
Position of zeta potential, dζ (Å) 7.2 6.8 6.5 5.8
Thickness of Gouy plane, dg (Å) 11.1 9.3 7.9 7.0
Surface potential, φ0 (mV) − 210.2 − 197.0 − 182.0 − 169.7
Stern potential, φs (mV) − 80.2 − 67.0 − 52.0 − 39.7
Zeta potential, ζ (mV) − 46.8 − 38.8 − 26.9 − 22.3
H. Pei and S. Zhang
Applied Clay Science 212 (2021) 106212
5
parameter a seems to lack a clear physical significance resulting in an
uncertain origin of the density profile, compared to the peak located at
the Stern plane in the Stern model. Therefore, the Stern model combined
with a calculated Stern potential has good accuracy in predicting ion
density profiles in the EDL.
3.2. Electroosmotic flow and zeta potential
Fig. 4 shows the electroosmotic velocity profile of water in mont­
morillonite pore containing 0.20 mol/L NaCl solutions. The streaming
velocity follows the hydrodynamical Navier-Stokes equation (Eq. (6))
beyond the first peak of water density profile, whereas the velocity drops
rapidly within that. A clear slip velocity can be observed at the first peak
of water density profile. This is because only a small number of water
molecules before the first peak of water density profile that strongly
absorbed by the clay surface have different dynamics properties from
the bulk water. The same observation has also been presented in the
NEMD simulation of an electroosmotic flow of CsCl solution in a charged
silica slit by Siboulet et al. (2017). Celebi and Beskok (2018); Celebi
et al. (2019) have considered the shear plane is located at the first peak
of the water density profile. Although the streaming velocity disobeys
the NS equation, there is still a non-zero velocity within 2.7 Å in this
simulation as observed in Fig. 4. Therefore, the first peak of the water
density profile cannot be regarded as the shear plane simply. Siboulet
et al. (2017) also demonstrated that the stagnant layer close to the
surface belongs to the ion hydration sphere. This paper confirms that the
flow domain starts at the first water density peak which is 2.7 Å away
from the clay surface in the simulation. A function was proposed to fit
the streaming velocity profile from MD simulations, which the deriva­
tion is shown in Supporting Information. Then the slip velocity and slip
length are obtained based on the fitting velocity profile. Both slip length
and slip mobility decreases with the concentration increases of NaCl,
which the calculated values are listed in Table 2. Rezaei et al. (2018)
also proved such tendency of slip length and slip velocity. The electro­
osmotic mobility (velocity profile normalized by field strength) along z-
direction is exhibited in Fig. 5. The plug-like profiles for NaCl-
montmorillonite systems are observed as expected. The average
mobility at the plateau is significantly affected by the ionic concentra­
tion, decreasing with the increasing of concentration. The zeta potential
of 0.20–1.30 mol/L NaCl solutions calculated by Eq. (9) are − 46.8,
− 38.8, − 26.9 and − 22.3 mV respectively. Tournassat et al. (2009)
measured the zeta potential of − 38 mV for montmorillonite (σ= −
0.11C/m2
) in 0.12 molL− 1
NaCl solution. Mészáros et al. (2019)
measured the zeta potential of bentonite (σ= − 0.14C/m2
) in 0.1 mol/L
NaCl solution as − 37.4 to − 47 mV in the range of 2–12 pH. The zeta
potential of Na-montmorillonite (σ= − 0.16C/m2
, pH = 6.5) at 1.0 M
was determined as − 17.9 mV (Sondi et al., 1996). It shows that the
values of zeta potential in the NEMD simulation are slightly larger than
to the measurements. The zeta potential of 0.20–1.30 mol/L NaCl so­
lutions are calculated as − 126.7, − 116.5, − 100.8 and − 90.9 mV
respectively, without considering the slip length (Eq. (7)), which are
significantly larger than the measured values. Therefore, the slip length
should be considered when calculating the zeta potential to avoid an
overestimated value.
3.3. Zeta position and shear plane position
After obtaining potential distribution based on the Stern model and
the zeta potential in the EDL, the zeta position can be calculated, as
shown in Fig. 6. The zeta positions are calculated as 7.2, 6.8, 6.5 and 5.8
Å for montmorillonite at 0.20–1.30 mol/L. Results show that the posi­
tion of zeta potential cannot be identified as the Stern plane simply. The
zeta position is closer to the Gouy plane than the Stern plane. For
montmorillonite in NaCl solution, the zeta position from the clay surface
decreases as the ionic concentration increases because of the Debye
screening length.
The TLM can also obtain the zeta position after the determination of
zeta potential. The charge density at d-plane (zeta position) Qd can be
calculated by Eq. (11). Besides, Qd is equal to subtracting the compen­
sated charge within the d-plane from the surface charge of the
0 5 10 15 20 25
0
1
2
3
4
5
6
0 5 10 15 20 25
0.00
0.05
0.10
0.15
0.20
Cl
-
concentration
(mol/L)
Na
+
concentration
(mol/L)
Distance from the clay surface (Å)
MD
Stern model
GC model
MGC a=0.21 nm
MGC a=0.31 nm
MGC a=0.43 nm
MGC a=0.57 nm
Fig. 3. The comparison among Na+
(and Cl−
in the inserted graph) density
profile for montmorillonite at NaCl concentration of 0.20 mol/L in MD simu­
lation and the predictions of the Gouy-Chapman model, Stern model and
modified Gouy-Chapman model with fitted parameter a of 0.21, 0.31, 0.43 and
0.57 nm.
25 20 15 10 5 0 -5
0
5
10
15
20
25
30
35
40
45
50
55
60
0.0
0.5
1.0
1.5
2.0
2.5
w
u||
(z0
)
Water
density
(g/cm
3
)
ueo
b=2.9 Å
z=z0
0
|z z
du
k
dz
0
0
|
r
x z z
d
E
dz
uMD
uslip
ustick
u(z)
(m/s)
z (Å)
Fig. 4. Electroosmotic velocity profiles for montmorillonite in 0.20 mol/L NaCl
electrolytes. uMD is the velocity of the fluid along the aperture direction from
MD, uSlip from fitting uMD in a PB equation form as shown in Supporting In­
formation, and uStick by assuming a vanishing velocity at z = 2.7 Å. u‖(z0) is the
slip velocity. b is the slip length.
Table 2
The slip length and slip mobility for montmorillonite in NaCl electrolytes ob­
tained by MD and Eq. (8).
Concentration
(mol/L)
Slip length
(Å)
Slip mobility
(×10− 8
m2
/V s)
0.20 2.90 6.90
0.37 2.50 6.73
0.70 1.95 6.44
1.30 1.90 5.96
H. Pei and S. Zhang
Applied Clay Science 212 (2021) 106212
6
montmorillonite basal surface (Eq. (10)). Therefore, by combining two
equations, the d-plane positions are 6.0, 5.5, 5.3 and 4.8 Å for mont­
morillonite at 0.20, 0.37, 0.70 and 1.30 mol/L, as listed in Table 3.
Results show that the position of the d-plane is located beyond the
β-plane, in accordance with the TLM. The same tendency is exhibited in
the TLM that the distance of zeta position from the clay surface de­
creases with electrolyte concentration. However, the distance between
the zeta position and the Stern plane in the TLM is less than in the Stern
model. Fig. 7 shows the relationship between the zeta position from the
interface and diffuse layer, zeta position and ionic strength from MD
simulations. The results from the Stern model and the TLM indicate that
the zeta position from the clay surface has a strictly linear relationship
with the Debye length, which agrees with the conclusion demonstrated
by Charlton and Doherty (1999) through experiments. Moreover, the
results also demonstrate that the zeta position (log scale) is linearly
dependent on the log value of the ionic strength, in accordance with the
conclusions of Bourikas et al. (2001) and Panagiotou et al. (2008).
After the zeta position determination, the position of the shear plane
was also investigated in this paper. In the Stern model, the position of
zeta potential is located at the shear plane, at which the flow has a nil
velocity under no-slip conditions. In the simulation, however, there is no
stagnant layer within the zeta potential position. Similarly, more and
more MD simulations have proved that the stagnant layer near a solid
surface cannot be observed in the EDL (Lylema et al., 1998; Freund,
2002; Hartkamp et al., 2005; Lorenz et al., 2008; Zhang et al., 2011).
The distinction between the standard hydrodynamics and the MD results
may contribute to the following reason. In standard hydrodynamics and
Stern model, the concept of shear plane strictly separating two neighbor
regions with an infinite and a constant viscosity of water respectively is
in contradiction with the findings that the viscosity changes continu­
ously near the interface by experiments and MD simulations (Schmickler
and Henderson, 1986; Evans and Wennerström, 1999; Bazant et al.,
2009; Hartkamp et al., 2018). Therefore, although the Stern model can
predict well the static potential and ion distributions on the global EDL,
the local physical properties in electrokinetic phenomenon cannot be
accurately described by the NS equation and Stern model.
Xu (1998) has proposed a method to measure the shear plane
thickness of micellar particles based on their self-diffusion coefficients.
The hydrodynamic radius Rh can be calculated by self-diffusion coeffi­
cient Ds using the Stokes-Einstein relationship:
Ds =
kBT
6πηRh
(16)
The thickness of the shear layer can be obtained by subtracting the
particle radius from the hydrodynamic radius. Jain et al. (2021) deter­
mined the shear layer thickness of nano-particles by calculating the self-
diffusion coefficient of cations in MD simulations. Therefore, the diffu­
sion coefficients of water oxygen and cations in the slit pore divided into
slices of 2 Å thickness were calculated in the simulations to evaluate
whether the coefficients can be used to determine the shear plane. The
time interval is an important parameter in the calculation of the diffu­
sion coefficient. The influence of the time interval was discussed in
Supporting Information. After comparison, the length of time intervals
of 10 ps was chosen used in mean-square displacement calculation in
MD simulations. The diffusion coefficients of water oxygen and Na+
parallel to the clay surface along z-direction are illustrated in Fig. 8,
normalized by the values in the bulk 0.20 mol/L NaCl solution. Diffusion
coefficients decrease linearly within 15 Å from the interface and reach a
stable value beyond that, in agreement with the other MD results of
montmorillonite nanopore (Bourg and Sposito, 2011). It can be noted
that both coefficients of water and Na+
slightly decrease as ionic con­
centration increases. The Shear plane position is assumed located at the
turning point with the diffusion value equals to the bulk value, which is
15 Å away from the clay surface. However, the calculated position is
0 10 20 30 40 50
0
3
6
9
12
15
u(z)/E
x
(m
2
/V·s)
z (Å)
0.20 mol/L
0.37 mol/L
0.70 mol/L
1.30 mol/L
×10-8
2.7 Å
Fig. 5. Electroosmotic velocity profiles normalized by the strength of the
external electric field for montmorillonite in 1.30, 0.70, 0.37 and 0.20 mol/L
NaCl electrolytes. The flow domains start at 2.7 Å away from the clay surface.
Fig. 6. The electric potential distributions of montmorillonite with different
ion concentrations. The position of the zeta position is obtained from the Stern
model and zeta potential.
Table 3
Calculated parameters in the triple-layer model for montmorillonite in NaCl
solutions. The capacitance C1 in the Stern layer is 1.0 F m− 2
.
0.20 mol/
L
0.37 mol/
L
0.70 mol/
L
1.30 mol/
L
Surface charge at 0-plane,
Q0 (C/m2
)
− 0.1293 − 0.1297 − 0.1294 − 0.1288
Surface charge at d-plane,
Qd (C/m2
)
− 0.0562 − 0.0613 − 0.0573 − 0.0614
Capacitance C2 (F⋅m− 2
) 3.34 4.49 5.09 8.04
Position of β-plane, zβ (Å) 4.3 4.3 4.3 4.3
Position of d-plane, zd (Å) 6.0 5.5 5.3 4.8
Potential at 0-plane,
φ0 (mV)
− 193.0 − 181.4 − 167.5 − 156.6
Potential at β-plane, φβ
(mV)
− 63.7 − 51.7 − 38.1 − 27.8
Potential at d-plane,
φd (mV)
− 46.8 − 38.8 − 26.9 − 22.3
H. Pei and S. Zhang
Applied Clay Science 212 (2021) 106212
7
beyond the Gouy plane in all concentrations of NaCl solutions, which
contradicts the Stern model. Although the diffusion coefficients of water
and cations have reduced mobility near the surface, a stagnant layer
with zero diffusion coefficient cannot be observed. To sum up, the re­
sults confirmed that the shear plane position cannot be observed from
the electroosmotic flow and diffusion coefficient in MD simulations.
3.4. Discussion on the TLM
The TLM consists of several equations and parameters that make the
solution difficult, which simplex algorithm is commonly used to solve
the model (Leroy and Revil, 2004; Leroy et al., 2015). Parameters such
as equilibrium constants at corresponding planes and capacitances of the
inner and outer part of the Stern layer need to be estimated before the
solution. Previous studies have evaluated these parameters based on
experiments and MD simulations, especially for capacitances C1 and C2.
A basic agreement was achieved for C1 based on the proposed values of
0.6–1.3 F m− 2
by experiments (Machesky et al., 1998; Sverjensky, 2001)
and 0.8–1.2 F m− 2
by MD simulations (Tournassat et al., 2009). How­
ever, the value of C2 in the range of 0.2–5.5 F m− 2
(Yates et al., 1974;
Hiemstra and Van Riemsdijk, 2006; Wolthers et al., 2008; Bourg and
Sposito, 2011) is still in dispute. In this paper, the capacitance C2
between β-plane and d-plane was investigated. As the obtained position
of β-plane and d-plane in the TLM (Table 2), the capacitance C2 can be
estimated by the relation (Sahai and Sverjensky, 1997):
Ci = εiε0/Δzi (17)
where εi and Δzi are the relative permittivity and thickness of each
“capacitor”. The spaces between β-plane and d-plane for montmoril­
lonite in 0.20–1.30 mol/L NaCl solutions are 1.7, 1.2, 1.0 and 0.6 Å.
Wander and Clark (2008) have calculated the dielectric constants for the
quartz-water interface as a function of the distance from the surface. In
this paper, the permittivity of the outer Stern layer is calculated based on
the proposed curve (permittivity scaled using reference value of 78 for
bulk water), given as 64.2, 60.9, 57.5 and 54.5, respectively. Thus, the
capacitance C2 in the TLM can be obtained by Eq. (17) as 3.34, 4.49,
5.09 and 8.04 F m− 2
, respectively. The results show that the value of
capacitance between β-plane and d-plane has a large range depending
on the ionic concentration, which is in disagreement with the constant
value calculated by Bourg and Sposito (2011) but in agreement with the
conclusion of Nishimura et al. (2002). After obtaining the capacitance
between β-plane and d-plane, the TLM can be solved. Fig. S4 shows the
potential distribution from the clay surface based on the TLM. The
capacitance within the Stern layer is set as 1.0 F m− 2
. The potential
2 3 4 5 6 7 8
4
5
6
7
8
9
by Stern model
by TLM model
Zeta
position
from
the
surface
(Å)
-1(Å)
0.28 / 5.37
d
0.25 / 4.28
d
2
0.98
R
2
0.96
R
0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.80.9 1
4
5
6
7
8
9
10
2
0.96
R
2
0.93
R
log -0.112log 0.664
d I
log -0.111log 0.751
d I
By Stern model
By triple-layer model
logd
(Å)
logI (M)
Fig. 7. The distance of the zeta potential position from the clay surface calculated by the Stern model and triple-layer model as a function of the Debye length and
ionic strength.
Fig. 8. Self-diffusion coefficients of water oxygen and Na+
parallel to the clay surface as a function of distance from the basal surface. Error bars show the values of
95% confidence interval.
H. Pei and S. Zhang
Applied Clay Science 212 (2021) 106212
8
calculated by the TLM is smaller than the Stern model, especially for the
potential between β-plane and d-plane. The positions of zeta potential
calculated by the TLM are also smaller than those by the Stern model, as
mentioned above.
4. Limitations
This paper only studies the EDL structure and electroosmosis prop­
erties of the basal surface of montmorillonite. It is well known, mont­
morillonite nanoparticle has two surfaces including basal and edge
surface. Due to the large specific surface area, the basal surface is the
domain part of the surface of montmorillonite, which indicates the re­
sults in this paper can generally reflect the properties of montmoril­
lonite. However, surface complexes on the edge site can also develop in
the EDL. Cations exchange on edge surfaces shows complex, thermo­
dynamically non-ideal behavior much different from exchange on the
basal surface (Lammers et al., 2017). Besides, the orientation (Kraevsky
et al., 2020) and cleavage type (Shen and Bourg, 2021) of the edge in­
fluence the stability of the edge structure and its adsorption character­
istics. The investigations of surface complexation of edge surfaces can be
found in Newton et al. (2016, 2017) and Kraevsky et al. (2020). Besides,
the edge surface of montmorillonite also has an influence on the value of
zeta potential. Edges surfaces have a pH-dependent charge, while basal
surfaces bear a negative charge generated by isomorphous substitution
(Delgado et al., 1985; Durán et al., 2000). This is a significant reason
that the zeta potential of montmorillonite changes as a function of pH
value. The effect of particle morphologies on the zeta potential was also
not considered in this paper. Those reasons may cause the gap between
the zeta potential measurement and the MD simulation result.
This paper distinguished the zeta potential and the shear plane. The
latter one was confirmed not to exist from MD simulations. The zeta
potential position was measured and its properties were investigated in
the paper. Thus, two questions arise. What is the physical meaning of the
zeta potential if the shear plane does not exist? What is the physical
meaning of the zeta potential position? This paper attempts to answer
these questions by combing our thoughts with previous literature.
Because those are still arguments in collochemistry, more in-depth work
should be done in the future.
Firstly, although its definition is related to the shear plane, the
generation and physical meaning of zeta potential is independent of the
existence of the shear plane. The shear plane is an imaginary plane only
used to reconcile the Poisson-Boltzmann equation and the Navier-Stokes
equation (Eq. (6)). Předota et al. (2016) have demonstrated that the zeta
potential does not arise from the existence of a shear plane, but rather
precisely from the electrokinetically driven motion of ions within the
whole inhomogeneous interfacial region. Delgado et al. (2007) also said,
the zeta potential is fully defined by the nature of the surface, its charge,
the electrolyte concentration in the solution, and the nature of the
electrolyte and of the solvent. Therefore, the zeta potential is still
meaningful without the shear plane. Secondly, although the physical
interpretation of the zeta potential is still ambiguous, a widely accepted
meaning is proposed by Delgado et al. (2007). The zeta potential is the
observed electrokinetic signal crossing from the non-contributing region
for electrokinetic phenomenon to the double layer.
Following this idea, the physical meaning of the zeta potential po­
sition is given as which is the boundary of electrokinetic phenomenon
non-contributing region. The charges located between the surface and
the zeta potential position are electrokinetically inactive exhibiting only
electrostatic properties and contribute to the excess conductivity of the
double layer (Delgado et al., 2007). This is the reason that the bulk
streaming velocity is related to the zeta potential rather than the surface
potential in the H–S equation. Although no electrokinetic effect is
generated in this region under an external electrical field, water behind
the zeta potential position has non-zero velocity in the electroosmotic
tangential flow because of the viscosity. This also shows the difference
