ArticlePDF Available

Atomistic simulations of cation hydration in sodium and calcium montmorillonite nanopores

Authors:

Abstract and Figures

In this work, classical MD simulations using Gromacs with the CLAYFF force field plus the SPC/E water model and Joung-Cheatham ion model were performed by accounting for the explicit solvent molecules. MD simulations show that Na+ in Na-MMT nanopores have an affinity to the ditrigonal cavities of the clay layers and form transient inner-sphere complexes at about 3.8 Å from clay midplane at water contents less than the fifth-layer hydration state. However, these phenomena are not observed in Ca-MMT regardless of swelling states. Moreover, the MD results showed that the complete hydration shells are found to be nearly spherical (with an orthogonal coordination sphere). They could only be formed when the basal spacing d001 ≥ 18.7 Å, i.e., the interlayer separation h ≥ 10 Å approximately. Comparison between DFT and MD simulations shows that DFT failed to reproduce the outer-sphere complexes in the Stern-layer (within ~5.0 Å from the clay basal-plane), observed in the MD simulations.
Content may be subject to copyright.
Atomistic simulations of cation hydration in sodium and calcium montmorillonite
nanopores
Guomin Yang, Ivars Neretnieks, and Michael Holmboe
Citation: The Journal of Chemical Physics 147, 084705 (2017); doi: 10.1063/1.4992001
View online: http://dx.doi.org/10.1063/1.4992001
View Table of Contents: http://aip.scitation.org/toc/jcp/147/8
Published by the American Institute of Physics
THE JOURNAL OF CHEMICAL PHYSICS 147, 084705 (2017)
Atomistic simulations of cation hydration in sodium and calcium
montmorillonite nanopores
Guomin Yang,1,a) Ivars Neretnieks,1and Michael Holmboe2
1Department of Chemical Engineering, Royal Institute of Technology, S-100 44 Stockholm, Sweden
2Department of Chemistry, Ume˚
a University, Ume˚
a, Sweden
(Received 25 June 2017; accepted 1 August 2017; published online 23 August 2017)
During the last four decades, numerous studies have been directed to the swelling smectite-rich clays
in the context of high-level radioactive waste applications and waste-liners for contaminated sites.
The swelling properties of clay mineral particles arise due to hydration of the interlayer cations
and the diffuse double layers formed near the negatively charged montmorillonite (MMT) surfaces.
To accurately study the cation hydration in the interlayer nanopores of MMT, solvent-solute and
solvent-clay surface interactions (i.e., the solvation effects and the shape effects) on the atomic level
should be taken into account, in contrast to many recent electric double layer based methodologies
using continuum models. Therefore, in this research we employed fully atomistic simulations using
classical molecular dynamics (MD) simulations, the software package GROMACS along with the
CLAYFF forcefield and the SPC/E water model. We present the ion distributions and the deformation
of the hydrated coordination structures, i.e., the hydration shells of Na+and Ca2+ in the interlayer,
respectively, for MMT in the first-layer, the second-layer, the third-layer, the fourth-layer, and the fifth-
layer (1W, 2W, 3W, 4W, and 5W) hydrate states. Our MD simulations show that Na+in Na-MMT
nanopores have an affinity to the ditrigonal cavities of the clay layers and form transient inner-sphere
complexes at about 3.8 Å from clay midplane at water contents less than the 5W hydration state.
However, these phenomena are not observed in Ca-MMT regardless of swelling states. For Na-MMT,
each Na+is coordinated to four water molecules and one oxygen atom of the clay basal-plane in
the first hydration shell at the 1W hydration state, and with five to six water molecules in the first
hydration shell within a radius of 3.1 Å at all higher water contents. In Ca-MMT, however each Ca2+
is coordinated to approximately seven water molecules in the first hydration shell at the 1W hydration
state and about eight water molecules in the first hydration shell within a radius of 3.3 Å at all higher
hydration states. Moreover, theMD results show that the complete hydration shells are nearly spherical
with an orthogonal coordination sphere. They could only be formed when the basal spacing d001
18.7 Å, i.e., approximately, the interlayer separation h10 Å. Comparison between DFT and MD
simulations shows that DFT failed to reproduce the outer-sphere complexes in the Stern-layer (within
5.0 Å from the clay basal-plane), observed in the MD simulations. Published by AIP Publishing.
[http://dx.doi.org/10.1063/1.4992001]
I. INTRODUCTION
Hydration of the interlayer ions and their ion distribu-
tions near the clay-water interface govern the underlying
mechanisms of clay swelling,18clay mineral precipitations,
as well as radionuclide migration and retardation in argilla-
ceous rocks. In these systems, the electric double layers
(EDL) formed near the charged surfaces, e.g., the negatively
charged clays make the clay swell or collapse.911 The sig-
nificance of EDL for biological systems and in industrial
technical processes has been well illustrated in the work of
Angell et al.12 Recently, the EDL of ionic liquids has been
cumulatively relevant in designing lithium ion batteries and
super-capacitors.13
In the theoretical study of the EDL of electrolyte solutions
and ionic liquids, the simplest model of the underlying EDL
a)Electronic mail: guomin@kth.se
theories is the so-called primitive model (PM), where ions
typically are modeled as hard spheres and where the water
molecules are only represented implicitly as a continuum,
using a uniform dielectric constant. In the PM, the hydration
effects simply are accounted for by increasing the radius of
the ions. The enlarged ionic radius thus corresponds to the
“hydrated ionic radius” representing the size of the ion plus
the hydration shell of tightly coordinated and oriented water
molecules around it. With decreasing interlayer distances (i.e.,
decreasing water content) the hydration shell may deform,
eventually resulting in water molecules being squeezed out
from the inner hydration shell. Under the condition when the
interlayer distances are smaller than 1 nm, equivalent to a basal
spacing of approximately 2 nm, using a “hydrated ionic radius”
for the ions and particles, therefore, might become incredibly
unphysical.10
In an effort to account for the molecular nature of the
solvent, the PM was extended to a three-component model
0021-9606/2017/147(8)/084705/8/$30.00 147, 084705-1 Published by AIP Publishing.
084705-2 Yang, Neretnieks, and Holmboe J. Chem. Phys. 147, 084705 (2017)
(3CM) in which the solvent molecules are represented by
neutral hard spheres, and the electrostatic part of the sol-
vent molecules is still described as a continuum with uniform
dielectric constant.14 The ion-ion interaction in the 3CM thus
remains exactly the same as in the PM, whereas the ion-
solvent and solvent-solvent interactions are simply represented
by the hard sphere repulsions. The structural and thermo-
dynamic properties of the simplest PM and 3CM have been
studied.10,11,14,15
Yet another useful model of solvated electrolytes or
ionic liquids that goes beyond the primitive model is the
dimer electrolyte model,1624 in which the solvated cations
or ionic liquids are represented by a dimer consisting of two
tangentially tethered hard spheres one of which is neutral
and the other is positively charged. The ions are described
by a negatively charged hard sphere monomers in a sol-
vent that is a dielectric continuum. This dimer model has
successfully been implemented for canonical Monte Carlo
(CMC),25 and density functional theory of fluids (DFT),26
techniques for the prediction of structure and differential
capacitance of a planar EDL. Results from CMC simula-
tions have shown that the shape of cations influences the
double layer structure and capacitance substantially for a neg-
atively charged surface, but less so for a positively charged
wall.25
In all of the above models, the dielectric constant of water
is constant. In reality it decreases with increasing ion concen-
tration27,28 since the screening ability of the solvent (dielectric
permittivity) decreases as the concentration of ions increases.
Later, the so-called II+IW theory was proposed by taking into
account the contributions of ion-ion (II) and ion-water (IW)
interactions to the excess chemical potentials. In the II+IW
theory, the IW interaction contains the information of dielec-
tric constant about the ability of water to screen the ions.2932 In
their work, the II interaction was obtained with Grand Canon-
ical Monte Carlo (GCMC) simulations on the basis of the PM
by using the Pauling radii of the charged hard spheres. The IW
interaction was calculated from Born’s treatment of hydra-
tion by employing the Born radii, which are not a physical
radii, rather an effective parameter with which the theoretic
results agreed with the experimental hydration free energy.33
This II+IW model has been increasingly applied to predict the
activity coefficients of ions in bulk ionic liquid for a variety
of systems of interest. To the best of our knowledge, however
the work of the II+IW model for the EDL of ionic liquids and
electrolytes (in the presence of external field) are still rarely
investigated.
For more than two decades, molecular modeling such as
Monte Carlo (MC) and molecular dynamics (MD) simula-
tions1,2,46,8,3436 have proven to be valuable heuristic tech-
niques for studying the interlayer molecular structures and
the EDL properties of clay minerals, by explicitly considering
the water-water, clay-water, counter-ion-water, and counter-
ion-clay interactions on the atomic level. Nevertheless, MD
and MC simulations have limitations with regard to mod-
eling large system sizes and long time scales. Especially
MD simulations have been increasingly used for studying
the EDL in nanopores3,3739 because of its ability to give
explicit and atomistic insight into water-mineral interface
phenomena, with software packages such as GROMACS4044
and LAMMPS.45
With this in mind, this study was motivated by a pre-
vious work that used DFT and showed that the attractive
forces between the sheets under certain circumstances can be
switched to repulsive forces with increasing ion sizes.10 More-
over, the hydrated ionic radii used in DFT agree quite well with
experimental results in predicting the ion exchange equilibria
only when the interlayer separations are larger than about 10 Å.
It has been realized that the hard-sphere model (in which the
ionic diameters are adjustable parameters) used in DFT9,11
failed in describing the hydration shells of water molecules,
whereas this can be achieved by MD simulations. To study how
the deformation of hydration shells influences the properties of
compacted smectite clays, we employed classical MD simula-
tions taking into account all atom-atom interactions including
the water molecule interactions. Furthermore, the MD simu-
lations demonstrated under which conditions DFT modeling
becomes significantly inaccurate, and when it still can give
accurate results.
To summarize, in this research we report MD simula-
tion of the cation hydration in the interlayer nanopores of
Na and Ca-MMT at different water contents. In Sec. II, the
methodologies of MD simulations implemented with GRO-
MACS are described and illustrated. The interlayer structure
of cations and water molecules in Na- and Ca-MMT are dis-
cussed in Sec. III. The deformation of the hydration shells
is captured by MD when the swelling state is varied from
the fifth-layer hydrate (5W) to the first-layer hydrate (1W).
The structures of interlayer hydration and water molecules are
quantitatively described by the radial distribution functions
(RDF). In Sec. IV, the EDL structures computed by the MD
simulations are compared with the corresponding results from
the DFT calculations, to demonstrate under which conditions
DFT predictions become inaccurate.
II. MOLECULAR DYNAMIC SIMULATION METHOD
In this study, all-atom MD simulations were carried out
with the open source package GROMACS.41,44 All interatomic
interactions for and between water, the clay lattice, and the
interlayer ions, were modeled using the most popular SPC/E
water model,46 the well established CLAYFF forcefield47 and
the Åqvist48 and Joung-Cheatham parameters,49 respectively.
It has been verified that the combination of forcefields and
water models50,51 is superior in reproducing the overall density
profiles (by deriving them from the X-ray structure factors for
the hydrated clays plus adsorbed waters and ions and compar-
ing with experimental XRD data) for the crystalline hydrates
of smectites.
The CLAYFF forcefield uses partially charged atom types
and considers all interactions as non-bonded “ionic” inter-
actions, except for the hydroxyl O–H bond. The generated
charge defects (due to the isomorphic substitution of Al iii
with Mg ii) are distributed over the Mg atoms and all first-
neighbor oxygens, which carry charges that approximate the
water-oxygen in the SPC/E model, with which they also share
their Lennard-Jones parameters. Beyond a cutoff value of
1.0 nm, all interactions due to the Lennard-Jones potential
084705-3 Yang, Neretnieks, and Holmboe J. Chem. Phys. 147, 084705 (2017)
(VdW interactions) were truncated whereas the electrostatic
interactions were computed with the Particle-Mesh-Ewald
(PME) method.52
In this study we used a Wyoming-type MMT,5,53 having
a generic unit cell chemical formula of X+
0.66 Al3.33Mg0.66
[Si8]O20[OH]4, with a surface charge density of approxi-
mately 0.11 C m 2, which was compensated by the incor-
poration of cations into the interlayer region, e.g., 0.33 Ca2+
or 0.66 Na+per unit cell. X +
0.66 denotes the average 0.66 mono-
valent ions (corresponding to 0.33 divalent ion) needed to
counterbalance the negative lattice charge. In all the MD simu-
lations, the systems are modeled with two fully flexible parallel
MMT sheets (6 ×6×2 unit cells) stacked in the zdirection
(Fig. 1).