between the shear plane and zeta potential position and proves the non-
existence of the shear plane.
5. Conclusions
A novel approach based on molecular dynamics was proposed in this
paper to determine the zeta potential and the position of the shear plane
for montmorillonite in NaCl electrolyte with a concentration of
0.20–1.30 mol/L. It combines an electrostatic surface complexation
model (Stern model or TLM) that determines the electric potential dis­
tribution in the EDL from the ion density profiles in the EMD with the
electroosmotic flow NEMD simulation that calculates the zeta potential.
The conclusions can be summarized as follows:
(1) The density profiles of ion species clearly show the existence of
ion adsorption complexes. The Stern model has better accuracy in
predicting the ions density profiles after the Stern potential
determination, compare with the Gouy-Chapman and modified
Gouy-Chapman model.
(2) The zeta potential is calculated based on the electroosmotic ve­
locity profile, considering the slip length, close to the experi­
mental measurement. Methods including NEMD and self-
diffusion coefficients cannot observe the shear plane. The posi­
tion of zeta potential was determined by the Stern model and
TLM, showing a certain distance from the Stern plane. The dis­
tance of the zeta position from the clay surface has a strictly linear
relationship with the Debye length and the ionic strength in the
log scale, both in agreement with the experimental observation.
(3) The capacitance between the Stern and shear plane in the TLM
was calculated indicating that the capacitance is variable
depending on the ionic strength rather than a constant value.
Declaration of Competing Interest
The authors declare that they have no known competing financial
interests or personal relationships that could have appeared to influence
the work reported in this paper.
Acknowledgements
This research has been supported by the China National Key R&D
Program during the 13th Five-year Plan Period (Grant No.
2018YFC1505104 and 2017YFC1503103), National Natural Science
Foundation of China (Grants No. 51778107) and Liao Ning Revitaliza­
tion Talents Program (Grants No. XLYC1807263).
Appendix A. Supplementary data
Supplementary data to this article can be found online at https://doi.
org/10.1016/j.clay.2021.106212.
References
Balasubramanian, S., Mundy, C.J., Klein, M.L., 1996. Shear viscosity of polar fluids:
Molecular dynamics calculations of water. J. Chem. Phys. 105, 11190–11195.
https://doi.org/10.1063/1.472918.
Bazant, M.Z., Kilic, M.S., Storey, B.D., Ajdari, A., 2009. Towards an understanding of
induced-charge electrokinetics at large applied voltages in concentrated solutions.
Adv. Colloid Interface 152, 48–88.
Biriukov, D., Fibich, P., Předota, M., 2020. Zeta potential determination from molecular
simulations. J. Phys. Chem. C 124, 3159–3170. https://doi.org/10.1021/acs.
jpcc.9b11371.
Bourg, I.C., Sposito, G., 2011. Molecular dynamics simulations of the electrical double
layer on smectite surfaces contacting concentrated mixed electrolyte (NaCl-CaCl2)
solutions. J. Colloid Interface Sci. 360, 701–715. https://doi.org/10.1016/j.
jcis.2011.04.063.
Bourg, I.C., Lee, S.S., Fenter, P., Tournassat, C., 2017. Stern layer structure and energetics
at mica-water interfaces. J. Phys. Chem. C 121, 9402–9412. https://doi.org/
10.1021/acs.jpcc.7b01828.
Bourikas, K., Hiemstra, T., Van Riemsdijk, W.H., 2001. Ion pair formation and primary
charging behavior of titanium oxide (anatase and rutile). Langmuir 17, 749–756.
H. Pei and S. Zhang
Applied Clay Science 212 (2021) 106212
9
Celebi, A.T., Beskok, A., 2018. Molecular and continuum transport perspectives on
electroosmotic slip flows. J. Phys. Chem. C 122, 9699–9709. https://doi.org/
10.1021/acs.jpcc.8b02519.
Celebi, A.T., Cetin, B., Beskok, A., 2019. Molecular and continuum perspectives on
intermediate and flow reversal regimes in electroosmotic transport. J. Phys. Chem.
C. https://doi.org/10.1021/acs.jpcc.9b02432.
Chang, F.R.C., Skipper, N.T., Sposito, G., 1998. Monte Carlo and molecular dynamics
simulations of electrical double-layer structure in potassium-montmorillonite
hydrates. Langmuir 14, 1201–1207. https://doi.org/10.1021/la9704720.
Charlton, I.D., Doherty, A.P., 1999. Locating the micellar shear plane and its relationship
with the Debye screening length. J. Phys. Chem. B 103, 5081–5083. https://doi.org/
10.1021/jp9843914.
Cheng, L., Fenter, P., Nagy, K.L., Schlegel, M.L., Sturchio, N.C., 2001. Molecular-scale
density oscillations in water adjacent to a mica surface. Phys. Rev. Lett. 87 https://
doi.org/10.1103/PhysRevLett.87.156103, 156103-156103–4.
Cygan, R.T., Liang, J.-J., Kalinichev, A.G., 2004. Molecular models of hydroxide,
oxyhydroxide, and clay phases and the development of a general force field. J. Phys.
Chem. B 108, 1255–1266. https://doi.org/10.1021/jp0363287.
Delgado, A.V., Gonzalez-Cabllero, F., Bruque, J.M., 1985. On the zeta potential and
surface charge density of montmorillonite in aqueous electrolyte solutions. J. Colloid
Interface Sci. 13 (1), 203–211.
Delgado, A.V., Gonzalez-Cabllero, F., Hunter, R.J., Koopal, L., Lyklema, J., 2007.
Measurement and interpretation of electrokinetic phenomena. J. Colloid Interface
Sci. 309, 194–224.
Ding, W., Liu, X., Song, L., Li, Q., Zhu, Q., Zhu, H., Hu, F., Luo, Y., Zhu, L., Li, H., 2015.
An approach to estimate the position of the shear plane for colloidal particles in an
electrophoresis experiment. Surf. Sci. 632, 50–59. https://doi.org/10.1016/j.
susc.2014.08.024.
Dufreche, J.F., Marry, V., Malikova, N., Turq, P., 2005. Molecular hydrodynamics for
electro-osmosis in clays: from Kubo to Smoluchowski. J. Mol. Liq. 118, 145–153.
https://doi.org/10.1016/j.molliq.2004.07.076.
Durán, J.D.G., Ramos-Tejada, M.M., Arroyo, F.J., Gonzalez-Cabllero, F., 2000.
Rheological and electrokinetic properties of sodium montmorillonite suspensions I.
Rheological properties and interparticle energy of interaction. J. Colloid Interface
Sci. 229, 107–117.
Evans, D.F., Wennerström, H., 1999. The Colloidal Domain: Where Physics, Chemistry,
Biology, and Technology Meet, 2nd ed. Wiley-VCH, New York.
Freund, J.B., 2002. Electro-osmosis in a nanometer-scale channel studied by atomistic
simulation. J. Chem. Phys. 116, 2194–2200. https://doi.org/10.1063/1.1431543.
Gereben, O., Pusztai, L., 2011. On the accurate calculation of the dielectric constant from
molecular dynamics simulations: the case of SPC/E and SWM4-DP water. Chem.
Phys. Lett. 507, 80–83. https://doi.org/10.1016/j.cplett.2011.02.064.
Greathouse, J.A., Hart, D.B., Bowers, G.M., Kirkpatrick, R.J., Cygan, R.T., 2015.
Molecular simulation of structure and diffusion at smectite-water interfaces: using
expanded clay interlayers as model nanopores. J. Phys. Chem. C 119, 17126–17136.
https://doi.org/10.1021/acs.jpcc.5b03314.
Greenwood, R., 2003. Review of the measurement of zeta potentials in concentrated
aqueous suspensions using electroacoustics. Adv. Colloid Interf. Sci. 106, 55–81.
https://doi.org/10.1016/S0001-8686(03)00105-2.
Hartkamp, R., Siboulet, B., Dufreche, J.F., Coasne, B., 2005. Ion-specific adsorption and
electroosmosis in charged amorphous porous silica. Phys. Chem. Chem. Phys. 17
(2005), 24683–24695.
Hartkamp, R., Biance, A.L., Fu, L., Dufreche, J.F., Bonhomme, O., Joly, L., 2018.
Measuring surface charge: why experimental characterization and molecular
modeling should be coupled. Curr. Opin. Colloid Interface Sci. 37, 101–114.
Hiemstra, T., Van Riemsdijk, W.H., 2006. On the relationship between charge
distribution, surface hydration, and the structure of the interface of metal
hydroxides. J. Colloid Interface Sci. 301, 1–18. https://doi.org/10.1016/j.
jcis.2006.05.008.
Hocine, S., Hartkamp, R., Siboulet, B., Duvail, M., Coasne, B., Turq, P., Dufrêche, J.F.,
2016. How ion condensation occurs at a charged surface: a molecular dynamics
investigation of the Stern layer for water-silica interfaces. J. Phys. Chem. C 120,
963–973. https://doi.org/10.1021/acs.jpcc.5b08836.
Hockney, R., Eastwood, J., 1988. Computer Simulation Using Particles. CRC Press.
Hou, J., Li, H., Zhu, H., Wu, L., 2009. Determination of clay surface potential: a more
reliable approach. Soil Sci. Soc. Am. J. 73, 1658–1663. https://doi.org/10.2136/
sssaj2008.0017.
Hunter, R.J., 1981. Zeta Potential in Colloidal Science: Principles and Applications.
Academic Press, New York.
Jain, K., Mehandzhiyski, A.Y., Zozoulenko, I., Wågberg, L., 2021. PEDOT:PSS nano-
particles in aqueous media: a comparative experimental and molecular dynamics
study of particle size, morphology and z-potential. J. Colloid Interface Sci. 584,
57–66. https://doi.org/10.1016/j.jcis.2020.09.070.
Kraevsky, S.V., Tournassat, C., Vayer, M., Warmont, F., et al., 2020. Indentification of
montmorillonite particle edge orientations by atomic-force microscopy. Appl. Clay
Sci. 186, 105442.
Lammers, L.N., Bourg, I.C., Okumura, M., Kolluri, K., et al., 2017. Molecular dynamics
simulations of cesium adsorption on illite nanoparticles. J. Colloid Interface Sci. 490,
608–620.
Le Crom, S., Tournassat, C., Robinet, J.C., Marry, V., 2020. Influence of polarizability on
the prediction of the electrical double layer structure in a clay mesopore: a molecular
dynamics study. J. Phys. Chem. C 124, 6221–6232.
Leroy, P., Revil, A., 2004. A triple-layer model of the surface electrochemical properties
of clay minerals. J. Colloid Interface Sci. 270, 371–380. https://doi.org/10.1016/j.
jcis.2003.08.007.
Leroy, P., Revil, A., Kemna, A., Cosenza, P., Ghorbani, A., 2008. Complex conductivity of
water-saturated packs of glass beads. J. Colloid Interface Sci. 321, 103–117.
Leroy, P., Tournassat, C., Bernard, O., Devau, N., Azaroual, M., 2015. The electrophoretic
mobility of montmorillonite. Zeta potential and surface conductivity effects.
J. Colloid Interface Sci. 451, 21–39. https://doi.org/10.1016/j.jcis.2015.03.047.
Li, G.X., 2016. Advanced Soil Mechanics, second edition. Tsinghua University Press (in
Chinese).
Li, H., Wei, S., Qing, C., Yang, J., 2003. Discussion on the position of the shear plane.
J. Colloid Interface Sci. 258, 40–44. https://doi.org/10.1016/S0021-9797(02)
00077-2.
Li, H., Qing, C.L., Wei, S.Q., Jiang, X.J., 2004. An approach to the method for
determination of surface potential on solid/liquid interface: theory. J. Colloid
Interface Sci. 275, 172–176.
Liu, X., Hu, F., Ding, W., Tian, R., Li, R., Li, H., 2015. A how-to approach for estimation of
surface/Stern potentials considering ionic size and polarization. Analyst 140,
7217–7224. https://doi.org/10.1039/c5an01053e.
Liu, X., Ding, W., Tian, R., Li, R., Li, H., 2016. How ionic polarization affects Stern
potential: an insight into hofmeister effects. Soil Sci. Soc. Am. J. 80, 1181–1189.
https://doi.org/10.2136/sssaj2016.04.0095.
Liu, X., Ding, W., Tian, R., Du, W., Li, H., 2017. Position of shear plane at the clay-water
interface: strong polarization effects of counterions. Soil Sci. Soc. Am. J. 81,
268–276. https://doi.org/10.2136/sssaj2016.08.0261.
Liu, X.T., Yang, S., Gu, P.K., Liu, S., Yang, G., 2021. Adsorption and Removal of Metal
Ions by Smectites Nanoparticles: Mechanistic Aspects, and Impacts of Charge
Location and Edge Structure.
Lorenz, C.D., Travesset, A., 2007. Charge inversion of divalent ionic solutions in silica
channels. Phys. Rev. E 75, 061202.
Lorenz, C.D., Crozier, P.S., Anderson, J.A., Travesset, A., 2008. Molecular dyncamics of
ionic transport and electrokinetic effects in realistic silica channels. J. Chem. Phys. C
112, 10222–10232.
Lyklema, J. 1995. Fundamentals of Interface and Colloid Science, Academic Press, 2:
1–232.
Lylema, J., Rovillard, S., de Coninck, J., 1998. Electrokinetics: the properties of the
stagnant layer unraveled. Langmuir 14, 5659–5663.
Machesky, M.L., Wesolowski, D.J., Palmer, D.A., Ichiro-Hayashi, K., 1998.
Potentiometric titrations of rutile suspensions to 250◦
C. J. Colloid Interface Sci. 200,
298–309. https://doi.org/10.1006/jcis.1997.5401.
Mészáros, R., Jobbik, A., Varga, G., Bárány, S., 2019. Electrosurface properties of Na-
bentonite particles in electrolytes and surfactants solution. Appl. Clay Sci. 178,
105127. https://doi.org/10.1016/j.clay.2019.105127.
Miller, S.E., Low, P.F., 1990. Characterization of the electrical double layer of
montmorillonite. Langmuir 6, 572–578. https://doi.org/10.1021/la00093a010.
Moussa, C., Xu, J., Wang, X., Zhang, J., Chen, Z., Li, X., 2017. Molecular dynamics
simulation of hydrated Na-montmorillonite with inorganic salts addition at high
temperature and high pressure. Appl. Clay Sci. 146, 206–215. https://doi.org/
10.1016/j.clay.2017.05.045.
Newton, A.G., Kwon, K.D., Cheong, D.K., 2016. Edge structure of montmorillonite from
atomistic simulations. Minerals 6 (25).
Newton, A.G., Kwon, K.D., Cheong, D.K., 2017. Na-montmorillonite edge structure and
surface complexes: an atomistic perspective. Minerals 7 (78).
Nishimura, S., Yao, K., Kodama, M., Imai, Y., et al., 2002. Electrokinetc study of synthetic
smectites by flat plate streaming potential technique. Langmuir. 18, 188–193.
Ou, C.Y., Chien, S.C., Yang, C.C., Chen, C.T., 2015. Mechanism of soil cementation by
electroosmotic chemical treatment. Appl. Clay Sci. 104, 135–142. https://doi.org/
10.1016/j.clay.2014.11.020.
Panagiotou, G.D., Petsi, T., Bourikas, K., Garoufalis, C.S., et al., 2008. Mapping the
surface (hydr)oxo-groups of titanium oxide and its interface with an aqueous
solution: the state of the art and a new approach. Adv. Colloid Interface 142, 20–42.
Park, C., Fenter, P.A., Nagy, K.L., Sturchio, N.C., 2006. Hydration and distribution of ions
at the mica-water interface. Phys. Rev. Lett. 97, 1–4. https://doi.org/10.1103/
PhysRevLett.97.016101.
Park, S.H., Sposito, G., 2002. Structure of Water Adsorbed on a Mica Surface. Phys. Rev.
Lett. 89, 8–10. https://doi.org/10.1103/PhysRevLett.89.085501.
Parsons, D.F., Ninham, B.W., 2010. Charge reversal of surfaces in divalent electrolytes:
the role of ionic dispersion interactions. Langmuir 26, 6430–6436. https://doi.org/
10.1021/la9041265.
Peng, J., Ye, H., Alshawabkeh, A.N., 2015. Soil improvement by electroosmotic grouting
of saline solutions with vacuum drainage at the cathode. Appl. Clay Sci. 114, 53–60.
https://doi.org/10.1016/j.clay.2015.05.012.
Plimpton, S., 1995. Fast parallel algorithms for short-range molecular dynamics.
J. Comput. Phys. https://doi.org/10.1006/jcph.1995.1039.
Předota, M., Machesky, M.L., Wesolowski, D.J., 2016. Molecular origins of the zeta
potential. Langmuir 32, 10189–10198. https://doi.org/10.1021/acs.
langmuir.6b02493.
Rezaei, M., Azimian, A.R., Pishevar, A.R., 2018. Surface charge-dependent
hydrodynamic properties of an electroosmotic slip flow. Phys. Chem. Chem. Phys.
20, 30365–30375. https://doi.org/10.1039/c8cp06408c.
Ricci, M., Spijker, P., Stellacci, F., Molinari, J.F., Voïtchovsky, K., 2013. Direct
visualization of single ions in the Stern layer of calcite. Langmuir 29, 2207–2216.
https://doi.org/10.1021/la3044736.
Sahai, N., Sverjensky, D.A., 1997. Solvation and electrostatic model for specific
electrolyte adsorption. Geochim. Cosmochim. Acta 61, 2827–2848. https://doi.org/
10.1016/S0016-7037(97)00127-0.
Şans, B.E., Güven, O., Esenli, F., Çelik, M.S., 2017. Contribution of cations and layer
charges in the smectite structure on zeta potential of Ca-bentonites. Appl. Clay Sci.
143, 415–421. https://doi.org/10.1016/j.clay.2017.04.016.
H. Pei and S. Zhang
Applied Clay Science 212 (2021) 106212
10
Schmickler, W., Henderson, D., 1986. New models for the structure of the
electrochemical interface. Prog. Surf. Sci. 22, 323–420.
Shang, J.Q., 1997. Zeta potential and electroosmotic permeability of clay soils. Can.
Geotech. J. 34, 627–631. https://doi.org/10.1139/t97-28.
Shang, J.Q., Lo, K.Y., Quigley, R.M., 1994. Quantitative determination of potential
distribution in Stern-Gouy double-layer model. Can. Geotech. J. 31, 624–636.
https://doi.org/10.1139/t94-075.
Shen, X.Y., Bourg, I.C., 2021. Molecular dynamics simulations of the colloidal interaction
between smectite clay nanoparticles in liquid water. J. Colloid Interface Sci. 584,
610–621.
Siboulet, B., Hocine, S., Hartkamp, R., Dufreche, J.F., 2017. Scrutinizing electro-osmosis
and surface conductivity with molecular dynamics. J. Phys. Chem. C 121 (12),
6756–6769.
Sondi, I., Bišćan, J., Pravdić, V., 1996. Electrokinetics of pure clay minerals revisited.
J. Colloid Interface Sci. 178, 514–522. https://doi.org/10.1006/jcis.1996.0146.
Sverjensky, D.A., 2001. Interpretation and prediction of triple-layer model capacitances
and the structure of the oxide-electrolyte-water interface. Geochim. Cosmochim.
Acta 65, 3643–3655. https://doi.org/10.1016/S0016-7037(01)00709-8.
Sverjensky, D.A., 2005. Prediction of surface charge on oxides in salt solutions: Revisions
for 1:1 (M+L-) electrolytes. Geochim. Cosmochim. Acta 69, 225–257. https://doi.
org/10.1016/j.gca.2004.05.040.
Tombácz, E., Szekeres, M., 2004. Colloidal behavior of aqueous montmorillonite
suspensions: the specific role of pH in the presence of indifferent electrolytes. Appl.
Clay Sci. 27, 75–94. https://doi.org/10.1016/j.clay.2004.01.001.
Tournassat, C., Chapron, Y., Leroy, P., Bizi, M., Boulahya, F., 2009. Comparison of
molecular dynamics simulations with triple layer and modified Gouy-Chapman
models in a 0.1 M NaCl-montmorillonite system. J. Colloid Interface Sci. 339,
533–541. https://doi.org/10.1016/j.jcis.2009.06.051.
Vasu, N., De, S., 2010. Electroosmotic flow of power-law fluids at high zeta potentials.
Colloids Surf. A Physicochem. Eng. Asp. 368, 44–52. https://doi.org/10.1016/j.
colsurfa.2010.07.014.
Wander, M.C.F., Clark, A.E., 2008. Structural and Dielectric Properties of Quartz – Water
Interfaces 19986–19994. https://doi.org/10.1021/jp803642c.
Wang, J., Wang, M., Li, Z., 2006. Lattice Poisson-Boltzmann simulations of electro-
osmotic flows in microchannels. J. Colloid Interface Sci. 296, 729–736. https://doi.
org/10.1016/j.jcis.2005.09.042.
Wolthers, M., Charlet, L., Cappellen, P.V., 2008. The surface chemistry of divalent metal
carbonate minerals; a critical assessment of surface charge and poential data using
the charge distribution multi-site ion complexation model. Am. J. Sci. 308, 905–941.
Xie, Y., Fu, L., Niehaus, T., Joly, L., 2020. Liquid-solid slip on charged walls: the dramatic
impact of charge distribution. Phys. Rev. Lett. 125, 1–7. https://doi.org/10.1103/
PhysRevLett.125.014501.
Xu, R., 1998. Shear plane and hydrodynamic diameter of microspheres in suspension.
Langmuir 14, 2593–2597. https://doi.org/10.1021/la971404g.
Yates, D.E., Levine, S., Healy, T.W., 1974. Site-binding model of the electrical double
layer at the oxide/water interface. J. Chem. Soc. Farady Trans. 70, 1807.
Zhang, C., Liu, Z., Dong, Y., 2017. Effects of adsorptive water on the rupture of nanoscale
liquid bridges. Appl. Clay Sci. 146, 487–494. https://doi.org/10.1016/j.
clay.2017.07.002.
Zhang, H., Hassanali, A.A., Shin, Y.K., Knight, C., Singer, S.J., 2011. The water-
amorphous silica interface: analysis of the Stern layer and surface conduction.
J. Chem. Phys. 134 https://doi.org/10.1063/1.3510536.
Zhang, S., Pei, H., 2020. Rate of capillary rise in quartz nanochannels considering the
dynamic contact angle by using molecular dynamics. Powder Technol. 372,
477–485. https://doi.org/10.1016/j.powtec.2020.06.018.
Zhao, Q., Choo, H., Bhatt, A., Burns, S.E., Bate, B., 2017. Review of the fundamental
geochemical and physical behaviors of organoclays in barrier applications. Appl.
Clay Sci. 142, 2–20. https://doi.org/10.1016/j.clay.2016.11.024.
H. Pei and S. Zhang