All MD runs were simulated at 298 K using a time step of
1 fs, and the integration of the equations of motion were per-
formed using the leap-frog algorithm. To generate equilibrated
systems, the initial configurations were prepared by (1) energy
minimization, using the steepest descent algorithm, followed
by (2) simulations using position restraints of the clay lattices
in the canonical NVT ensemble for 5 ns, and (3) simulations
for 5 ns in the isothermal-isobaric NPT ensemble using the
Berendsen barostat,54 at 1 bar and 298 K, to relax the MMT
layers and interlayer solutes in the x,y, and zdirection. The
final production runs were performed in the NVT ensemble at
298 K for 125 ns.
Interlayer density profiles for all atoms were calculated
from MD trajectory data using
ρi(z)=hNii/(xyz), (1)
where ρiis the number density of the atoms in region i, i.e.,
a sliced section located between zand z+z, and where hNii
is the ensemble averaged number of atoms within this region,
having a bin volume of xyzin Å3. Similarly, the radial dis-
tribution function is calculated from the trajectories using the
equation
gij(r)=DNjE . (ρj4πr2r), (2)
where DNjEis the ensemble averaged number of atoms jlocated
within the spherical shells at rand r+rcentered on the atom
i, and ρjis the average density of atom j. The radial distribution
FIG. 1. Snapshot of a MD simulation cell showing Ca-MMT in the 3W hydra-
tion state with an 11.7 Å wide interlayer nanopore containing Ca ions (green)
ions and water molecules. The clay mineral structure contains Si (yellow), Al
(pink), Mg (blue), O (red), and H (white) atoms.
functions can be integrated cumulatively and radially along the
radius r, giving the coordination number, CNij(r), of j-atoms
about the central i-atom,
CNij (r)=4π ρjXgij (r)r2r. (3)
III. MOLECULAR DYNAMICS SIMULATIONS
OF CATION HYDRATION
In the simulations studying hydration of interlayer cations,
the average interlayer distances were solely determined by the
number of water molecules in each interlayer, since there could
be no exchange of the water molecules or ions between the dif-
ferent adjacent layers. The water contents were varied from the
first-layer hydrate to the fifth-layer hydrate (i.e., 1W, 2W, 3W,
4W, and 5W), having 5, 10, 15, 20, and 25 H2O per unit cell of
MMT as recommended by Holmboe and Bourg38 for a typical
Wyoming type MMT, since these water loadings was shown
to reproduce experimental interlayer distances measured by
XRD.38,55
A. Interlayer structures
Figure 2presents the MD simulations of the interlayer ion
distributions for water hydrogen (HW), water oxygen (OW),
and cations along the z-axis in Na- and Ca-MMT,respectively.
In Fig. 2(a), the smectite basal spacing (d001) is shown to be
about 12.5 Å for Na-MMT in the 1W hydration state. It is seen
that the interlayer structures of the water molecules and Na ions
are symmetric along the midplane of the interlayer. In Fig. 2(a),
two shoulders of the HW density profiles located around 5 Å
from the clay midplane indicate that HW are attracted to the
ditrigonal cavities of the siloxane surfaces, likely interacting
through hydrogen bond with basal-plane oxygen atoms (at
about 3.2 Å from the clay midplane). OW and Na ions are
found located roughly equally around two planes with a small
offset from the interlayer midplane. The Na ions appear to be
forming the inner-sphere complexes (peaks about 5.7 Å from
the clay midplane) interacting with the octahedral charge site.
Nevertheless, two minor features in the Na ion distributions
are found to reside at around 3.8 Å from the clay midplane
because they are attracted to ditrigonal cavities and forming
transient inner-sphere surface complexes. Figure 2(b) shows a
basal spacing of 13.7 Å (slightly larger than the experimental
FIG. 2. Interlayer ion distributions as a function of interlayer separations for
(a) Na-MMT and (b) Ca-MMT in the 1W hydration state. Inserts show the Na
and Ca ion distributions magnified.
084705-4 Yang, Neretnieks, and Holmboe J. Chem. Phys. 147, 084705 (2017)
value) for Ca-MMT in 1W. It is seen that two small OW
peaks reside at the same place as the HW peaks in the wings
of distributions around 5 Å from the clay midplane. It indi-
cates that OW atoms located near the ditrigonal cavities of the
siloxane surfaces interacting with the clay hydrogen atoms.
Comparison between Na-MMT and Ca-MMT illustrate that
the Ca ions do not penetrate the ditrigonal cavities and show
a preference for outer-sphere complexes over inner-sphere
complexes.
Figure 3shows the ion distributions in the interlayer of
MMT in the 2W hydration state, resulting in a basal spac-
ing for both Na- and Ca-MMT of 15.6 Å, which is well in
line with the experimental basal spacing data in the work of
Holmboe et al.55 It is seen that the interlayer structures of HW
and OW in Na-MMT and Ca-MMT [Figs. 3(a) and 3(b)] are
very similar, and that the HW and OW distributions are sym-
metrically distributed around two planes with a small offset
from the interlayer midplane. For neither Na- nor Ca-MMT,
the OW atoms show no preference for the ditrigonal cavities
of the siloxane surfaces.
Figure 3(a) shows that the Na ions are mainly located
at the interlayer midplane, forming outer-sphere complexes
(peaks about 7.8 Å from the clay midplane) but also inner-
sphere complexes (peaks about 5.9 Å from the clay midplane).
Interestingly, even for the 2W hydration state, the Na ions
can form transient inner-sphere complexes with the ditrigonal
cavities (peaks about 3.8 Å from clay midplane) as shown
in the inserted plot [Fig. 3(a)]. The Ca ions, however, show
exclusively outer-sphere complexes within the interlayer water
molecules [Fig. 3(b)].
Figure 4shows the corresponding 3W hydration state
resulting in a basal spacing of approximately 18.7 Å for both
Na- and Ca-MMT (well in line with the experimental basal
spacing data, see Holmboe et al., 2012). In Fig. 4(a), the broad
and seemingly diffuse layer of Na ions was found, associated
with water molecules at the interlayer midplane, with two lay-
ers of outer-sphere complexes residing at about 7.6 Å from
the clay midplane, but still with two small shoulders located
around 5.9 Å indicating some residual inner-sphere complexes.
The transient inner-sphere complexes (peaks about 3.8 Å from
clay midplane shown in the interested plot) interacting with
the ditrigonal cavities seen at lower hydration states were also
FIG. 3. Interlayerion distributions as a function of interlayer separations from
MD simulations for (a) Na-MMT and (b) Ca-MMT in the 2W hydration state.
Inserts show the Na and Ca ion distributions magnified.
FIG. 4. Interlayer density profiles as a function of interlayer separations from
MD simulations for (a) Na-MMT and (b) Ca-MMT in the 3W hydration state.
Inserts show the Na and Ca ion distributions magnified.
found for Na-MMT in the 3W hydration state [Fig. 4(a)]. The
Na ion distributions display asymmetry in the transient inner-
sphere complexes as shown in the embedded plots, which
can be explained by the isomorphic substitution sites being
distributed randomly in the x-yplane, possibly creating a cer-
tain amount of heterogeneity in the surface charge density.