More Related Content

What's hot

Study of Microstructural, Electrical and Dielectric Properties of La0.9Pb0.1M...
Study of Microstructural, Electrical and Dielectric Properties of La0.9Pb0.1M...Study of Microstructural, Electrical and Dielectric Properties of La0.9Pb0.1M...
Study of Microstructural, Electrical and Dielectric Properties of La0.9Pb0.1M...Scientific Review SR
 
Interrelationships between Characteristic Lengths of Local Scour Hole
Interrelationships between Characteristic Lengths of Local Scour HoleInterrelationships between Characteristic Lengths of Local Scour Hole
Interrelationships between Characteristic Lengths of Local Scour Holedrboon
 
Literature survey modeling of microfluidics devices
Literature survey modeling of microfluidics devicesLiterature survey modeling of microfluidics devices
Literature survey modeling of microfluidics devicesAweshkumarsingh
 
Numerical Model of Microwave Heating in a Saturated Non-Uniform Porosity Medi...
Numerical Model of Microwave Heating in a Saturated Non-Uniform Porosity Medi...Numerical Model of Microwave Heating in a Saturated Non-Uniform Porosity Medi...
Numerical Model of Microwave Heating in a Saturated Non-Uniform Porosity Medi...drboon
 
Viscoelasticity of Polyodmain to Monodomain transition in liquid crystal elas...
Viscoelasticity of Polyodmain to Monodomain transition in liquid crystal elas...Viscoelasticity of Polyodmain to Monodomain transition in liquid crystal elas...
Viscoelasticity of Polyodmain to Monodomain transition in liquid crystal elas...Valeria Vasconcellos
 
Fall MRS 2013 - MgO grain boundaries structure and transport
Fall MRS 2013 - MgO grain boundaries structure and transportFall MRS 2013 - MgO grain boundaries structure and transport
Fall MRS 2013 - MgO grain boundaries structure and transportKedarnath Kolluri
 
JChemPhy_2014
JChemPhy_2014JChemPhy_2014
JChemPhy_2014Bhuvana S
 
Modelling Simul. Mater. Sci. Eng. 8 (2000) 445 C A And C P F E M
Modelling  Simul.  Mater.  Sci.  Eng. 8 (2000) 445  C A And  C P  F E MModelling  Simul.  Mater.  Sci.  Eng. 8 (2000) 445  C A And  C P  F E M
Modelling Simul. Mater. Sci. Eng. 8 (2000) 445 C A And C P F E MDierk Raabe
 
Electrical resistivity data interpretation for groundwater detection in titta...
Electrical resistivity data interpretation for groundwater detection in titta...Electrical resistivity data interpretation for groundwater detection in titta...
Electrical resistivity data interpretation for groundwater detection in titta...eSAT Journals
 
Liquid crystalline semiconductors
Liquid crystalline semiconductorsLiquid crystalline semiconductors
Liquid crystalline semiconductorsSpringer
 
Application of electrical resistivity tomography (ert) and arial photographs ...
Application of electrical resistivity tomography (ert) and arial photographs ...Application of electrical resistivity tomography (ert) and arial photographs ...
Application of electrical resistivity tomography (ert) and arial photographs ...Alexander Decker
 
Bubble-Propelled Micromotors for Enhanced Transport of Passive Tracers
Bubble-Propelled Micromotors for Enhanced Transport of Passive TracersBubble-Propelled Micromotors for Enhanced Transport of Passive Tracers
Bubble-Propelled Micromotors for Enhanced Transport of Passive TracersMichael Galarnyk
 
Multi scale modeling of micro-coronas
Multi scale modeling of micro-coronasMulti scale modeling of micro-coronas
Multi scale modeling of micro-coronasFa-Gung Fan
 
STM Observation of the Si(111) - (7×7) Reconstructed Surface Modified by Exce...
STM Observation of the Si(111) - (7×7) Reconstructed Surface Modified by Exce...STM Observation of the Si(111) - (7×7) Reconstructed Surface Modified by Exce...
STM Observation of the Si(111) - (7×7) Reconstructed Surface Modified by Exce...IJECEIAES
 

What's hot (20)

Ch5
Ch5Ch5
Ch5
 
Study of Microstructural, Electrical and Dielectric Properties of La0.9Pb0.1M...
Study of Microstructural, Electrical and Dielectric Properties of La0.9Pb0.1M...Study of Microstructural, Electrical and Dielectric Properties of La0.9Pb0.1M...
Study of Microstructural, Electrical and Dielectric Properties of La0.9Pb0.1M...
 
Interrelationships between Characteristic Lengths of Local Scour Hole
Interrelationships between Characteristic Lengths of Local Scour HoleInterrelationships between Characteristic Lengths of Local Scour Hole
Interrelationships between Characteristic Lengths of Local Scour Hole
 
Literature survey modeling of microfluidics devices
Literature survey modeling of microfluidics devicesLiterature survey modeling of microfluidics devices
Literature survey modeling of microfluidics devices
 
Numerical Model of Microwave Heating in a Saturated Non-Uniform Porosity Medi...
Numerical Model of Microwave Heating in a Saturated Non-Uniform Porosity Medi...Numerical Model of Microwave Heating in a Saturated Non-Uniform Porosity Medi...
Numerical Model of Microwave Heating in a Saturated Non-Uniform Porosity Medi...
 
Fractals and Rocks
Fractals and RocksFractals and Rocks
Fractals and Rocks
 
Viscoelasticity of Polyodmain to Monodomain transition in liquid crystal elas...
Viscoelasticity of Polyodmain to Monodomain transition in liquid crystal elas...Viscoelasticity of Polyodmain to Monodomain transition in liquid crystal elas...
Viscoelasticity of Polyodmain to Monodomain transition in liquid crystal elas...
 
Diffusion
DiffusionDiffusion
Diffusion
 
ACS Applied & Mat Interfaces 2015
ACS Applied & Mat Interfaces 2015ACS Applied & Mat Interfaces 2015
ACS Applied & Mat Interfaces 2015
 
Fall MRS 2013 - MgO grain boundaries structure and transport
Fall MRS 2013 - MgO grain boundaries structure and transportFall MRS 2013 - MgO grain boundaries structure and transport
Fall MRS 2013 - MgO grain boundaries structure and transport
 
JChemPhy_2014
JChemPhy_2014JChemPhy_2014
JChemPhy_2014
 
Kronfeld asc2009-li
Kronfeld asc2009-liKronfeld asc2009-li
Kronfeld asc2009-li
 
Geophysical exploration
Geophysical explorationGeophysical exploration
Geophysical exploration
 
Modelling Simul. Mater. Sci. Eng. 8 (2000) 445 C A And C P F E M
Modelling  Simul.  Mater.  Sci.  Eng. 8 (2000) 445  C A And  C P  F E MModelling  Simul.  Mater.  Sci.  Eng. 8 (2000) 445  C A And  C P  F E M
Modelling Simul. Mater. Sci. Eng. 8 (2000) 445 C A And C P F E M
 
Electrical resistivity data interpretation for groundwater detection in titta...
Electrical resistivity data interpretation for groundwater detection in titta...Electrical resistivity data interpretation for groundwater detection in titta...
Electrical resistivity data interpretation for groundwater detection in titta...
 
Liquid crystalline semiconductors
Liquid crystalline semiconductorsLiquid crystalline semiconductors
Liquid crystalline semiconductors
 
Application of electrical resistivity tomography (ert) and arial photographs ...
Application of electrical resistivity tomography (ert) and arial photographs ...Application of electrical resistivity tomography (ert) and arial photographs ...
Application of electrical resistivity tomography (ert) and arial photographs ...
 
Bubble-Propelled Micromotors for Enhanced Transport of Passive Tracers
Bubble-Propelled Micromotors for Enhanced Transport of Passive TracersBubble-Propelled Micromotors for Enhanced Transport of Passive Tracers
Bubble-Propelled Micromotors for Enhanced Transport of Passive Tracers
 
Multi scale modeling of micro-coronas
Multi scale modeling of micro-coronasMulti scale modeling of micro-coronas
Multi scale modeling of micro-coronas
 
STM Observation of the Si(111) - (7×7) Reconstructed Surface Modified by Exce...
STM Observation of the Si(111) - (7×7) Reconstructed Surface Modified by Exce...STM Observation of the Si(111) - (7×7) Reconstructed Surface Modified by Exce...
STM Observation of the Si(111) - (7×7) Reconstructed Surface Modified by Exce...
 