Figure 4(b) shows that there are only two shoulders in the Ca
ion distributions, located about 7.8 Å from the clay midplane,
without any significant diffuse or broad layer in the interlayer
midplane.
For the 4W hydration state, the basal spacing was found
to be approximately 21.7 Å for both Na- and Ca-MMT, shown
in Fig. 5. It is seen that both outer-sphere complexes and
diffuse-layer exist for both Na- and Ca-MMT in the 4W hydra-
tion state. Again, Fig. 5(a) shows the transient inner-sphere
complexes in ditrigonal cavities reside at 3.8 Å and residue
inner-sphere complexes at 5.9 Å from the clay midplane for
Na-MMT. Figure 5(b) illustrates that the diffuse layers appear
in Ca-MMT when the swelling reaches the 4W hydration
state.
For the 5W hydration state, the basal spacing was found
to be approximately 24.8 Å for both Na- and Ca-MMT.
Figure 6(a) shows a significant fraction of the Na ions reside
in outer-sphere complexes (peaks about 7.8 Å from the clay
midplane). The small peaks of Na ions at about 10.8 Å from
the clay midplane [Fig. 6(a)] indicate a complex exchange
FIG. 5. Interlayerion distributions as a function of interlayer separations from
MD simulations for (a) Na-MMT and (b) Ca-MMT in the 4W hydration state.
Inserts show the Na and Ca ion distributions magnified.
084705-5 Yang, Neretnieks, and Holmboe J. Chem. Phys. 147, 084705 (2017)
FIG. 6. Interlayer density profiles as a function of interlayer separations from
MD simulations for (a) Na-MMT and (b) Ca-MMT in the 5W hydration state.
Inserts show the Na and Ca ion distributions magnified.
between the diffuse-layer and outer-sphere complexes. It is
worth noting that the Na ions show no affinity to the ditrig-
onal cavities at this hydration state. Furthermore, it is seen
that the peaks of the diffuse-layer at the interlayer midplane in
Na-MMT disappeared [Fig. 6(a)], indicating that the Na ions
are solvated in the diffuse layer region. The Ca ions mainly
reside in outer-sphere complexes (peaks at about 7.7 Å from
the clay midplane) and are solvated in the diffuse layer region
as well [Fig. 6(b)]. It is worth noting that the water concen-
tration for Ca-MMT in the 5W is approximately 55 M, i.e.,
the water density is about 1.0 g/cm3at the interlayer midplane
indicating that the start of the formation of the bulk water
reservoir.
B. Structure of interlayer hydration
Figure 7(a) shows the radial distribution function (RDF)
between Na-OW (Na is the reference solute, OW represents
the ligand) from the simulations for Na-MMT in the 1W, 2W,
3W, 4W, and 5W hydration states. It is seen that the first peak
in the RDF is located at 2.34 Å for the 1W hydrate state.
For higher hydration states, the primary peak in the RDF’s
is found to reside at about 2.36 Å. This initial shift indicates
that the water molecules have more space to orient themselves
with increasing the interlayer separation. Figure 7(b) shows
the corresponding coordination number (CN) of Na coordi-
nating to OW. It is seen that, in the 1W hydrate state, there are
FIG. 8. Polyhedral structure of Na ions hydrated in the interlayer of Na-MMT
in (a) 1W and (b) 5W. Only Na ions (dark blue) and water oxygen atoms OW
(red) of the first hydration shell in the interlayer are shown. The clay mineral
structures are shaded for clarity.
four water molecules (i.e., OW) in the first hydration shell of
Na ion. Figure 8(a) further illustrates that for the 1W hydra-
tion state, each Na ion is coordinated to four water molecules
and one basal oxygen atom from the surface of MMT.
Figure 7(b) shows that the number of water molecules in the
first hydration shell is increased from four to six when increas-
ing the water content from the 1W to the 5W hydration state.
It is seen that each Na ion in the interlayer is coordinated by
five to six water molecules forming orthogonal coordination
sphere [Fig. 8(b)] with a radius of 3.1 Å [which is the posi-
tion of the first minimum in the RDF for Na-OW shown in
Fig. 7(a)] in the 5W hydration state. Figure 7(c) presents the
average distances of the nearest-neighbor and next nearest-
neighbor water molecules about a reference water molecule
are 2.74 Å (the first peak) and 4.66 Å (the second peak),
respectively. The average distance of hydrogen bond in the
interlayer is found to be about 1.9 Å [the first peak shown in
Fig. 7(d)].
Figure 9shows the corresponding data presented in Fig. 7,
but here data are shown for Ca-MMT at different hydration
states. In Fig. 9(a), the first peak of the RDF of Ca-OW is
shifted from 2.36 Å to 2.38 Å when the water content is
increased from the 1W to the 5W hydration state. There are
seven water molecules in the first shell of the Ca hydration
complex [Fig. 10(a)] when Ca-MMT is in the 1W hydra-
tion state. It is seen that the coordination number of water
molecules in the first hydration shell is increased from seven
to eight when increasing the water contents from 1W to 5W
FIG. 7. Graphs of interlayer (a) RDF
for Na-OW, (b) coordinate number
CNNa-OW (r), (c) RDF for OW-OW, and
(d) RDF for HW-OW in Na-MMT.
084705-6 Yang, Neretnieks, and Holmboe J. Chem. Phys. 147, 084705 (2017)
FIG. 9. Graphs of the interlayer (a)
RDF for Ca-OW, (b) coordinate num-
ber CNCa-OW (r), (c) RDF for OW-OW,
and (d) RDF for HW-OW in Ca-MMT.
[Fig. 9(b)]. Furthermore, at the 5W hydration state, each Ca
ion is coordinated to eight water molecules, forming orthog-
onal coordination sphere [Fig. 10(b)] with a radius of 3.3 Å
[which is the position of the first minimum in the RDF for Ca-
OW shown in Fig. 9(a)]. Figure 9(c) shows that the average
distances of the nearest-neighbor and next nearest-neighbor
water molecules about a reference molecule are 2.8 Å (the
first peak) and 4.55 Å (the second peak). The average distance
of hydrogen bond is about 1.8 Å (the first peak) as shown in
Fig. 9(d).
C. Deformation of hydration shells
Figure 11 shows the deformation of the hydration shells
coordinated around Ca ions in the 1W, 2W, 3W, and 5W hydra-
tion states. In Fig. 11, only the interlayer Ca ions (green) and
water oxygen atoms OW (red) in the first coordination shell
are shown. Figure 11(a) shows that there are about seven water
molecules coordinated around each Ca ion in the first hydra-
tion shell when Ca-MMT is in the 1W hydration state. It is
seen that the water molecules are squeezed out from the top
and bottom of the first shell and prefer to reside in the center of
a ring in the xy-plane, due to the limited space. For Ca-MMT
in the 2W hydration state, eight water molecules coordinate to
Ca in the first shell, as shown in Fig. 11(b). Only a minor defor-
mation of the first coordination water shell was observed [such
as the one in the right corner in Fig. 10(b)], since the hydration
FIG. 10. Polyhedral structure of Ca ions hydrated in the interlayer for Ca-
MMT in (a) 1W and (b) 5W. Ca ions (light blue) and water oxygen atoms OW
(red) of the first hydration shell in the interlayer are shown. The clay mineral
structures are shaded for clarity.
shells were found to be nearly spherical. Figure 11(c) shows
that Ca ions are fully hydrated forming evenly distributed
spherical shells for Ca-MMT in the 3W hydration state. In
Fig. 11(d), there are nearly eight water molecules coordinated
around each Ca ion in the first hydration shell for Ca-MMT in
the 5W hydration state. In Fig. 11(d), the complete hydration
shells are observed forming orthogonal coordination spheres
[Fig. 10(b)] with radius of approximately 3.3 Å without any
deformation.
D. Comparison between MD and classical DFT
on the EDL
The Stern-layer effect plays an essential role in explain-
ing the swelling mechanism of clay minerals. In order to reveal
under which conditions our previous modeling using density
functional theory (DFT) becomes increasingly inaccurate with
decreasing interlayer separation, we performed a compari-
son with MD simulations conducted in this study. The blue
curves presented in Fig. 12 come from the reference fluid
FIG. 11. Snapshot of MD simulations showing deformation of hydration
shells in the interlayer of Ca-MMT in (a) 1W, (b) 2W, (c) 3W, and (d) 5W. Only
Ca ions (green) and water oxygen atoms OW (red) of the first hydration shell
are shown in the picture. The clay mineral structures are shaded for clarity.
084705-7 Yang, Neretnieks, and Holmboe J. Chem. Phys. 147, 084705 (2017)
density/weighted correlation approach (RFD/WCA) within
the framework of DFT. The detailed information of the DFT
method can be found in the papers.9,11 The red curves are from
MD simulations. The gray dashed lines represent the midplane
to basal-plane distance (half a MMT layer). In the DFT cal-
culations, the diameters of Na and Ca ions from CLAYFF
forcefield47 were used. For simplicity, the ions were described
as hard spheres of equal size with point charges assumed
to be located at the ion centers, and the solvent represented
by a continuum with a uniform dielectric permittivity (εr
= 78.4) at room temperature (T= 298 K), next to charged
surfaces in the DFT calculations for the restricted primitive
model.
Figure 12(a) illustrates the MD simulation cell with
Na-MMT having an excess electrolyte concentration with
24 excess Na–Cl ions in a ten-layer hydration state (10 W),
containing 3600 H2O water molecules (containing 50 water
molecules per unit cell), and 13 776 atoms in total (including
atoms in clay mineral structure). The size of the simulation
cell was 31.0 Å ×54.2 Å ×82.9 Å (x×y×z), result-
ing in an equilibrated interlayer separation of approximately
30 Å. The corresponding DFT calculations were performed
for a similar system where Na-MMT system is in equilibrium
with a NaCl bulk solution of 0.62M. The bulk ion concentra-
tions were determined from the density profiles in middle of
the pore. The ionic diameters and the surface charge density
used in DFT system were 4.848 Å and 0.11 C m 2, respec-
tively, hence similar surface charge density used in the MD
simulations. It was shown that the DFT calculations success-
fully capture the diffuse ion swarm from approximately 5.0 Å
beyond the basal-plane in comparison with MD simulations.
However, the features of the main outer-sphere complexes
(within 5.0 Å from the basal-plane) and inner-sphere com-
plexes observed by MD simulations were not predicted by the
DFT results.
The MD simulations presented in Fig. 12(b) show a cor-
responding Ca-MMT MD system in the ten-layer hydration
state (10W), containing an additional 6 Ca and 12 Cl ions,
resulting in the cell dimensions of 31.0 Å ×54.2 Å ×83.2 Å
FIG. 12. Comparison of interlayer density profile from MD (red curves) and
DFT (blue curves) for (a) Na-MMT and (b) Ca-MMT. Solid curves repre-
sent the counter-ion distributions, and dashed curves represent the co-ion
distributions.
FIG. 13. Snapshot of MD simulations cell showing Ca-MMT in the ten-layer
hydrate. Ca2+ (green) ions, Cl (dark blue), and water molecules. The clay
mineral structure contains Si (yellow), Al (pink), Mg (light blue), O (red), and
H (white) atoms.
(x×y×z), see also Fig. 13. After reaching equilibrium, the
bulk CaCl2salt concentration was about 0.23M, and the inter-
layer separation approximately 31.6 Å. It is compared with the
corresponding DFT system in equilibrium with a 0.23M CaCl2
solution. Here, the ionic diameters and the surface charge den-
sity were 5.3 Å and 0.11 C m 2, respectively. Analogously to
the case with Na ions, the steric and hydration effects on the
Stern-layer are absent in the DFT predictions.
IV. CONCLUSIONS
To better understand the mechanism causing the swelling
of smectite clays and interlayer hydration, and the effects of ion
hydration on the molecular level for the EDL, we modeled Na-
and Ca-MMT at different water loadings using MD, explicitly
taking into account the interactions of the water molecules. The
MD results showed that the smectite basal spacing (d001) was
about 12.5 Å, with Na ion coordinated to four water molecules
and one basal oxygen atom from the surface of Na-MMT in
the 1W hydration state. It was found that Na-MMT resulted in
d001 of about 15.6 Å, 18.7 Å, 21.7 Å, and 24.8 Å in the 2W, 3W,
4W, and 5W hydration states, respectively, in which each Na
ion was coordinated to nearly six water molecules in the first
hydration shell within a radius of 3.1 Å. It was shown that Ca-
MMT in the 1W hydration state resulted in d001 of about 13.7 Å
(slightly larger than the experimental value), with Ca ion being
hydrated by approximately seven water molecules in its first
084705-8 Yang, Neretnieks, and Holmboe J. Chem. Phys. 147, 084705 (2017)
hydration shell, which was deformed and non-symmetrical. In
the 2W hydration state, d001 was found to be about 15.6 Å
for Ca-MMT (well in line with the experimental data), with
each Ca ion being hydrated by nearly eight water molecules
in the first shell. Only a minor deformation of the first coor-
dination water shell was observed for Ca-MMT in 2W, since
the hydration shells were found to be nearly spherical. For the
higher hydration states, it was found that the Ca ions were fully
hydrated with evenly distributed spherical coordination shells,
resulting in d001 of 18.7 Å, 21.7 Å, and 24.8 Å, respectively,
in the 3W, 4W, and 5W hydration state. Moreover, eight water
molecules were coordinated to Ca ion in the first coordination
shell, forming orthogonal coordination spheres with a radius
of about 3.3 Å. Hence, it was concluded that complete hydra-
tion shells could only be formed when the basal spacing is
larger than approximately 18.7 Å.