Similar to Pei, mds zeta potential shear plane montmorillonite appl clay sci. 2021

Constraints on Ceres’ internal structure and evolution from its shape and gra...
Constraints on Ceres’ internal structure and evolution from its shape and gra...Constraints on Ceres’ internal structure and evolution from its shape and gra...
Constraints on Ceres’ internal structure and evolution from its shape and gra...Sérgio Sacani
 
Lattice Energy LLC - Two Facets of W-L Theorys LENR-active Sites Supported b...
Lattice Energy LLC -  Two Facets of W-L Theorys LENR-active Sites Supported b...Lattice Energy LLC -  Two Facets of W-L Theorys LENR-active Sites Supported b...
Lattice Energy LLC - Two Facets of W-L Theorys LENR-active Sites Supported b...Lewis Larsen
 
Computational electromagnetics in plasmonic nanostructures
Computational electromagnetics in plasmonic nanostructuresComputational electromagnetics in plasmonic nanostructures
Computational electromagnetics in plasmonic nanostructuresAliakbarMonfared1
 
SURFACE POLARITONS IN GAAS/ALGAAS/LH HETROJUNCTION STRUCTURE IN A HIGH MAGNET...
SURFACE POLARITONS IN GAAS/ALGAAS/LH HETROJUNCTION STRUCTURE IN A HIGH MAGNET...SURFACE POLARITONS IN GAAS/ALGAAS/LH HETROJUNCTION STRUCTURE IN A HIGH MAGNET...
SURFACE POLARITONS IN GAAS/ALGAAS/LH HETROJUNCTION STRUCTURE IN A HIGH MAGNET...ijrap
 
Compositional and thermal state of the lower mantle from joint 3D inversion w...
Compositional and thermal state of the lower mantle from joint 3D inversion w...Compositional and thermal state of the lower mantle from joint 3D inversion w...
Compositional and thermal state of the lower mantle from joint 3D inversion w...Sérgio Sacani
 
The Effect of High Zeta Potentials on the Flow Hydrodynamics in Parallel-Plat...
The Effect of High Zeta Potentials on the Flow Hydrodynamics in Parallel-Plat...The Effect of High Zeta Potentials on the Flow Hydrodynamics in Parallel-Plat...
The Effect of High Zeta Potentials on the Flow Hydrodynamics in Parallel-Plat...CSCJournals
 
Finding the Spontaneous/Self Potential of the Surface
Finding the Spontaneous/Self Potential of the SurfaceFinding the Spontaneous/Self Potential of the Surface
Finding the Spontaneous/Self Potential of the SurfaceIRJESJOURNAL
 
Stability Of Magnetostatic Surface Waves In A Semiconductor-Ferrite-Left-Hand...
Stability Of Magnetostatic Surface Waves In A Semiconductor-Ferrite-Left-Hand...Stability Of Magnetostatic Surface Waves In A Semiconductor-Ferrite-Left-Hand...
Stability Of Magnetostatic Surface Waves In A Semiconductor-Ferrite-Left-Hand...ijrap
 
Structural, electronic, elastic, optical and thermodynamical properties of zi...
Structural, electronic, elastic, optical and thermodynamical properties of zi...Structural, electronic, elastic, optical and thermodynamical properties of zi...
Structural, electronic, elastic, optical and thermodynamical properties of zi...Alexander Decker
 
Topology of charge density from pseudopotential density functional theory cal...
Topology of charge density from pseudopotential density functional theory cal...Topology of charge density from pseudopotential density functional theory cal...
Topology of charge density from pseudopotential density functional theory cal...Alexander Decker
 
10.1016-j.mssp.2014.10.034-Graphene nanosheets as electrode materials for sup...
10.1016-j.mssp.2014.10.034-Graphene nanosheets as electrode materials for sup...10.1016-j.mssp.2014.10.034-Graphene nanosheets as electrode materials for sup...
10.1016-j.mssp.2014.10.034-Graphene nanosheets as electrode materials for sup...Mahdi Robat Sarpoushi
 
Stability Of Magnetostatic Surface Waves In A Semiconductor-Ferrite-Left-Hand...
Stability Of Magnetostatic Surface Waves In A Semiconductor-Ferrite-Left-Hand...Stability Of Magnetostatic Surface Waves In A Semiconductor-Ferrite-Left-Hand...
Stability Of Magnetostatic Surface Waves In A Semiconductor-Ferrite-Left-Hand...ijrap
 
11.electronic properties of nanostructured quantum dots
11.electronic properties of nanostructured quantum dots11.electronic properties of nanostructured quantum dots
11.electronic properties of nanostructured quantum dotsAlexander Decker
 
Electronic properties of nanostructured quantum dots
Electronic properties of nanostructured quantum dotsElectronic properties of nanostructured quantum dots
Electronic properties of nanostructured quantum dotsAlexander Decker
 
Piezoelectricity : Think Locally, Act Globally
Piezoelectricity : Think Locally, Act GloballyPiezoelectricity : Think Locally, Act Globally
Piezoelectricity : Think Locally, Act GloballyZaahir Salam
 

Similar to Pei, mds zeta potential shear plane montmorillonite appl clay sci. 2021 (20)

Constraints on Ceres’ internal structure and evolution from its shape and gra...
Constraints on Ceres’ internal structure and evolution from its shape and gra...Constraints on Ceres’ internal structure and evolution from its shape and gra...
Constraints on Ceres’ internal structure and evolution from its shape and gra...
 
1-s2.0-S1369800114000055-main
1-s2.0-S1369800114000055-main1-s2.0-S1369800114000055-main
1-s2.0-S1369800114000055-main
 
Ikard et al 2013a
Ikard et al 2013aIkard et al 2013a
Ikard et al 2013a
 
Lattice Energy LLC - Two Facets of W-L Theorys LENR-active Sites Supported b...
Lattice Energy LLC -  Two Facets of W-L Theorys LENR-active Sites Supported b...Lattice Energy LLC -  Two Facets of W-L Theorys LENR-active Sites Supported b...
Lattice Energy LLC - Two Facets of W-L Theorys LENR-active Sites Supported b...
 
Oe3424742482
Oe3424742482Oe3424742482
Oe3424742482
 
Computational electromagnetics in plasmonic nanostructures
Computational electromagnetics in plasmonic nanostructuresComputational electromagnetics in plasmonic nanostructures
Computational electromagnetics in plasmonic nanostructures
 
SURFACE POLARITONS IN GAAS/ALGAAS/LH HETROJUNCTION STRUCTURE IN A HIGH MAGNET...
SURFACE POLARITONS IN GAAS/ALGAAS/LH HETROJUNCTION STRUCTURE IN A HIGH MAGNET...SURFACE POLARITONS IN GAAS/ALGAAS/LH HETROJUNCTION STRUCTURE IN A HIGH MAGNET...
SURFACE POLARITONS IN GAAS/ALGAAS/LH HETROJUNCTION STRUCTURE IN A HIGH MAGNET...
 
Spr in a thin metal film
Spr in a thin metal filmSpr in a thin metal film
Spr in a thin metal film
 
Compositional and thermal state of the lower mantle from joint 3D inversion w...
Compositional and thermal state of the lower mantle from joint 3D inversion w...Compositional and thermal state of the lower mantle from joint 3D inversion w...
Compositional and thermal state of the lower mantle from joint 3D inversion w...
 
The Effect of High Zeta Potentials on the Flow Hydrodynamics in Parallel-Plat...
The Effect of High Zeta Potentials on the Flow Hydrodynamics in Parallel-Plat...The Effect of High Zeta Potentials on the Flow Hydrodynamics in Parallel-Plat...
The Effect of High Zeta Potentials on the Flow Hydrodynamics in Parallel-Plat...
 
Finding the Spontaneous/Self Potential of the Surface
Finding the Spontaneous/Self Potential of the SurfaceFinding the Spontaneous/Self Potential of the Surface
Finding the Spontaneous/Self Potential of the Surface
 
Stability Of Magnetostatic Surface Waves In A Semiconductor-Ferrite-Left-Hand...
Stability Of Magnetostatic Surface Waves In A Semiconductor-Ferrite-Left-Hand...Stability Of Magnetostatic Surface Waves In A Semiconductor-Ferrite-Left-Hand...
Stability Of Magnetostatic Surface Waves In A Semiconductor-Ferrite-Left-Hand...
 
Structural, electronic, elastic, optical and thermodynamical properties of zi...
Structural, electronic, elastic, optical and thermodynamical properties of zi...Structural, electronic, elastic, optical and thermodynamical properties of zi...
Structural, electronic, elastic, optical and thermodynamical properties of zi...
 
Topology of charge density from pseudopotential density functional theory cal...
Topology of charge density from pseudopotential density functional theory cal...Topology of charge density from pseudopotential density functional theory cal...
Topology of charge density from pseudopotential density functional theory cal...
 
10.1016-j.mssp.2014.10.034-Graphene nanosheets as electrode materials for sup...
10.1016-j.mssp.2014.10.034-Graphene nanosheets as electrode materials for sup...10.1016-j.mssp.2014.10.034-Graphene nanosheets as electrode materials for sup...
10.1016-j.mssp.2014.10.034-Graphene nanosheets as electrode materials for sup...
 
Eg35750753
Eg35750753Eg35750753
Eg35750753
 
Stability Of Magnetostatic Surface Waves In A Semiconductor-Ferrite-Left-Hand...
Stability Of Magnetostatic Surface Waves In A Semiconductor-Ferrite-Left-Hand...Stability Of Magnetostatic Surface Waves In A Semiconductor-Ferrite-Left-Hand...
Stability Of Magnetostatic Surface Waves In A Semiconductor-Ferrite-Left-Hand...
 
11.electronic properties of nanostructured quantum dots
11.electronic properties of nanostructured quantum dots11.electronic properties of nanostructured quantum dots
11.electronic properties of nanostructured quantum dots
 
Electronic properties of nanostructured quantum dots
Electronic properties of nanostructured quantum dotsElectronic properties of nanostructured quantum dots
Electronic properties of nanostructured quantum dots
 
Piezoelectricity : Think Locally, Act Globally
Piezoelectricity : Think Locally, Act GloballyPiezoelectricity : Think Locally, Act Globally
Piezoelectricity : Think Locally, Act Globally
 

Recently uploaded

Thermal Engineering Unit - I & II . ppt
Thermal Engineering  Unit - I & II . pptThermal Engineering  Unit - I & II . ppt
Thermal Engineering Unit - I & II . pptDineshKumar4165
 
AKTU Computer Networks notes --- Unit 3.pdf
AKTU Computer Networks notes ---  Unit 3.pdfAKTU Computer Networks notes ---  Unit 3.pdf
AKTU Computer Networks notes --- Unit 3.pdfankushspencer015
 
VIP Model Call Girls Kothrud ( Pune ) Call ON 8005736733 Starting From 5K to ...
VIP Model Call Girls Kothrud ( Pune ) Call ON 8005736733 Starting From 5K to ...VIP Model Call Girls Kothrud ( Pune ) Call ON 8005736733 Starting From 5K to ...
VIP Model Call Girls Kothrud ( Pune ) Call ON 8005736733 Starting From 5K to ...SUHANI PANDEY
 
BSides Seattle 2024 - Stopping Ethan Hunt From Taking Your Data.pptx
BSides Seattle 2024 - Stopping Ethan Hunt From Taking Your Data.pptxBSides Seattle 2024 - Stopping Ethan Hunt From Taking Your Data.pptx
BSides Seattle 2024 - Stopping Ethan Hunt From Taking Your Data.pptxfenichawla
 
The Most Attractive Pune Call Girls Manchar 8250192130 Will You Miss This Cha...
The Most Attractive Pune Call Girls Manchar 8250192130 Will You Miss This Cha...The Most Attractive Pune Call Girls Manchar 8250192130 Will You Miss This Cha...
The Most Attractive Pune Call Girls Manchar 8250192130 Will You Miss This Cha...ranjana rawat
 
Call Girls Pimpri Chinchwad Call Me 7737669865 Budget Friendly No Advance Boo...
Call Girls Pimpri Chinchwad Call Me 7737669865 Budget Friendly No Advance Boo...Call Girls Pimpri Chinchwad Call Me 7737669865 Budget Friendly No Advance Boo...
Call Girls Pimpri Chinchwad Call Me 7737669865 Budget Friendly No Advance Boo...roncy bisnoi
 
UNIT - IV - Air Compressors and its Performance
UNIT - IV - Air Compressors and its PerformanceUNIT - IV - Air Compressors and its Performance
UNIT - IV - Air Compressors and its Performancesivaprakash250
 
Vivazz, Mieres Social Housing Design Spain
Vivazz, Mieres Social Housing Design SpainVivazz, Mieres Social Housing Design Spain
Vivazz, Mieres Social Housing Design Spaintimesproduction05
 
Java Programming :Event Handling(Types of Events)
Java Programming :Event Handling(Types of Events)Java Programming :Event Handling(Types of Events)
Java Programming :Event Handling(Types of Events)simmis5
 
KubeKraft presentation @CloudNativeHooghly
KubeKraft presentation @CloudNativeHooghlyKubeKraft presentation @CloudNativeHooghly
KubeKraft presentation @CloudNativeHooghlysanyuktamishra911
 
Top Rated Pune Call Girls Budhwar Peth ⟟ 6297143586 ⟟ Call Me For Genuine Se...
Top Rated  Pune Call Girls Budhwar Peth ⟟ 6297143586 ⟟ Call Me For Genuine Se...Top Rated  Pune Call Girls Budhwar Peth ⟟ 6297143586 ⟟ Call Me For Genuine Se...
Top Rated Pune Call Girls Budhwar Peth ⟟ 6297143586 ⟟ Call Me For Genuine Se...Call Girls in Nagpur High Profile
 
PVC VS. FIBERGLASS (FRP) GRAVITY SEWER - UNI BELL
PVC VS. FIBERGLASS (FRP) GRAVITY SEWER - UNI BELLPVC VS. FIBERGLASS (FRP) GRAVITY SEWER - UNI BELL
PVC VS. FIBERGLASS (FRP) GRAVITY SEWER - UNI BELLManishPatel169454
 
Thermal Engineering -unit - III & IV.ppt
Thermal Engineering -unit - III & IV.pptThermal Engineering -unit - III & IV.ppt
Thermal Engineering -unit - III & IV.pptDineshKumar4165
 
ONLINE FOOD ORDER SYSTEM PROJECT REPORT.pdf
ONLINE FOOD ORDER SYSTEM PROJECT REPORT.pdfONLINE FOOD ORDER SYSTEM PROJECT REPORT.pdf
ONLINE FOOD ORDER SYSTEM PROJECT REPORT.pdfKamal Acharya
 
Extrusion Processes and Their Limitations
Extrusion Processes and Their LimitationsExtrusion Processes and Their Limitations
Extrusion Processes and Their Limitations120cr0395
 
Online banking management system project.pdf
Online banking management system project.pdfOnline banking management system project.pdf
Online banking management system project.pdfKamal Acharya
 

Recently uploaded (20)

Roadmap to Membership of RICS - Pathways and Routes
Roadmap to Membership of RICS - Pathways and RoutesRoadmap to Membership of RICS - Pathways and Routes
Roadmap to Membership of RICS - Pathways and Routes
 
(INDIRA) Call Girl Meerut Call Now 8617697112 Meerut Escorts 24x7
(INDIRA) Call Girl Meerut Call Now 8617697112 Meerut Escorts 24x7(INDIRA) Call Girl Meerut Call Now 8617697112 Meerut Escorts 24x7
(INDIRA) Call Girl Meerut Call Now 8617697112 Meerut Escorts 24x7
 
Thermal Engineering Unit - I & II . ppt
Thermal Engineering  Unit - I & II . pptThermal Engineering  Unit - I & II . ppt
Thermal Engineering Unit - I & II . ppt
 
AKTU Computer Networks notes --- Unit 3.pdf
AKTU Computer Networks notes ---  Unit 3.pdfAKTU Computer Networks notes ---  Unit 3.pdf
AKTU Computer Networks notes --- Unit 3.pdf
 
VIP Model Call Girls Kothrud ( Pune ) Call ON 8005736733 Starting From 5K to ...
VIP Model Call Girls Kothrud ( Pune ) Call ON 8005736733 Starting From 5K to ...VIP Model Call Girls Kothrud ( Pune ) Call ON 8005736733 Starting From 5K to ...
VIP Model Call Girls Kothrud ( Pune ) Call ON 8005736733 Starting From 5K to ...
 
BSides Seattle 2024 - Stopping Ethan Hunt From Taking Your Data.pptx
BSides Seattle 2024 - Stopping Ethan Hunt From Taking Your Data.pptxBSides Seattle 2024 - Stopping Ethan Hunt From Taking Your Data.pptx
BSides Seattle 2024 - Stopping Ethan Hunt From Taking Your Data.pptx
 
The Most Attractive Pune Call Girls Manchar 8250192130 Will You Miss This Cha...
The Most Attractive Pune Call Girls Manchar 8250192130 Will You Miss This Cha...The Most Attractive Pune Call Girls Manchar 8250192130 Will You Miss This Cha...
The Most Attractive Pune Call Girls Manchar 8250192130 Will You Miss This Cha...
 