ACKNOWLEDGMENTS
The authors would like to acknowledge the financial
support of the Swedish Nuclear Fuel and Waste Manage-
ment Company (SKB). The simulations were performed on
resources provided by the Swedish National Infrastructure
for Computing (SNIC) at PDC and HPC2N Centers for High
Performance Computing.
1F. R. C. Chang, N. T. Skipper, and G. Sposito, Langmuir 11, 2734–2741
(1995).
2F. R. C. Chang, N. T. Skipper, and G. Sposito, Langmuir 14, 1201–1207
(1998).
3L. Tao, T. Xiao-Feng, Z. Yu, and G. Tao, Chin. Phys. B 19, 109101 (2010).
4N. T. Skipper, K. Refson, and J. D. C. McConnell, J. Chem. Phys. 94, 7434
(1991).
5N. T.Skipper, F. R. C. Chang, and G. Sposito, Clays Clay Miner.43, 285–293
(1995).
6N. T.Skipper, F. R. C. Chang, and G. Sposito, Clays Clay Miner.43, 294–303
(1995).
7N. T. Skipper, P. A. Lock, J. O. Titiloye, J. Swenson, Z. A. Mirza, W.
S. Howells, and F. Fernandez-Alonso, Chem. Geol. 230, 182–196 (2006).
8G. Sposito, S. H. Park, and R. Sutton, Clays Clay Miner. 47, 192–200
(1999).
9G. Yang and L. Liu, J. Chem. Phys. 142, 194110 (2015).
10G. Yang, I. Neretnieks, L. Moreno, and S. Wold, Appl. Clay Sci. 132-133,
561–570 (2016).
11G. Yang, I. Neretnieks, L. Moreno, and S. Wold, Appl. Clay Sci. 135,
526–531 (2017).
12C. A. Angell, Y. Ansari, and Z. Zhao, Farad. Discuss. 154, 9–27 (2012).
13M. Gali ´
nski, A. Lewandowski, and I. Ste˛pniak, Electrochim. Acta 51,
5567–5580 (2006).
14Z. Tang, L. E. Scriven, and H. T. Davis, J. Chem. Phys. 97, 494 (1992).
15D. Boda, D. Henderson, A. Patrykiejew, and S. Sokołowski, J. Colloid
Interface Sci. 239, 432–439 (2001).
16L. B. Bhuiyan, S. Lamperski, J. Wu, and D. Henderson, J. Phys. Chem. B
116, 10364 (2012).
17M. V. Fedorov, N. Georgi, and A. A. Kornyshev, Electrochem. Commun.
12, 296 (2010).
18D. Henderson and J. Wu, J. Phys. Chem. B 116, 2520 (2012).
19D. Henderson, S. Lamperski, Z. Jin, and J. Wu, J. Phys. Chem. B 115, 12911
(2011).
20M. Kaja, W. Silvestre-Alcantara, S. Lamperski, D. Henderson, and
L. B. Bhuiyan, Mol. Phys. 113, 1043–1052 (2015).
21O. Matsuoka, E. Clementi, and M. Yoshimine, J. Chem. Phys. 64, 1351
(1976).
22S. Lamperski, M. Kaja, L. B. Bhuiyan, J. Wu, and D. Henderson, J. Chem.
Phys. 139, 054703 (2013).
23J. Wu, T. Jiang, D.-e. Jiang, Z. Jin, and D. Henderson, Soft Matter 7, 11222
(2011).
24W. Silvestre-Alcantara, D. Henderson, J. Wu, M. Kaja, S. Lamperski, and
L. B. Bhuiyan, J. Colloid Interface Sci. 449, 175–179 (2015).
25W. Silvestre-Alcantara, M. Kaja, D. Henderson, S. Lamperski, and
L. B. Bhuiyan, Mol. Phys. 114, 53–60 (2016).
26D. Henderson, W. Silvestre-Alcantara, M. Kaja, S. Lamperski, J. Wu, and
L. B. Bhuiyan, J. Mol. Liq. 228, 236–242 (2016).
27J. Barthel, R. Buchner, P. N. Ebersp¨
acher, M. M¨
unsterer, J. Stauber, and
B. Wurm, J. Mol. Liq. 78, 83–109 (1998).
28J.-L. Liu and B. Eisenberg, Chem. J. Phys. Lett. 637, 1–6 (2015).
29J. Vincze, M. Valisk´
o, and D. Boda, J. Chem. Phys. 133, 154507 (2010).
30J. Vincze, M. Valisk´
o, and D. Boda, J. Chem. Phys. 134, 157102 (2011).
31M. Valisk´
o and D. Boda, J. Chem. Phys. 140, 234508 (2014).
32M. Valisk´
o and D. Boda, J. Phys. Chem. B 119, 1546–1557 (2015).
33M. Valisk´
o and D. Boda, Mol. Phys. 115, 1245 (2017).
34J. A. Greathouse and R. T. Cygan, Environ. Sci. Technol. 40, 3865
(2006).
35V. Marry, B. Rotenberg, and P. Turq, Phys. Chem. Chem. Phys. 10, 4802
(2008).
36B. Rotenberg, V.Marry, R. Vuilleumier, N. Malikova, C. Simon, and P. Turq,
Geochim. Cosmochim. Acta 71, 5089–5101 (2007).
37J. W. Lee, J. A. Templeton, K. K. Mandadapu, and J. A. Zimmerman,
J. Chem. Theory Comput. 9, 3051–3061 (2013).
38M. Holmboe and I. C. Bourg, J. Phys. Chem. C 118, 1001–1013 (2014).
39R. M. Tinnacher, M. Holmboe, C. Tournassat, I. C. Bourg, and J. A. Davis,
Geochim. Cosmochim. Acta 117, 130–149 (2016).
40M. J. Abraham, T. Murtola, R. Schulz, S. P´
all, J. C. Smith, B. Hess, and
E. Lindahl, SoftwareX 1-2, 19–25 (2015).
41H. J. C. Berendsen, D. van der Spoel, and R. van Drunen, Comput. Phys.
Commun. 91, 43–56 (1995).
42B. Hess, C. Kutzner, D. van der Spoel, and E. Lindahl, J. Chem. Theory
Comput. 4, 435–447 (2008).
43E. Lindahl, B. Hess, and D. van der Spoel, J. Mol. Model. 7, 306–317
(2001).
44D. van der Spoel, E. Lindahl, B. Hess, G. Groenhof, A. E. Mark, and
H. J. C. Berendsen, J. Comput. Chem. 26, 1701–1718 (2005).
45S. Plimpton, J. Comput. Phys. 117, 1–42 (1995).
46H. J. C. Berendsen, J. R. Grigera, and T. P. Straatsma, J. Phys. Chem. 91,
6269–6271 (1987).
47R. T. Cygan, J. J. Liang, and A. G. Kalinichev, J. Phys. Chem. B 108,
1255–1266 (2004).
48J. Åqvist, J. Phys. Chem. 94, 8021–8024 (1990).
49I. S. Joung and T. E. Cheatham III, J. Phys. Chem. B 112, 9020–9041 (2008).
50E. Ferrage, B. A. Sakharov, L. J. Michot, A. Delville, A. Bauer, B. Lanson,
S. Grangeon, G. Frapper, M. Jim´
enez-Ruiz, and G. J. Cuello, J. Phys. Chem.
C115, 1867–1881 (2011).
51B. Dazas, E. Ferrage, A. Delville, and B. Lanson, Am. Mineral. 99,
1724–1735 (2014).
52T. Darden, D. York, and L. Pederson, J. Chem. Phys. 98, 10089 (1993).
53T. J. Tambach, E. J. M. Hensen, and B. Smit, J. Phys. Chem. B 108,
7586–7596 (2004).
54H. J. C. Berendsen, J. P. M. Postma, W. F. van Gunsteren, A. DiNola, and
J. R. Haak, J. Chem. Phys. 81, 3684–3690 (1984).
55M. Holmboe, S. Wold, and M. Jonsson, J. Contam. Hydrol. 128, 19–32
(2012).
... The smectite clay chosen for this study is Wyoming Na-montmorillonite (SWy) -a wellcharacterised and most commonly studied clay in the laboratory due to its easy purchase from the Clay Minerals Society. While the molecular simulations of montmorillonite clays have also been done frequently, 23 4 , ensuring that the isomorphic substitutions are random, but are not neighbouring. Since the valance of octahedral iron in clay may be both II and III, based on the known charge per unit cell, the SWy MMT only features Fe III . ...
... 27,28 Therefore, accounting for hydration shells of Na + and Ca 2+ cations, we used 12 water molecules per counter ion, which resulted in the desired interlayer spacing. 23,27 We performed molecular dynamics simulations of the 0. With nearly all molecules fully protonated under a pH of 6.5, Api 0 precipitates, forming orderly π-π stacks with a crystalline structure. No effect of cations is observed. ...
Preprint
Full-text available
Interactions between organic species and natural minerals are fundamental to the processes around us. With the aid of molecular dynamics simulations, we identify key adsorption mechanisms of apigenin on smectite clay minerals. The mechanism is highly sensitive to the pH -- changing from co-crystallisation in acidic-to-neutral solutions to the ion-bridging in mild-alkaline. The ionic species play a significant role in alkaline environments: the deprotonated apigenin species chelate metals, which, in turn, leads to the formation of a stable organic-metal-mineral complex and stronger adsorption in the presence of divalent cations.Smectite clays buffer the solution to mildly alkaline; hence, the type of exchangeable cations in the clay will be critical in determining the adsorption mechanism and organic retention capacity. Overall, our study showcases a computational strategy that can be transferred to a wide variety of organic-mineral systems in the natural environment.
... This occurs because Ca 2+ is displaced from sorption sites on flocculated clays by Na + . This outcome is due to the thin Na + hydration shell (3.1 Å) that attaches to a hydrated clay surface with one oxygen atom and four water molecules, whereas Ca 2+ has a thicker hydration shell (3.3 Å) containing seven water molecules in its outer-sphere complex (Yang et al., 2017). Physically, Na + is small enough to enter the interlattice and expand the double diffuse layer within clay sheets, which causes swelling of clay minerals. ...
Article
Sodium-rich water may deteriorate calcite-stabilize soil. The objective of this work was to determine whether calcite solution prepared with sodium-rich water were capable of forming stable soil macroaggregates. Calcite solutions were 2.6 g Ca2+ L-1 of CaCO3 mixed with water containing 2.6, 5 or 10 g Na+ L-1 from NaCl. Arid desert soil (5 cm layer, packed at 1.1 g cm-3) was moistened with calcite solution at 50% or 75% of the water holding capacity. Then, the calcite-treated soil was incubated for up to 30 d at average 35 ◦C and 19% relative air humidity. Clay flocculation rate was estimated from the decline in extractable Ca2+ and Na+ concentrations during macroaggregate formation. We observed carbonate precipitation in voids of the macroaggregates using micro-CT scanning, which is evidence of cementation. Aggregate stability was the mean weight diameter of macroaggregates. Macroaggregate stability depends upon calcite interactions with clay, so a higher proportion of Ca2+ to Na+ favors the formation of flocculants and cements the flocculants within macroaggregates with carbonate materials, which is surmised to be sodium carbonate with low resistance to abrasion. There was 40% more cementing material within macroaggregates treated with pure calcite and Ca2+: Na+ solutions than with solutions containing Ca2+: 2 Na+, Ca2+: 4 Na+ and Na+. We conclude that calcite-stabilized macroaggregates will persist longer when formed by a calcite solution with equal proportions of Ca2+:Na+, since macroaggregates disintegrate more quickly when formed from calcite made with brackish water at ratios of Ca2+ ≤ 2 Na+.
... This might be because of the stronger Ca 2+ −clay interactions in the 1L state. 6, 43,[45][46][47]49,50 Our studies show that the solvent type significantly influences the clay swelling process and the species mobilities and can guide further research in this field. Such an understanding of the interaction between the clay and the 44 96.0 43 (1L) 38 10.0−30.0 ...
Article
Molecular simulations were performed to explore the adsorption and transport mechanism of ethanol and ions in Na- and Ca-montmorillonite clays. Our results show that the uptake of ethanol by montmorillonite increases with increasing relative pressure (RP)/basal d-spacing, consistent with experimental observations. The basal d-spacing of montmorillonite grows in the presence of ethanol to about 13.0 Å with a monolayer arrangement of ethanol (1L). Further uptake of ethanol allows the basal d-spacing to grow to about 16.5 Å with a bilayer arrangement of ethanol (2L). For a given solvation state, the amount of ethanol uptake is almost independent of RP and the type of the counterion. The stable basal d-spacings of the montmorillonite+ethanol system are larger than those of the montmorillonite+water system. Also, the swelling transitions are relatively shifted to higher RP values in the montmorillonite+ethanol system. This may be because the clay has a weaker affinity for the less polar ethanol molecules as compared with water. The mobility of ethanol and ions in the interlayers increases with increasing RP, due to the associated swelling of montmorillonite. For a given solvation state, the mobility of these species is almost unaffected by changes in RP, as in the case of adsorption. The mobility of ethanol and ions in 1L and 2L states is about 3 order of magnitude lower than that in the bulk. The species mobility in the montmorillonite+ethanol system is generally much lower than those in the montmorillonite+water system. This may be attributed to the higher steric hindrance in the alcohol molecules. However, the mobility of the Ca2+ ion in the 1L state is almost similar in both the montmorillonite+ethanol and the montmorillonite+water systems. This is possibly due to the stronger Ca2+-clay interactions in the 1L state.