Call Girls Pimpri Chinchwad Call Me 7737669865 Budget Friendly No Advance Boo...
Call Girls Pimpri Chinchwad Call Me 7737669865 Budget Friendly No Advance Boo...Call Girls Pimpri Chinchwad Call Me 7737669865 Budget Friendly No Advance Boo...
Call Girls Pimpri Chinchwad Call Me 7737669865 Budget Friendly No Advance Boo...
 
UNIT - IV - Air Compressors and its Performance
UNIT - IV - Air Compressors and its PerformanceUNIT - IV - Air Compressors and its Performance
UNIT - IV - Air Compressors and its Performance
 
Vivazz, Mieres Social Housing Design Spain
Vivazz, Mieres Social Housing Design SpainVivazz, Mieres Social Housing Design Spain
Vivazz, Mieres Social Housing Design Spain
 
Java Programming :Event Handling(Types of Events)
Java Programming :Event Handling(Types of Events)Java Programming :Event Handling(Types of Events)
Java Programming :Event Handling(Types of Events)
 
KubeKraft presentation @CloudNativeHooghly
KubeKraft presentation @CloudNativeHooghlyKubeKraft presentation @CloudNativeHooghly
KubeKraft presentation @CloudNativeHooghly
 
Top Rated Pune Call Girls Budhwar Peth ⟟ 6297143586 ⟟ Call Me For Genuine Se...
Top Rated  Pune Call Girls Budhwar Peth ⟟ 6297143586 ⟟ Call Me For Genuine Se...Top Rated  Pune Call Girls Budhwar Peth ⟟ 6297143586 ⟟ Call Me For Genuine Se...
Top Rated Pune Call Girls Budhwar Peth ⟟ 6297143586 ⟟ Call Me For Genuine Se...
 
(INDIRA) Call Girl Aurangabad Call Now 8617697112 Aurangabad Escorts 24x7
(INDIRA) Call Girl Aurangabad Call Now 8617697112 Aurangabad Escorts 24x7(INDIRA) Call Girl Aurangabad Call Now 8617697112 Aurangabad Escorts 24x7
(INDIRA) Call Girl Aurangabad Call Now 8617697112 Aurangabad Escorts 24x7
 
PVC VS. FIBERGLASS (FRP) GRAVITY SEWER - UNI BELL
PVC VS. FIBERGLASS (FRP) GRAVITY SEWER - UNI BELLPVC VS. FIBERGLASS (FRP) GRAVITY SEWER - UNI BELL
PVC VS. FIBERGLASS (FRP) GRAVITY SEWER - UNI BELL
 
Thermal Engineering -unit - III & IV.ppt
Thermal Engineering -unit - III & IV.pptThermal Engineering -unit - III & IV.ppt
Thermal Engineering -unit - III & IV.ppt
 
Call Girls in Ramesh Nagar Delhi 💯 Call Us 🔝9953056974 🔝 Escort Service
Call Girls in Ramesh Nagar Delhi 💯 Call Us 🔝9953056974 🔝 Escort ServiceCall Girls in Ramesh Nagar Delhi 💯 Call Us 🔝9953056974 🔝 Escort Service
Call Girls in Ramesh Nagar Delhi 💯 Call Us 🔝9953056974 🔝 Escort Service
 
ONLINE FOOD ORDER SYSTEM PROJECT REPORT.pdf
ONLINE FOOD ORDER SYSTEM PROJECT REPORT.pdfONLINE FOOD ORDER SYSTEM PROJECT REPORT.pdf
ONLINE FOOD ORDER SYSTEM PROJECT REPORT.pdf
 
Extrusion Processes and Their Limitations
Extrusion Processes and Their LimitationsExtrusion Processes and Their Limitations
Extrusion Processes and Their Limitations
 
Online banking management system project.pdf
Online banking management system project.pdfOnline banking management system project.pdf
Online banking management system project.pdf
 

Pei, mds zeta potential shear plane montmorillonite appl clay sci. 2021

  • 1. Applied Clay Science 212 (2021) 106212 Available online 22 July 2021 0169-1317/© 2021 Elsevier B.V. All rights reserved. Molecular dynamics study on the zeta potential and shear plane of montmorillonite in NaCl solutions Huafu Pei , Siqi Zhang * School of Civil Engineering, State Key Lab of Coastal and Offshore Engineering, Dalian University of Technology, Dalian 116026, China A R T I C L E I N F O Keywords: Clay mineral Zeta potential Shear plane Electrical double layer Stern model Molecular dynamics A B S T R A C T Zeta potential and the position of the shear plane are key physical properties characterizing the behavior of clay minerals in the colloidal system, and have been applied in many fields such as electroosmotic consolidation and electro-kinetic decontamination. In the past decades, numerous studies have been conducted on measuring and calculating the zeta potential. Nevertheless, few researchers have been reported to achieve a systematic un­ derstanding and predicting the zeta potential and the shear plane's position, especially for clay particles. This paper provided a molecular dynamics (MD)-based method to determine the zeta potential and shear layer thickness simultaneously to fill the gap. In the paper, the structure of the electrical double layer (EDL) was investigated for montmorillonite mesopore containing NaCl electrolyte in the concentration of 0.20–1.30 mol/L. The density profiles of ion species were well predicted by the Stern model, combining the Stern potential's determination. The calculated zeta potential based on the electroosmotic velocity profile in nonequilibrium molecular dynamics (NEMD) simulation was improved by introducing the slip length and was found to be closely comparable to the experimental values. Furthermore, the results confirmed that the shear plane cannot be observed from the electroosmotic velocity profile and self-diffusion coefficient of species in the MD simulation. The position of the zeta potential was determined by the Stern model and triple-layer model (TLM), showing a certain distance from the Stern plane. The zeta position was found that has a linear relationship with the Debye length and linearly depends on the ionic strength in the log scale, in agreement with previous investigations. The findings provided a systematic insight into the electrical double layer structure, zeta potential and shear plane for montmorillonite. 1. Introduction As negatively charged particles, clay minerals are ubiquitous in the natural world. The clay-water interface strongly attracts cation species and forms the electrical double layer (EDL) structure (Lyklema, 1995). The structure greatly influences not only the electrochemical properties of clay minerals that characterize the efficiency of electroosmosis and electro-kinetic decontamination of clay (Shang, 1997; Ou et al., 2015) but also the behavior of aggregation, complexation and sedimentation (Tombácz and Szekeres, 2004; Peng et al., 2015). Serval EDL models including the Gouy-Chapman model, Stern-Gouy-Chapman model (Stern model in this paper for short) (Shang et al., 1994) and triple-layer model (TLM) (Leroy and Revil, 2004; Sverjensky, 2005) have been developed to describe the ions and electric potential distribution in the colloidal system. For the Stern model, the EDL consists of the Stern layer that the electrostatic potential within decreases linearly with distance and diffuse layer (Gouy layer) in which the potential follows the Poisson-Boltzmann (PB) equation. The Stern layer contains two kinds of adsorbate species including inner-sphere surface complexes (ISSC) and outer-sphere surface complexes (OSSC). In the TLM, the electrostatic potential drops linearly among the 0-, β- and d-planes controlled by two capacitance factors, while the potential is modeled with the PB equation beyond the d-planes. The detailed introductions of the Stern model and TLM are shown in Supporting Information. Zeta potential at the shear plane is a significant macroscopic mea­ surement used to estimate the stability of colloidal systems and numerous electro-kinetic applications (Şans et al., 2017; Zhao et al., 2017). The electric potential at the supposed shear plane that is the boundary between an immobile layer strongly adsorbed by the particle surface and the mobile fluid. When investigating the nanoparticle agglomeration, zeta potential can evaluate the critical coagulation concentration (Hunter, 1981). The potential at the shear plane is used to * Corresponding author. E-mail address: zhangsiqi9315@mail.dlut.edu.cn (S. Zhang). Contents lists available at ScienceDirect Applied Clay Science journal homepage: www.elsevier.com/locate/clay https://doi.org/10.1016/j.clay.2021.106212 Received 25 January 2021; Received in revised form 6 July 2021; Accepted 13 July 2021
  • 2. Applied Clay Science 212 (2021) 106212 2 study the transport properties of clay minerals (Leroy et al., 2008). The shear plane is also an essential parameter that characterizes the hy­ drodynamic properties of colloidal systems. The shear layer (the space between the clay surface and shear plane) greatly impacts the hydro­ dynamic motion of pore fluid and suspended particles (Xu, 1998). As a boundary condition, the shear plane is used to calculate the electroos­ motic flow velocity (Wang et al., 2006; Vasu and De, 2010). In soil mechanics, water in the shear layer is regarded as tightly bound water (Li, 2016). Compared to several measurement techniques such as elec­ trophoresis, electroosmosis and electroacoustics of the zeta potential determination (Greenwood, 2003), to date, the measurement of the shear layer thickness mainly follows two methods. The primary method is based on the PB equation or modified PB equation. The potential profile in the EDL is obtained, and then the shear plane's position is calculated according to the experimental measurement of zeta potential (Ding et al., 2015; Liu et al., 2017). In another way, only a few in­ vestigations directly measured the shear layer thickness according to its hydrodynamic definition. In electrophoresis experiments, a surface of shear beyond the particle interface can be observed due to the electro­ static interactions between the nanoparticle and the applied electric field (Hunter, 1981). Besides, diffusion coefficients of the colloidal particles, counterions or water were measured to determine the shear plane thickness without applying the electric field (Xu, 1998; Jain et al., 2021). Bourikas et al. (2001) and Panagiotou et al. (2008) referred to the position by those two methods as the shear plane uniformly in common. However, the obtained position in the first method is actually the po­ sition of zeta potential rather than that of the shear plane. Henceforth, to distinguish the two positions, this study defines the zeta position measured by the first method, which indicates the potential at the calculated position corresponding to the zeta potential, and shear plane position measured by the second method, at which the streaming ve­ locity is zero. Although a number of observations have been achieved, based on the mentioned methods, the accurate position of the zeta potential in the EDL remains controversial. To simplify, many researchers have assumed that the zeta position is close to the Stern plane or even locate at it; namely, the capacity value between two planes is much large in the TLM (Hiemstra and Van Riemsdijk, 2006). However, Li et al. (2003) and Liu et al. (2017) demonstrated that the zeta position is much closer to the Gouy plane than the Stern plane after comparing the zeta potential with surface potential and Stern potential in montmorillonite-water systems. Molecular dynamics (MD) is a computational method that models the molecular structure and predicts the behaviors of materials at the atomic scale (Zhang and Pei, 2020). MD was used to explore the micro- mechanism of clay minerals (Moussa et al., 2017; Zhang et al., 2017; Liu et al., 2021). Chang et al. (1998), Parsons and Ninham (2010), and Bourg and Sposito (2011) successfully applied the MD technique to investigate the structure of EDL for clay-water systems. MD simulations can determine the profiles of species in the diffuse layer and bulk elec­ trolyte. Furthermore, MD can be utilized to simulate the electroosmotic flows between two planar-charged surfaces. Based on the electroosmosis simulation, scholars have also proposed methods to determine the zeta potential of materials in different electrolyte solutions (Předota et al., 2016; Biriukov et al., 2020). Therefore, MD is an appropriate tool to study the zeta potential and shear plane position of clay. In this paper, MD was performed to determine the zeta potential and position of the shear plane in NaCl-montmorillonite systems with different concentrations. The electric potential profiles, particularly for the Stern potential and zeta potential, were calculated based on MD and the EDL model. The shear layer thickness and zeta position were dis­ cussed based on several measurement methods in the paper. The ca­ pacitances in the TLM model were finally discussed. 2. Theoretical background 2.1. Stern model In the classical EDL theory, the Poisson-Boltzmann equation de­ scribes the distributions of ions in electrolyte solutions (Lyklema, 1995). d2 φ dz2 = − 4π ε0εr ∑ i c0 i viFexp ( − viFφ(z) RT ) (1) where φ(z) is the potential distribution in the diffuse layer; ε0 is vacuum dielectric constant; εr is the relative permittivity of bulk water; ci 0 is the ion concentration within the bulk solution; vi is the valence of the ith cation species. F is the Faraday constant; R is the gas constant. In the Stern model, by solving Eq. (1), the distribution of electric potential away from the clay surface in the Stern layer and diffuse layer can be expressed as (Shang et al., 1994): φ(z) = ⎧ ⎪ ⎪ ⎪ ⎪ ⎪ ⎨ ⎪ ⎪ ⎪ ⎪ ⎪ ⎩ φ0 − φ0 − φs ds z, z ≤ ds 2RT vF ln eκ(z− ds) + tanh vFφs 4RT eκ(z− ds) − tanh vFφs 4RT , z > ds (2) φ0 = σ Cs + φs (3) where Debye-Huckel parameter κ = ̅̅̅̅̅̅̅̅̅̅̅̅ 2v2F2c0 ε0εrRT √ (m− 1 ) represents the reciprocal of the thickness of the diffuse layer; ds is the thickness of the Stern layer; φsrepresents the Stern layer potential; φ0is the surface po­ tential and can be calculated for a given surface charge density σ and constant capacitance of the Stern layer Cs. Once ds and φs are deter­ mined, the potential at any distance from the clay surface in the EDL can be calculated. In other words, given a zeta potential, the position of the zeta potential can be located by the Stern model. Previous investigations (Bourg and Sposito, 2011; Ricci et al., 2013; Bourg et al., 2017; Hocine et al., 2016) have demonstrated that the Stern plane is located at the position of the OSSC. Therefore, the Stern layer thickness can be determined from the density profiles of cations in MD simulation. Li et al. (2004) and Hou et al. (2009) have proposed a method to calculate the potential at the top end of the diffuse layer based on the Poisson-Boltzmann equation. The potential is only determined by the mean concentration of counterions in the diffuse layer ci and the concentration in the equilibrium solution (Hou et al., 2009): Fφs = − 2RTln 1 − 0.5 [ − A ( ci ) + ̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅ A ( ci )2 − 10.873 √ ] 1 + 0.5 [ − A ( ci ) + ̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅ A ( ci )2 − 10.873 √ ] (4) where A ( ci ) = 3.7183 + 6.8731c0 i ci − c0 i (5) The derivation process of Eq. (4) is detailed in the Supporting In­ formation. Liu et al. (2015, 2016) have applied the method to calculate the Stern potential. For the first time, this method is combined with MD simulations to solve the potential distribution in the EDL. The average concentration of cations in the diffuse layer can be calculated from the density profiles in MD simulation after locating the onset of the diffuse layer. Thus, by means of Eqs. (2)-(5), the potential distribution is determined using only the cations density profile near the clay surface. H. Pei and S. Zhang
  • 3. Applied Clay Science 212 (2021) 106212 3 2.2. Zeta potential The electroosmotic velocity profile is governed by the Navier-Stokes (NS) equation together with the PB equation (Lorenz and Travesset, 2007): u(z) = ε0εr η Ex[φ(z) − ζ ] (6) where ζ is the zeta potential. When the fluid far away from the surface, the electroosmotic velocity is linearly related to zeta potential, namely the Helmholtz-Smoluchowski (HS) equation (Delgado et al., 2007): ζ = - ueoη ε0εrEx (7) where ueo represents the streaming velocity of the bulk solution under the electric field; η is the viscosity of the solution; Ex is the external electric field along the x-direction. In Eq. (6), the electroosmotic flow is defined as a plug-like flow under the no-slip condition. However, MD simulations of electroosmotic flows showed a significant slippage phe­ nomenon in previous literature (Rezaei et al., 2018; Xie et al., 2020). Dufreche et al. (2005) have investigated the electro-osmosis in Na- montmorillonite by MD simulations, which shows a slip length should be taken into account to describe the profile of electroosmotic flow. An overestimated value of the zeta potential may be obtained when considering the no-slip condition. A Navier-type slip condition is intro­ duced to describe the slip velocity of electroosmotic flow at the interface (Celebi and Beskok, 2018): u‖(z0) = b du dz ⃒ ⃒z=z0 = ε0εr η Exb dφ dz ⃒ ⃒z=z0 (8) where u‖(z0) is the slip velocity at the slip wall of z = z0; b is the slip length. Thus, zeta potential can be calculated as (Celebi and Beskok, 2018): ζ = η ε0εrEx ( − ueo + b dφ dz ⃒ ⃒z=z0 ) (9) Nonequilibrium molecular dynamics (NEMD) can be applied to simulate the electroosmotic flow in montmorillonite nanopore, and the zeta potential can be calculated by Eq. (9). Finally, the zeta position can be determined by the zeta potential together with the potential profile from the Stern model. 2.3. Triple-layer model The triple-layer model is a popular model proposed to characterize the electrochemical properties of colloidal systems, especially for clay- solution systems. Taking the example of montmorillonite in a NaCl electrolyte, the main equations of the TLM are expressed as follows (Tournassat et al., 2009; Leroy et al., 2015): Q0 + Qβ + Qd = 0 (10) Qd = − ̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅ 8ε0εrkbTci0NA √ × sinh ( Fφd 2RT ) (11) φβ = φd − Qd C2 (12) φ0 = φβ + Q0 C1 (13) KNa = - Q0 + Qβ Qβ aNaexp ( − Fφβ RT ) (14) where Q0, Qβ and Qd are the surface charge densities at the 0, β-plane and d-plane respectively and Q0is equal to the surface density of the basal surface of montmorillonite minus that of ISCC. φ0, φβ and φd are the corresponding potential. Notably, φd corresponds to zeta potential in TLM. kb is the Boltzmann constant.NA is Avogadro number. KNa is the equilibrium constant at β-plane. C1 and C2 are the capacitances within β-plane and between β-plane and d-plane. The TLM can be simplified as the basic Stern model by assuming φβ equal to φd. 2.4. Simulation methodology In this paper, the zeta potential and the shear plane of montmoril­ lonite in NaCl solution were investigated. The molecular model of montmorillonite sheets was constructed with the unit cell formula of Si8(Al3.25Mg0.75)O20(OH)4(Greathouse et al., 2015). Isomorphous sub­ stitution occurs in the octahedral sheet, that is, Mg2+ substitution for Al3+ . The clay layer used in the simulations consists of 4 × 4 unit cells and has the size 20.66 Å × 35.85 Å × 6.54 Å and a surface charge density of − 0.13C/m2 due to 12 Mg for Al substitutions. The layer was divided in two halves located at the top and bottom of the simulation box to form a slit-like pore with a width of 5.0 nm, which prevents overlapping of the EDL. Water molecules with the number of 1187 were added in the pore to reach a bulk density of 0.997 g/cm3 . Na+ was chosen as the coun­ terion. Counterions were randomly distributed in the pore to compen­ sate for the charge deficit of the clay layer. Besides, Na+ and Cl− atoms with the number of 2, 5, 11 and 22 were inserted in the water box, leading to concentrations in the center of 0.20, 0.37, 0.70 and 1.30 mol/ L. The initial configuration of montmorillonite in 0.20 mol/L NaCl is shown in Fig. 1. Periodical boundaries were set in all directions. The interactions of minerals-minerals and minerals-water were described by the ClayFF force field in this study (Cygan et al., 2004). The extended simple point charge (SPC/E) water model was utilized to calculate the water-water interactions. LAMMPS was used to perform all the MD simulations (Plimpton, 1995). The MD simulation was conducted in two parts. First, an equilibrium molecular dynamics (EMD) simulation was performed under NVT ensemble at 300 K. After a relaxation time of 10 ns, the trajectories of atoms were recorded every 1 ns for another 200 ns. To prevent un­ wanted movements of the clay particles, Al and Mg atoms in the clay layer were fixed during the simulations. The density profiles and mean square displacement of ion species and water molecules were obtained in EMD. An electric field of 5 × 108 V/m was then applied parallel to the clay surface acting on water molecules, cations and mobile surface atoms (surface O atoms) in the NEMD simulation. Notably, the intensity of the electric field is higher than the strength used in experiments, while it is in the common range in MD studies. The purpose of the high electric Fig. 1. Snapshot of MD model for a 50 Å mesopore constructed by the parallel montmorillonite surfaces containing 0.20 mol/L NaCl. The width of the pore is the distance between siloxane oxygens on each side of the clay layer. H. Pei and S. Zhang
  • 4. Applied Clay Science 212 (2021) 106212 4 field is to avoid excessive noise and obtain a reasonable accuracy. The influence of electric field intensity on NEMD was discussed in Sup­ porting Information. The system was relaxed by 20 ns under NVT ensemble at 300 K to reach the equilibrium state. The streaming velocity of water along the clay surface was recorded for 200 ns. The relative permittivity and viscosity of the SPC/E water model (72.4 and 7.29 × 10− 4 Pa⋅s) were used in Eq. (9) to calculate the zeta potential (Balasu­ bramanian et al., 1996; Gereben and Pusztai, 2011). In this study, a standard MD thermostat (Nosé-Hoover thermostat) was used in both EMD and NEMD simulations. However, the thermostat used in NEMD simulations is coupled only to the degrees of freedoms perpendicular to the flow direction. The Lennard Jones (LJ) interactions were truncated to 12 Å, and the Lorentz-Bertholet mixing rule was adopted for the LJ interactions between different atoms. The Coulomb interactions were calculated by the PPPM method (Hockney and Eastwood, 1988) with 99.99% accuracy. The time step of 1 fs was set to integrate Newton's motion equations in the simulations. 3. Results and discussion 3.1. Ion distributions and stern potential The density distributions of water along the aperture direction in the montmorillonite mesopore containing different NaCl solutions present similar tendencies, as shown in Fig. S2. It is noted that the location of surface O atoms in the montmorillonite layer is the origin of the z-axis. Pronounced water layers occur near the clay surface and dissipate beyond 10 Å, which is in agreement with water on the clay surfaces from previous investigations (Park and Sposito, 2002). The first peak of water density is located at 2.7 Å, close to the value for other MD simulations (2.7 ± 0.5 Å) (Bourg and Sposito, 2011) and X-ray experiments of mica (2.5 ± 0.2 Å) (Cheng et al., 2001). For each NaCl-montmorillonite sys­ tem, the density profiles of Na+ and Cl− as a function of distance from the clay surface are presented in Fig. 2. The convergence and error analysis of ions density profile were discussed in Supporting Informa­ tion. The density profile of Na+ close to the surface shows the first two peaks corresponding to the plane of ISSC and OSSC, respectively. The stern plane positions for 0.20, 0.37, 0.70 and 1.30 mol/L are determined as the distance of 4.3 Å away from the clay surface. The thickness of Stern layer in MD simulations of montmorillonite in mixed NaCl-CaCl2 solutions (4.35–4.55 Å) (Bourg and Sposito, 2011) and measurement values of divalent cations OSSCs on mica (4.52 ± 0.24 Å) by X-ray (Park et al., 2006) are close to the results in this study. The simulations show that the z-coordination of OSSC is independent of the ionic concentration. The diffuse layer region is determined based on the Debye length after locating the Stern plane. The mean concentration of cations in the diffuse layer can be calculated from its density profiles, which are listed in Table 1. The Stern potential of 0.20–1.30 mol/L NaCl solutions are calculated as − 80.2, − 67.0, − 52.0 and − 39.7 mV by Eq. (4), respec­ tively. The Stern potential decreases with increasing ion concentration, which is consistent with the EDL theory. Miller and Low (1990) used several methods combined with experiments and theories to determine the Stern potential of montmorillonite in distilled water. The Stern po­ tentials of Na-montmorillonite with the cation exchange capacity of 90 meq/100 g were measured in the range of − 55.5 to − 59.1 mV. The Stern potential of montmorillonite with − 0.1 to − 0.15C/m2 surface charge density in 0.1 mol/L NaCl solution was calculated in the range of − 79.5 to − 96.8 mV in theory (Liu et al., 2016). Compared with their results, the calculated Stern potential is within a reasonable range. Once the Stern potential is determined, the electric potential distri­ bution in the diffuse layer can be calculated by Eq. (2). The density profiles of ion species beyond the Stern plane predicted by the Stern model are consistent with MD results, as shown in Fig. 2. The Stern model with a simple way of the Stern potential determination can well match the molar density of diffuse swarm (DS) species, however, slightly underestimates the density of the OSSC. The difference at the Stern plane may be attributed to the PB equation neglecting some effects such as ion polarizability. However, this study still focuses on the classical EDL theory in this paper. A popular model named the modified Gouy- Chapman (MGC) model has been proposed to describe quantitatively the ion concentration profiles in the EDL (Tournassat et al., 2009; Le Crom et al., 2020). The main equations are expressed as: ⎧ ⎪ ⎪ ⎪ ⎨ ⎪ ⎪ ⎪ ⎩ Fφ(a) RT = − 2 × sinh− 1 ( F|σ| 2κε0εrRT ) Fφ(z) RT = 4 × tanh− 1 [ tanh ( Fφ(a) 4RT ) × exp( − κ(z − a) ) ] (15) where a is a distance from the surface. σ is the surface charge minus the ions charge within a. For montmorillonite in 0.20 M NaCl solution, the comparison of the GC, MGC and Stern model is shown in Fig. 3. The Na+ density profile predicted by the GC model has the maximum molar density at the clay surface, which shows a distinct difference from MD simulation. Although the profiles by the MGC model with different a are close to the MD result beyond 10 Å, the values of calculated profiles are in disagreement with each other near the clay surface. In addition, the predicted profiles of the MGC model at the Stern plane are less accurate than the Stern model. The density profile of the Stern model is slightly larger than that of the GC and MGC model beyond 10 Å. The fitted 0 5 10 15 20 25 0 1 2 3 4 5 6 7 8 Na+ by MD Na+ by Stern model Na + concentration (mol/L) Distance from the clay surface (Å) 0 1 2 Cl- by MD Cl- by Stern model Cl - concentration (mol/L) Stern plane (OSSC) 1.30 mol·L-1 Fig. 2. Molar density profiles of Na, Cl at the clay surface as obtained from the MD model of montmorillonite in 1.30 mol/L NaCl electrolyte. The light dash lines are the density profiles of Na and Cl predicted by the Stern model. The density profiles of 0.70, 0.37 and 0.20 mol/L NaCl electrolytes are shown in Fig. S3. Table 1 Calculated parameters in the Stern model for montmorillonite in NaCl solutions. The capacitance in the Stern layer is 1.0 F m− 2 . 0.20 mol/L 0.37 mol/L 0.70 mol/L 1.30 mol/L Mean concentration of Na+ in diffuse layer, c(mol/L) 1.43 1.94 2.55 3.50 Electroosmotic mobility of bulk water, μeo/Ex (m2 /V⋅s) 11.02 10.13 8.77 7.90 Thickness of Stern plane, ds (Å) 4.3 4.3 4.3 4.3 Position of zeta potential, dζ (Å) 7.2 6.8 6.5 5.8 Thickness of Gouy plane, dg (Å) 11.1 9.3 7.9 7.0 Surface potential, φ0 (mV) − 210.2 − 197.0 − 182.0 − 169.7 Stern potential, φs (mV) − 80.2 − 67.0 − 52.0 − 39.7 Zeta potential, ζ (mV) − 46.8 − 38.8 − 26.9 − 22.3 H. Pei and S. Zhang
  • 5. Applied Clay Science 212 (2021) 106212 5 parameter a seems to lack a clear physical significance resulting in an uncertain origin of the density profile, compared to the peak located at the Stern plane in the Stern model. Therefore, the Stern model combined with a calculated Stern potential has good accuracy in predicting ion density profiles in the EDL. 3.2. Electroosmotic flow and zeta potential Fig. 4 shows the electroosmotic velocity profile of water in mont­ morillonite pore containing 0.20 mol/L NaCl solutions. The streaming velocity follows the hydrodynamical Navier-Stokes equation (Eq. (6)) beyond the first peak of water density profile, whereas the velocity drops rapidly within that. A clear slip velocity can be observed at the first peak of water density profile. This is because only a small number of water molecules before the first peak of water density profile that strongly absorbed by the clay surface have different dynamics properties from the bulk water. The same observation has also been presented in the NEMD simulation of an electroosmotic flow of CsCl solution in a charged silica slit by Siboulet et al. (2017). Celebi and Beskok (2018); Celebi et al. (2019) have considered the shear plane is located at the first peak of the water density profile. Although the streaming velocity disobeys the NS equation, there is still a non-zero velocity within 2.7 Å in this simulation as observed in Fig. 4. Therefore, the first peak of the water density profile cannot be regarded as the shear plane simply. Siboulet et al. (2017) also demonstrated that the stagnant layer close to the surface belongs to the ion hydration sphere. This paper confirms that the flow domain starts at the first water density peak which is 2.7 Å away from the clay surface in the simulation. A function was proposed to fit the streaming velocity profile from MD simulations, which the deriva­ tion is shown in Supporting Information. Then the slip velocity and slip length are obtained based on the fitting velocity profile. Both slip length and slip mobility decreases with the concentration increases of NaCl, which the calculated values are listed in Table 2. Rezaei et al. (2018) also proved such tendency of slip length and slip velocity. The electro­ osmotic mobility (velocity profile normalized by field strength) along z- direction is exhibited in Fig. 5. The plug-like profiles for NaCl- montmorillonite systems are observed as expected. The average mobility at the plateau is significantly affected by the ionic concentra­ tion, decreasing with the increasing of concentration. The zeta potential of 0.20–1.30 mol/L NaCl solutions calculated by Eq. (9) are − 46.8, − 38.8, − 26.9 and − 22.3 mV respectively. Tournassat et al. (2009) measured the zeta potential of − 38 mV for montmorillonite (σ= − 0.11C/m2 ) in 0.12 molL− 1 NaCl solution. Mészáros et al. (2019) measured the zeta potential of bentonite (σ= − 0.14C/m2 ) in 0.1 mol/L NaCl solution as − 37.4 to − 47 mV in the range of 2–12 pH. The zeta potential of Na-montmorillonite (σ= − 0.16C/m2 , pH = 6.5) at 1.0 M was determined as − 17.9 mV (Sondi et al., 1996). It shows that the values of zeta potential in the NEMD simulation are slightly larger than to the measurements. The zeta potential of 0.20–1.30 mol/L NaCl so­ lutions are calculated as − 126.7, − 116.5, − 100.8 and − 90.9 mV respectively, without considering the slip length (Eq. (7)), which are significantly larger than the measured values. Therefore, the slip length should be considered when calculating the zeta potential to avoid an overestimated value. 3.3. Zeta position and shear plane position After obtaining potential distribution based on the Stern model and the zeta potential in the EDL, the zeta position can be calculated, as shown in Fig. 6. The zeta positions are calculated as 7.2, 6.8, 6.5 and 5.8 Å for montmorillonite at 0.20–1.30 mol/L. Results show that the posi­ tion of zeta potential cannot be identified as the Stern plane simply. The zeta position is closer to the Gouy plane than the Stern plane. For montmorillonite in NaCl solution, the zeta position from the clay surface decreases as the ionic concentration increases because of the Debye screening length. The TLM can also obtain the zeta position after the determination of zeta potential. The charge density at d-plane (zeta position) Qd can be calculated by Eq. (11). Besides, Qd is equal to subtracting the compen­ sated charge within the d-plane from the surface charge of the 0 5 10 15 20 25 0 1 2 3 4 5 6 0 5 10 15 20 25 0.00 0.05 0.10 0.15 0.20 Cl - concentration (mol/L) Na + concentration (mol/L) Distance from the clay surface (Å) MD Stern model GC model MGC a=0.21 nm MGC a=0.31 nm MGC a=0.43 nm MGC a=0.57 nm Fig. 3. The comparison among Na+ (and Cl− in the inserted graph) density profile for montmorillonite at NaCl concentration of 0.20 mol/L in MD simu­ lation and the predictions of the Gouy-Chapman model, Stern model and modified Gouy-Chapman model with fitted parameter a of 0.21, 0.31, 0.43 and 0.57 nm. 25 20 15 10 5 0 -5 0 5 10 15 20 25 30 35 40 45 50 55 60 0.0 0.5 1.0 1.5 2.0 2.5 w u|| (z0 ) Water density (g/cm 3 ) ueo b=2.9 Å z=z0 0 |z z du k dz 0 0 | r x z z d E dz uMD uslip ustick u(z) (m/s) z (Å) Fig. 4. Electroosmotic velocity profiles for montmorillonite in 0.20 mol/L NaCl electrolytes. uMD is the velocity of the fluid along the aperture direction from MD, uSlip from fitting uMD in a PB equation form as shown in Supporting In­ formation, and uStick by assuming a vanishing velocity at z = 2.7 Å. u‖(z0) is the slip velocity. b is the slip length. Table 2 The slip length and slip mobility for montmorillonite in NaCl electrolytes ob­ tained by MD and Eq. (8). Concentration (mol/L) Slip length (Å) Slip mobility (×10− 8 m2 /V s) 0.20 2.90 6.90 0.37 2.50 6.73 0.70 1.95 6.44 1.30 1.90 5.96 H. Pei and S. Zhang
  • 6. Applied Clay Science 212 (2021) 106212 6 montmorillonite basal surface (Eq. (10)). Therefore, by combining two equations, the d-plane positions are 6.0, 5.5, 5.3 and 4.8 Å for mont­ morillonite at 0.20, 0.37, 0.70 and 1.30 mol/L, as listed in Table 3. Results show that the position of the d-plane is located beyond the β-plane, in accordance with the TLM. The same tendency is exhibited in the TLM that the distance of zeta position from the clay surface de­ creases with electrolyte concentration. However, the distance between the zeta position and the Stern plane in the TLM is less than in the Stern model. Fig. 7 shows the relationship between the zeta position from the interface and diffuse layer, zeta position and ionic strength from MD simulations. The results from the Stern model and the TLM indicate that the zeta position from the clay surface has a strictly linear relationship with the Debye length, which agrees with the conclusion demonstrated by Charlton and Doherty (1999) through experiments. Moreover, the results also demonstrate that the zeta position (log scale) is linearly dependent on the log value of the ionic strength, in accordance with the conclusions of Bourikas et al. (2001) and Panagiotou et al. (2008). After the zeta position determination, the position of the shear plane was also investigated in this paper. In the Stern model, the position of zeta potential is located at the shear plane, at which the flow has a nil velocity under no-slip conditions. In the simulation, however, there is no stagnant layer within the zeta potential position. Similarly, more and more MD simulations have proved that the stagnant layer near a solid surface cannot be observed in the EDL (Lylema et al., 1998; Freund, 2002; Hartkamp et al., 2005; Lorenz et al., 2008; Zhang et al., 2011). The distinction between the standard hydrodynamics and the MD results may contribute to the following reason. In standard hydrodynamics and Stern model, the concept of shear plane strictly separating two neighbor regions with an infinite and a constant viscosity of water respectively is in contradiction with the findings that the viscosity changes continu­ ously near the interface by experiments and MD simulations (Schmickler and Henderson, 1986; Evans and Wennerström, 1999; Bazant et al., 2009; Hartkamp et al., 2018). Therefore, although the Stern model can predict well the static potential and ion distributions on the global EDL, the local physical properties in electrokinetic phenomenon cannot be accurately described by the NS equation and Stern model. Xu (1998) has proposed a method to measure the shear plane thickness of micellar particles based on their self-diffusion coefficients. The hydrodynamic radius Rh can be calculated by self-diffusion coeffi­ cient Ds using the Stokes-Einstein relationship: Ds = kBT 6πηRh (16) The thickness of the shear layer can be obtained by subtracting the particle radius from the hydrodynamic radius. Jain et al. (2021) deter­ mined the shear layer thickness of nano-particles by calculating the self- diffusion coefficient of cations in MD simulations. Therefore, the diffu­ sion coefficients of water oxygen and cations in the slit pore divided into slices of 2 Å thickness were calculated in the simulations to evaluate whether the coefficients can be used to determine the shear plane. The time interval is an important parameter in the calculation of the diffu­ sion coefficient. The influence of the time interval was discussed in Supporting Information. After comparison, the length of time intervals of 10 ps was chosen used in mean-square displacement calculation in MD simulations. The diffusion coefficients of water oxygen and Na+ parallel to the clay surface along z-direction are illustrated in Fig. 8, normalized by the values in the bulk 0.20 mol/L NaCl solution. Diffusion coefficients decrease linearly within 15 Å from the interface and reach a stable value beyond that, in agreement with the other MD results of montmorillonite nanopore (Bourg and Sposito, 2011). It can be noted that both coefficients of water and Na+ slightly decrease as ionic con­ centration increases. The Shear plane position is assumed located at the turning point with the diffusion value equals to the bulk value, which is 15 Å away from the clay surface. However, the calculated position is 0 10 20 30 40 50 0 3 6 9 12 15 u(z)/E x (m 2 /V·s) z (Å) 0.20 mol/L 0.37 mol/L 0.70 mol/L 1.30 mol/L ×10-8 2.7 Å Fig. 5. Electroosmotic velocity profiles normalized by the strength of the external electric field for montmorillonite in 1.30, 0.70, 0.37 and 0.20 mol/L NaCl electrolytes. The flow domains start at 2.7 Å away from the clay surface. Fig. 6. The electric potential distributions of montmorillonite with different ion concentrations. The position of the zeta position is obtained from the Stern model and zeta potential. Table 3 Calculated parameters in the triple-layer model for montmorillonite in NaCl solutions. The capacitance C1 in the Stern layer is 1.0 F m− 2 . 0.20 mol/ L 0.37 mol/ L 0.70 mol/ L 1.30 mol/ L Surface charge at 0-plane, Q0 (C/m2 ) − 0.1293 − 0.1297 − 0.1294 − 0.1288 Surface charge at d-plane, Qd (C/m2 ) − 0.0562 − 0.0613 − 0.0573 − 0.0614 Capacitance C2 (F⋅m− 2 ) 3.34 4.49 5.09 8.04 Position of β-plane, zβ (Å) 4.3 4.3 4.3 4.3 Position of d-plane, zd (Å) 6.0 5.5 5.3 4.8 Potential at 0-plane, φ0 (mV) − 193.0 − 181.4 − 167.5 − 156.6 Potential at β-plane, φβ (mV) − 63.7 − 51.7 − 38.1 − 27.8 Potential at d-plane, φd (mV) − 46.8 − 38.8 − 26.9 − 22.3 H. Pei and S. Zhang
  • 7. Applied Clay Science 212 (2021) 106212 7 beyond the Gouy plane in all concentrations of NaCl solutions, which contradicts the Stern model. Although the diffusion coefficients of water and cations have reduced mobility near the surface, a stagnant layer with zero diffusion coefficient cannot be observed. To sum up, the re­ sults confirmed that the shear plane position cannot be observed from the electroosmotic flow and diffusion coefficient in MD simulations. 3.4. Discussion on the TLM The TLM consists of several equations and parameters that make the solution difficult, which simplex algorithm is commonly used to solve the model (Leroy and Revil, 2004; Leroy et al., 2015). Parameters such as equilibrium constants at corresponding planes and capacitances of the inner and outer part of the Stern layer need to be estimated before the solution. Previous studies have evaluated these parameters based on experiments and MD simulations, especially for capacitances C1 and C2. A basic agreement was achieved for C1 based on the proposed values of 0.6–1.3 F m− 2 by experiments (Machesky et al., 1998; Sverjensky, 2001) and 0.8–1.2 F m− 2 by MD simulations (Tournassat et al., 2009). How­ ever, the value of C2 in the range of 0.2–5.5 F m− 2 (Yates et al., 1974; Hiemstra and Van Riemsdijk, 2006; Wolthers et al., 2008; Bourg and Sposito, 2011) is still in dispute. In this paper, the capacitance C2 between β-plane and d-plane was investigated. As the obtained position of β-plane and d-plane in the TLM (Table 2), the capacitance C2 can be estimated by the relation (Sahai and Sverjensky, 1997): Ci = εiε0/Δzi (17) where εi and Δzi are the relative permittivity and thickness of each “capacitor”. The spaces between β-plane and d-plane for montmoril­ lonite in 0.20–1.30 mol/L NaCl solutions are 1.7, 1.2, 1.0 and 0.6 Å. Wander and Clark (2008) have calculated the dielectric constants for the quartz-water interface as a function of the distance from the surface. In this paper, the permittivity of the outer Stern layer is calculated based on the proposed curve (permittivity scaled using reference value of 78 for bulk water), given as 64.2, 60.9, 57.5 and 54.5, respectively. Thus, the capacitance C2 in the TLM can be obtained by Eq. (17) as 3.34, 4.49, 5.09 and 8.04 F m− 2 , respectively. The results show that the value of capacitance between β-plane and d-plane has a large range depending on the ionic concentration, which is in disagreement with the constant value calculated by Bourg and Sposito (2011) but in agreement with the conclusion of Nishimura et al. (2002). After obtaining the capacitance between β-plane and d-plane, the TLM can be solved. Fig. S4 shows the potential distribution from the clay surface based on the TLM. The capacitance within the Stern layer is set as 1.0 F m− 2 . The potential 2 3 4 5 6 7 8 4 5 6 7 8 9 by Stern model by TLM model Zeta position from the surface (Å) -1(Å) 0.28 / 5.37 d 0.25 / 4.28 d 2 0.98 R 2 0.96 R 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.80.9 1 4 5 6 7 8 9 10 2 0.96 R 2 0.93 R log -0.112log 0.664 d I log -0.111log 0.751 d I By Stern model By triple-layer model logd (Å) logI (M) Fig. 7. The distance of the zeta potential position from the clay surface calculated by the Stern model and triple-layer model as a function of the Debye length and ionic strength. Fig. 8. Self-diffusion coefficients of water oxygen and Na+ parallel to the clay surface as a function of distance from the basal surface. Error bars show the values of 95% confidence interval. H. Pei and S. Zhang
  • 8. Applied Clay Science 212 (2021) 106212 8 calculated by the TLM is smaller than the Stern model, especially for the potential between β-plane and d-plane. The positions of zeta potential calculated by the TLM are also smaller than those by the Stern model, as mentioned above. 4. Limitations This paper only studies the EDL structure and electroosmosis prop­ erties of the basal surface of montmorillonite. It is well known, mont­ morillonite nanoparticle has two surfaces including basal and edge surface. Due to the large specific surface area, the basal surface is the domain part of the surface of montmorillonite, which indicates the re­ sults in this paper can generally reflect the properties of montmoril­ lonite. However, surface complexes on the edge site can also develop in the EDL. Cations exchange on edge surfaces shows complex, thermo­ dynamically non-ideal behavior much different from exchange on the basal surface (Lammers et al., 2017). Besides, the orientation (Kraevsky et al., 2020) and cleavage type (Shen and Bourg, 2021) of the edge in­ fluence the stability of the edge structure and its adsorption character­ istics. The investigations of surface complexation of edge surfaces can be found in Newton et al. (2016, 2017) and Kraevsky et al. (2020). Besides, the edge surface of montmorillonite also has an influence on the value of zeta potential. Edges surfaces have a pH-dependent charge, while basal surfaces bear a negative charge generated by isomorphous substitution (Delgado et al., 1985; Durán et al., 2000). This is a significant reason that the zeta potential of montmorillonite changes as a function of pH value. The effect of particle morphologies on the zeta potential was also not considered in this paper. Those reasons may cause the gap between the zeta potential measurement and the MD simulation result. This paper distinguished the zeta potential and the shear plane. The latter one was confirmed not to exist from MD simulations. The zeta potential position was measured and its properties were investigated in the paper. Thus, two questions arise. What is the physical meaning of the zeta potential if the shear plane does not exist? What is the physical meaning of the zeta potential position? This paper attempts to answer these questions by combing our thoughts with previous literature. Because those are still arguments in collochemistry, more in-depth work should be done in the future. Firstly, although its definition is related to the shear plane, the generation and physical meaning of zeta potential is independent of the existence of the shear plane. The shear plane is an imaginary plane only used to reconcile the Poisson-Boltzmann equation and the Navier-Stokes equation (Eq. (6)). Předota et al. (2016) have demonstrated that the zeta potential does not arise from the existence of a shear plane, but rather precisely from the electrokinetically driven motion of ions within the whole inhomogeneous interfacial region. Delgado et al. (2007) also said, the zeta potential is fully defined by the nature of the surface, its charge, the electrolyte concentration in the solution, and the nature of the electrolyte and of the solvent. Therefore, the zeta potential is still meaningful without the shear plane. Secondly, although the physical interpretation of the zeta potential is still ambiguous, a widely accepted meaning is proposed by Delgado et al. (2007). The zeta potential is the observed electrokinetic signal crossing from the non-contributing region for electrokinetic phenomenon to the double layer. Following this idea, the physical meaning of the zeta potential po­ sition is given as which is the boundary of electrokinetic phenomenon non-contributing region. The charges located between the surface and the zeta potential position are electrokinetically inactive exhibiting only electrostatic properties and contribute to the excess conductivity of the double layer (Delgado et al., 2007). This is the reason that the bulk streaming velocity is related to the zeta potential rather than the surface potential in the H–S equation. Although no electrokinetic effect is generated in this region under an external electrical field, water behind the zeta potential position has non-zero velocity in the electroosmotic tangential flow because of the viscosity. This also shows the difference between the shear plane and zeta potential position and proves the non- existence of the shear plane. 5. Conclusions A novel approach based on molecular dynamics was proposed in this paper to determine the zeta potential and the position of the shear plane for montmorillonite in NaCl electrolyte with a concentration of 0.20–1.30 mol/L. It combines an electrostatic surface complexation model (Stern model or TLM) that determines the electric potential dis­ tribution in the EDL from the ion density profiles in the EMD with the electroosmotic flow NEMD simulation that calculates the zeta potential. The conclusions can be summarized as follows: (1) The density profiles of ion species clearly show the existence of ion adsorption complexes. The Stern model has better accuracy in predicting the ions density profiles after the Stern potential determination, compare with the Gouy-Chapman and modified Gouy-Chapman model. (2) The zeta potential is calculated based on the electroosmotic ve­ locity profile, considering the slip length, close to the experi­ mental measurement. Methods including NEMD and self- diffusion coefficients cannot observe the shear plane. The posi­ tion of zeta potential was determined by the Stern model and TLM, showing a certain distance from the Stern plane. The dis­ tance of the zeta position from the clay surface has a strictly linear relationship with the Debye length and the ionic strength in the log scale, both in agreement with the experimental observation. (3) The capacitance between the Stern and shear plane in the TLM was calculated indicating that the capacitance is variable depending on the ionic strength rather than a constant value. Declaration of Competing Interest The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper. Acknowledgements This research has been supported by the China National Key R&D Program during the 13th Five-year Plan Period (Grant No. 2018YFC1505104 and 2017YFC1503103), National Natural Science Foundation of China (Grants No. 51778107) and Liao Ning Revitaliza­ tion Talents Program (Grants No. XLYC1807263). Appendix A. Supplementary data Supplementary data to this article can be found online at https://doi. org/10.1016/j.clay.2021.106212. References Balasubramanian, S., Mundy, C.J., Klein, M.L., 1996. Shear viscosity of polar fluids: Molecular dynamics calculations of water. J. Chem. Phys. 105, 11190–11195. https://doi.org/10.1063/1.472918. Bazant, M.Z., Kilic, M.S., Storey, B.D., Ajdari, A., 2009. Towards an understanding of induced-charge electrokinetics at large applied voltages in concentrated solutions. Adv. Colloid Interface 152, 48–88. Biriukov, D., Fibich, P., Předota, M., 2020. Zeta potential determination from molecular simulations. J. Phys. Chem. C 124, 3159–3170. https://doi.org/10.1021/acs. jpcc.9b11371. Bourg, I.C., Sposito, G., 2011. Molecular dynamics simulations of the electrical double layer on smectite surfaces contacting concentrated mixed electrolyte (NaCl-CaCl2) solutions. J. Colloid Interface Sci. 360, 701–715. https://doi.org/10.1016/j. jcis.2011.04.063. Bourg, I.C., Lee, S.S., Fenter, P., Tournassat, C., 2017. Stern layer structure and energetics at mica-water interfaces. J. Phys. Chem. C 121, 9402–9412. https://doi.org/ 10.1021/acs.jpcc.7b01828. Bourikas, K., Hiemstra, T., Van Riemsdijk, W.H., 2001. Ion pair formation and primary charging behavior of titanium oxide (anatase and rutile). Langmuir 17, 749–756. H. Pei and S. Zhang
  • 9. Applied Clay Science 212 (2021) 106212 9 Celebi, A.T., Beskok, A., 2018. Molecular and continuum transport perspectives on electroosmotic slip flows. J. Phys. Chem. C 122, 9699–9709. https://doi.org/ 10.1021/acs.jpcc.8b02519. Celebi, A.T., Cetin, B., Beskok, A., 2019. Molecular and continuum perspectives on intermediate and flow reversal regimes in electroosmotic transport. J. Phys. Chem. C. https://doi.org/10.1021/acs.jpcc.9b02432. Chang, F.R.C., Skipper, N.T., Sposito, G., 1998. Monte Carlo and molecular dynamics simulations of electrical double-layer structure in potassium-montmorillonite hydrates. Langmuir 14, 1201–1207. https://doi.org/10.1021/la9704720. Charlton, I.D., Doherty, A.P., 1999. Locating the micellar shear plane and its relationship with the Debye screening length. J. Phys. Chem. B 103, 5081–5083. https://doi.org/ 10.1021/jp9843914. Cheng, L., Fenter, P., Nagy, K.L., Schlegel, M.L., Sturchio, N.C., 2001. Molecular-scale density oscillations in water adjacent to a mica surface. Phys. Rev. Lett. 87 https:// doi.org/10.1103/PhysRevLett.87.156103, 156103-156103–4. Cygan, R.T., Liang, J.-J., Kalinichev, A.G., 2004. Molecular models of hydroxide, oxyhydroxide, and clay phases and the development of a general force field. J. Phys. Chem. B 108, 1255–1266. https://doi.org/10.1021/jp0363287. Delgado, A.V., Gonzalez-Cabllero, F., Bruque, J.M., 1985. On the zeta potential and surface charge density of montmorillonite in aqueous electrolyte solutions. J. Colloid Interface Sci. 13 (1), 203–211. Delgado, A.V., Gonzalez-Cabllero, F., Hunter, R.J., Koopal, L., Lyklema, J., 2007. Measurement and interpretation of electrokinetic phenomena. J. Colloid Interface Sci. 309, 194–224. Ding, W., Liu, X., Song, L., Li, Q., Zhu, Q., Zhu, H., Hu, F., Luo, Y., Zhu, L., Li, H., 2015. An approach to estimate the position of the shear plane for colloidal particles in an electrophoresis experiment. Surf. Sci. 632, 50–59. https://doi.org/10.1016/j. susc.2014.08.024. Dufreche, J.F., Marry, V., Malikova, N., Turq, P., 2005. Molecular hydrodynamics for electro-osmosis in clays: from Kubo to Smoluchowski. J. Mol. Liq. 118, 145–153. https://doi.org/10.1016/j.molliq.2004.07.076. Durán, J.D.G., Ramos-Tejada, M.M., Arroyo, F.J., Gonzalez-Cabllero, F., 2000. Rheological and electrokinetic properties of sodium montmorillonite suspensions I. Rheological properties and interparticle energy of interaction. J. Colloid Interface Sci. 229, 107–117. Evans, D.F., Wennerström, H., 1999. The Colloidal Domain: Where Physics, Chemistry, Biology, and Technology Meet, 2nd ed. Wiley-VCH, New York. Freund, J.B., 2002. Electro-osmosis in a nanometer-scale channel studied by atomistic simulation. J. Chem. Phys. 116, 2194–2200. https://doi.org/10.1063/1.1431543. Gereben, O., Pusztai, L., 2011. On the accurate calculation of the dielectric constant from molecular dynamics simulations: the case of SPC/E and SWM4-DP water. Chem. Phys. Lett. 507, 80–83. https://doi.org/10.1016/j.cplett.2011.02.064. Greathouse, J.A., Hart, D.B., Bowers, G.M., Kirkpatrick, R.J., Cygan, R.T., 2015. Molecular simulation of structure and diffusion at smectite-water interfaces: using expanded clay interlayers as model nanopores. J. Phys. Chem. C 119, 17126–17136. https://doi.org/10.1021/acs.jpcc.5b03314. Greenwood, R., 2003. Review of the measurement of zeta potentials in concentrated aqueous suspensions using electroacoustics. Adv. Colloid Interf. Sci. 106, 55–81. https://doi.org/10.1016/S0001-8686(03)00105-2. Hartkamp, R., Siboulet, B., Dufreche, J.F., Coasne, B., 2005. Ion-specific adsorption and electroosmosis in charged amorphous porous silica. Phys. Chem. Chem. Phys. 17 (2005), 24683–24695. Hartkamp, R., Biance, A.L., Fu, L., Dufreche, J.F., Bonhomme, O., Joly, L., 2018. Measuring surface charge: why experimental characterization and molecular modeling should be coupled. Curr. Opin. Colloid Interface Sci. 37, 101–114. Hiemstra, T., Van Riemsdijk, W.H., 2006. On the relationship between charge distribution, surface hydration, and the structure of the interface of metal hydroxides. J. Colloid Interface Sci. 301, 1–18. https://doi.org/10.1016/j. jcis.2006.05.008. Hocine, S., Hartkamp, R., Siboulet, B., Duvail, M., Coasne, B., Turq, P., Dufrêche, J.F., 2016. How ion condensation occurs at a charged surface: a molecular dynamics investigation of the Stern layer for water-silica interfaces. J. Phys. Chem. C 120, 963–973. https://doi.org/10.1021/acs.jpcc.5b08836. Hockney, R., Eastwood, J., 1988. Computer Simulation Using Particles. CRC Press. Hou, J., Li, H., Zhu, H., Wu, L., 2009. Determination of clay surface potential: a more reliable approach. Soil Sci. Soc. Am. J. 73, 1658–1663. https://doi.org/10.2136/ sssaj2008.0017. Hunter, R.J., 1981. Zeta Potential in Colloidal Science: Principles and Applications. Academic Press, New York. Jain, K., Mehandzhiyski, A.Y., Zozoulenko, I., Wågberg, L., 2021. PEDOT:PSS nano- particles in aqueous media: a comparative experimental and molecular dynamics study of particle size, morphology and z-potential. J. Colloid Interface Sci. 584, 57–66. https://doi.org/10.1016/j.jcis.2020.09.070. Kraevsky, S.V., Tournassat, C., Vayer, M., Warmont, F., et al., 2020. Indentification of montmorillonite particle edge orientations by atomic-force microscopy. Appl. Clay Sci. 186, 105442. Lammers, L.N., Bourg, I.C., Okumura, M., Kolluri, K., et al., 2017. Molecular dynamics simulations of cesium adsorption on illite nanoparticles. J. Colloid Interface Sci. 490, 608–620. Le Crom, S., Tournassat, C., Robinet, J.C., Marry, V., 2020. Influence of polarizability on the prediction of the electrical double layer structure in a clay mesopore: a molecular dynamics study. J. Phys. Chem. C 124, 6221–6232. Leroy, P., Revil, A., 2004. A triple-layer model of the surface electrochemical properties of clay minerals. J. Colloid Interface Sci. 270, 371–380. https://doi.org/10.1016/j. jcis.2003.08.007. Leroy, P., Revil, A., Kemna, A., Cosenza, P., Ghorbani, A., 2008. Complex conductivity of water-saturated packs of glass beads. J. Colloid Interface Sci. 321, 103–117. Leroy, P., Tournassat, C., Bernard, O., Devau, N., Azaroual, M., 2015. The electrophoretic mobility of montmorillonite. Zeta potential and surface conductivity effects. J. Colloid Interface Sci. 451, 21–39. https://doi.org/10.1016/j.jcis.2015.03.047. Li, G.X., 2016. Advanced Soil Mechanics, second edition. Tsinghua University Press (in Chinese). Li, H., Wei, S., Qing, C., Yang, J., 2003. Discussion on the position of the shear plane. J. Colloid Interface Sci. 258, 40–44. https://doi.org/10.1016/S0021-9797(02) 00077-2. Li, H., Qing, C.L., Wei, S.Q., Jiang, X.J., 2004. An approach to the method for determination of surface potential on solid/liquid interface: theory. J. Colloid Interface Sci. 275, 172–176. Liu, X., Hu, F., Ding, W., Tian, R., Li, R., Li, H., 2015. A how-to approach for estimation of surface/Stern potentials considering ionic size and polarization. Analyst 140, 7217–7224. https://doi.org/10.1039/c5an01053e. Liu, X., Ding, W., Tian, R., Li, R., Li, H., 2016. How ionic polarization affects Stern potential: an insight into hofmeister effects. Soil Sci. Soc. Am. J. 80, 1181–1189. https://doi.org/10.2136/sssaj2016.04.0095. Liu, X., Ding, W., Tian, R., Du, W., Li, H., 2017. Position of shear plane at the clay-water interface: strong polarization effects of counterions. Soil Sci. Soc. Am. J. 81, 268–276. https://doi.org/10.2136/sssaj2016.08.0261. Liu, X.T., Yang, S., Gu, P.K., Liu, S., Yang, G., 2021. Adsorption and Removal of Metal Ions by Smectites Nanoparticles: Mechanistic Aspects, and Impacts of Charge Location and Edge Structure. Lorenz, C.D., Travesset, A., 2007. Charge inversion of divalent ionic solutions in silica channels. Phys. Rev. E 75, 061202. Lorenz, C.D., Crozier, P.S., Anderson, J.A., Travesset, A., 2008. Molecular dyncamics of ionic transport and electrokinetic effects in realistic silica channels. J. Chem. Phys. C 112, 10222–10232. Lyklema, J. 1995. Fundamentals of Interface and Colloid Science, Academic Press, 2: 1–232. Lylema, J., Rovillard, S., de Coninck, J., 1998. Electrokinetics: the properties of the stagnant layer unraveled. Langmuir 14, 5659–5663. Machesky, M.L., Wesolowski, D.J., Palmer, D.A., Ichiro-Hayashi, K., 1998. Potentiometric titrations of rutile suspensions to 250◦ C. J. Colloid Interface Sci. 200, 298–309. https://doi.org/10.1006/jcis.1997.5401. Mészáros, R., Jobbik, A., Varga, G., Bárány, S., 2019. Electrosurface properties of Na- bentonite particles in electrolytes and surfactants solution. Appl. Clay Sci. 178, 105127. https://doi.org/10.1016/j.clay.2019.105127. Miller, S.E., Low, P.F., 1990. Characterization of the electrical double layer of montmorillonite. Langmuir 6, 572–578. https://doi.org/10.1021/la00093a010. Moussa, C., Xu, J., Wang, X., Zhang, J., Chen, Z., Li, X., 2017. Molecular dynamics simulation of hydrated Na-montmorillonite with inorganic salts addition at high temperature and high pressure. Appl. Clay Sci. 146, 206–215. https://doi.org/ 10.1016/j.clay.2017.05.045. Newton, A.G., Kwon, K.D., Cheong, D.K., 2016. Edge structure of montmorillonite from atomistic simulations. Minerals 6 (25). Newton, A.G., Kwon, K.D., Cheong, D.K., 2017. Na-montmorillonite edge structure and surface complexes: an atomistic perspective. Minerals 7 (78). Nishimura, S., Yao, K., Kodama, M., Imai, Y., et al., 2002. Electrokinetc study of synthetic smectites by flat plate streaming potential technique. Langmuir. 18, 188–193. Ou, C.Y., Chien, S.C., Yang, C.C., Chen, C.T., 2015. Mechanism of soil cementation by electroosmotic chemical treatment. Appl. Clay Sci. 104, 135–142. https://doi.org/ 10.1016/j.clay.2014.11.020. Panagiotou, G.D., Petsi, T., Bourikas, K., Garoufalis, C.S., et al., 2008. Mapping the surface (hydr)oxo-groups of titanium oxide and its interface with an aqueous solution: the state of the art and a new approach. Adv. Colloid Interface 142, 20–42. Park, C., Fenter, P.A., Nagy, K.L., Sturchio, N.C., 2006. Hydration and distribution of ions at the mica-water interface. Phys. Rev. Lett. 97, 1–4. https://doi.org/10.1103/ PhysRevLett.97.016101. Park, S.H., Sposito, G., 2002. Structure of Water Adsorbed on a Mica Surface. Phys. Rev. Lett. 89, 8–10. https://doi.org/10.1103/PhysRevLett.89.085501. Parsons, D.F., Ninham, B.W., 2010. Charge reversal of surfaces in divalent electrolytes: the role of ionic dispersion interactions. Langmuir 26, 6430–6436. https://doi.org/ 10.1021/la9041265. Peng, J., Ye, H., Alshawabkeh, A.N., 2015. Soil improvement by electroosmotic grouting of saline solutions with vacuum drainage at the cathode. Appl. Clay Sci. 114, 53–60. https://doi.org/10.1016/j.clay.2015.05.012. Plimpton, S., 1995. Fast parallel algorithms for short-range molecular dynamics. J. Comput. Phys. https://doi.org/10.1006/jcph.1995.1039. Předota, M., Machesky, M.L., Wesolowski, D.J., 2016. Molecular origins of the zeta potential. Langmuir 32, 10189–10198. https://doi.org/10.1021/acs. langmuir.6b02493. Rezaei, M., Azimian, A.R., Pishevar, A.R., 2018. Surface charge-dependent hydrodynamic properties of an electroosmotic slip flow. Phys. Chem. Chem. Phys. 20, 30365–30375. https://doi.org/10.1039/c8cp06408c. Ricci, M., Spijker, P., Stellacci, F., Molinari, J.F., Voïtchovsky, K., 2013. Direct visualization of single ions in the Stern layer of calcite. Langmuir 29, 2207–2216. https://doi.org/10.1021/la3044736. Sahai, N., Sverjensky, D.A., 1997. Solvation and electrostatic model for specific electrolyte adsorption. Geochim. Cosmochim. Acta 61, 2827–2848. https://doi.org/ 10.1016/S0016-7037(97)00127-0. Şans, B.E., Güven, O., Esenli, F., Çelik, M.S., 2017. Contribution of cations and layer charges in the smectite structure on zeta potential of Ca-bentonites. Appl. Clay Sci. 143, 415–421. https://doi.org/10.1016/j.clay.2017.04.016. H. Pei and S. Zhang
  • 10. Applied Clay Science 212 (2021) 106212 10 Schmickler, W., Henderson, D., 1986. New models for the structure of the electrochemical interface. Prog. Surf. Sci. 22, 323–420. Shang, J.Q., 1997. Zeta potential and electroosmotic permeability of clay soils. Can. Geotech. J. 34, 627–631. https://doi.org/10.1139/t97-28. Shang, J.Q., Lo, K.Y., Quigley, R.M., 1994. Quantitative determination of potential distribution in Stern-Gouy double-layer model. Can. Geotech. J. 31, 624–636. https://doi.org/10.1139/t94-075. Shen, X.Y., Bourg, I.C., 2021. Molecular dynamics simulations of the colloidal interaction between smectite clay nanoparticles in liquid water. J. Colloid Interface Sci. 584, 610–621. Siboulet, B., Hocine, S., Hartkamp, R., Dufreche, J.F., 2017. Scrutinizing electro-osmosis and surface conductivity with molecular dynamics. J. Phys. Chem. C 121 (12), 6756–6769. Sondi, I., Bišćan, J., Pravdić, V., 1996. Electrokinetics of pure clay minerals revisited. J. Colloid Interface Sci. 178, 514–522. https://doi.org/10.1006/jcis.1996.0146. Sverjensky, D.A., 2001. Interpretation and prediction of triple-layer model capacitances and the structure of the oxide-electrolyte-water interface. Geochim. Cosmochim. Acta 65, 3643–3655. https://doi.org/10.1016/S0016-7037(01)00709-8. Sverjensky, D.A., 2005. Prediction of surface charge on oxides in salt solutions: Revisions for 1:1 (M+L-) electrolytes. Geochim. Cosmochim. Acta 69, 225–257. https://doi. org/10.1016/j.gca.2004.05.040. Tombácz, E., Szekeres, M., 2004. Colloidal behavior of aqueous montmorillonite suspensions: the specific role of pH in the presence of indifferent electrolytes. Appl. Clay Sci. 27, 75–94. https://doi.org/10.1016/j.clay.2004.01.001. Tournassat, C., Chapron, Y., Leroy, P., Bizi, M., Boulahya, F., 2009. Comparison of molecular dynamics simulations with triple layer and modified Gouy-Chapman models in a 0.1 M NaCl-montmorillonite system. J. Colloid Interface Sci. 339, 533–541. https://doi.org/10.1016/j.jcis.2009.06.051. Vasu, N., De, S., 2010. Electroosmotic flow of power-law fluids at high zeta potentials. Colloids Surf. A Physicochem. Eng. Asp. 368, 44–52. https://doi.org/10.1016/j. colsurfa.2010.07.014. Wander, M.C.F., Clark, A.E., 2008. Structural and Dielectric Properties of Quartz – Water Interfaces 19986–19994. https://doi.org/10.1021/jp803642c. Wang, J., Wang, M., Li, Z., 2006. Lattice Poisson-Boltzmann simulations of electro- osmotic flows in microchannels. J. Colloid Interface Sci. 296, 729–736. https://doi. org/10.1016/j.jcis.2005.09.042. Wolthers, M., Charlet, L., Cappellen, P.V., 2008. The surface chemistry of divalent metal carbonate minerals; a critical assessment of surface charge and poential data using the charge distribution multi-site ion complexation model. Am. J. Sci. 308, 905–941. Xie, Y., Fu, L., Niehaus, T., Joly, L., 2020. Liquid-solid slip on charged walls: the dramatic impact of charge distribution. Phys. Rev. Lett. 125, 1–7. https://doi.org/10.1103/ PhysRevLett.125.014501. Xu, R., 1998. Shear plane and hydrodynamic diameter of microspheres in suspension. Langmuir 14, 2593–2597. https://doi.org/10.1021/la971404g. Yates, D.E., Levine, S., Healy, T.W., 1974. Site-binding model of the electrical double layer at the oxide/water interface. J. Chem. Soc. Farady Trans. 70, 1807. Zhang, C., Liu, Z., Dong, Y., 2017. Effects of adsorptive water on the rupture of nanoscale liquid bridges. Appl. Clay Sci. 146, 487–494. https://doi.org/10.1016/j. clay.2017.07.002. Zhang, H., Hassanali, A.A., Shin, Y.K., Knight, C., Singer, S.J., 2011. The water- amorphous silica interface: analysis of the Stern layer and surface conduction. J. Chem. Phys. 134 https://doi.org/10.1063/1.3510536. Zhang, S., Pei, H., 2020. Rate of capillary rise in quartz nanochannels considering the dynamic contact angle by using molecular dynamics. Powder Technol. 372, 477–485. https://doi.org/10.1016/j.powtec.2020.06.018. Zhao, Q., Choo, H., Bhatt, A., Burns, S.E., Bate, B., 2017. Review of the fundamental geochemical and physical behaviors of organoclays in barrier applications. Appl. Clay Sci. 142, 2–20. https://doi.org/10.1016/j.clay.2016.11.024. H. Pei and S. Zhang