... Therefore, it is necessary to understand the effect of hydration of interlayer cations on the size of the interlayer space of clay minerals. Many authors have studied the effect of different types of interlayer ions of alkali and alkaline earth metals on crystal swelling [86,87]. The results of these studies show that the degree of crystal swelling depends on the intensity of cation hydration in the order Na-MMT > K-MMT > Cs-MMT > Mg-MMT > Ca-MMT. ...
Article
Clay is a widespread natural mineral. The review considers physical and chemical properties of clay minerals which are important in terms of geological high-level radioactive waste disposal (HLRW). The articles under consideration present that the properties of clay as a barrier material for the isolation of radionuclides are influenced by temperature, density (external pressure), ionic strength and pH of the solution, and the presence of cations of stable elements. These conditions are studied by both experimental and calculation methods. However, there are not enough research results in the publications for conditions close to those of HLRW geological disposal. The research status worldwide is introduced in detail, and at the end of the article, the future directions of the research on radionuclide diffusion in HLRW disposal are discussed. https://rdcu.be/cQFRk
... With decreasing interlayer distances, the hydration shell may deform, eventually resulting in water molecules being squeezed out from the inner hydration shell. However, some other studies have shown that not all cations were hydrated, some ones stay on clay surface to balance the negative surface charges and build up the relatively strong electrostatic interaction to resist external loading even at high water content which depend on the category of cation, charge, the site of isomorphous substitution and so on Yang et al. (2017), Subramanian et al. (2020), Zhang et al. (2014), Teich-McGoldrick et al. (2012. In addition, hydrogen bond network, as a major contributor to mechanical strength at moist state in general, has been studied extensively. ...
Article
The mechanical behavior of clayey rocks is generally sensitive to water saturation degree, mainly due to the high content of swelling clay minerals. In this study, the mechanical behavior of Montmorillonite (MMT) crystal with different water content at the atomistic scale is investigated by using molecular dynamics (MD) simulation. In accordance with macroscopic laboratory tests, we consider here triaxial compression with constant mean stress respectively for the dry and hydrated MMT systems containing one, two and three water layers. While whole stress–strain responses are calculated, the emphasis is put on the key mechanisms which are related to water saturation effect on both failure and deformation properties. It is found that the mechanical behavior of a hydrated MMT system is not only dependent on the number but also the arrangement of interlayer water molecules. In particular, the water sensitivity of mechanical behavior is driven by both the clay–solution (water and Na cation mixture) and the solution–solution interactions. It seems that the MMT system with the two-layer hydrated state provides a more stable arrangement of water molecules and has a higher mechanical strength than the other two ones. The results issued from this study brings a microscopic insight onto the water sensitivity of macroscopic deformation and failure properties of clayey rocks.
Article
Clay is a nanomaterial by nature. The cracking in clay at the nanoscale can provide significant insight into the cracking mechanism in clay at the continuum scale. This article is devoted to understanding the atomic-scale mechanism of the formation of cracks in a dry clay sheet through molecular dynamics with a general force field for clay. We simulate the formations of mode I and mode II cracks in intact dry clay sheets under tension and shear loading conditions, respectively. The numerical results show that the mode I and II cracking in a clay sheet is brittle and strain-rate dependent. The notion of critical bond stretch is adopted to model the breakage of bonds between atoms in clay. The critical bond length for crack formation is determined by the radial distribution function and the force curve between a pair of atoms. We investigate the crack initiation under mode I and mode II loading conditions from the bond breakage analysis. We interpret the cracking mechanism in terms of the type and number of broken bonds in mode I and mode II cracking. We present the stress intensity factor and cracking energy release rate under both cracking modes. The numerical results of the stress intensity factor and cracking energy release rate are quantitatively comparable to the experimental and numerical data in the literature.
Article
This study was intended to explore the effects of Chlorhexidine di(acetate)-Na-Monotmorillonite (CA-MMT) nanoparticles on implantable venous access port (PORT) nursing of hematopathy patients, which was expected to provide a more effective alternative for PORT nursing and alleviate the suffering of hematopathy patients. CA-MMT nanoparticles were prepared and characterized by X-ray and infrared spectroscopy. Specifically, CA and Na-MMT were dissolved in double distilled water, and then added to a round-bottomed flask and mixed uniformly. After the mixture was condensed and refluxed, the product in the flask was filtered, and then washed with double distilled water for several times, followed by suction filtration. The obtained off-white block solid was dried at high temperature, and then ground and sieved to obtain CA-MMT nanoparticles. 64 hematopathy patients were selected as research subjects and were randomly divided into control group and nano group. After the PORT surgery, the nano group had CA-MMT nursing, with the control group having iodophor nursing. The routine blood indexes and the phlebitis incidence were compared between the two groups for analysis. The results showed that the CA was successfully intercalated between layers of MMT, which had a relatively good dispersion property with a regular shape. Before the puncture, the two groups showed no notable differences ( P > 0.05), while after the needle was removed, a notable difference was noted ( P < 0.05). The phlebitis probability in the control and the nano group was 37.5% and 15.6%, respectively, indicating a notable difference ( P < 0.05). In conclusion, the CA is successfully intercalated between MMT layers. The CA-MMT nanoparticle is instrumental in reducing the incidence of infections and phlebitis, demonstrating good effects on PORT nursing of hematopathy patients.
Article
Full-text available
Clay-rich media have been proposed as engineered barrier materials or host rocks for high level radioactive waste repositories in several countries. Hence, a detailed understanding of adsorption and diffusion in these materials is needed, not only for radioactive contaminants, but also for predominant earth metals, which can affect radionuclide speciation and diffusion. The prediction of adsorption and diffusion in clay-rich media, however, is complicated by the similarity between the width of clay nanopores and the thickness of the electrical double layer (EDL) at charged clay mineral–water interfaces. Because of this similarity, the distinction between ‘bulk liquid’ water and ‘surface’ water (i.e., EDL water) in clayey media can be ambiguous. Hence, the goal of this study was to examine the ability of existing pore scale conceptual models (single porosity models) to link molecular and macroscopic scale data on adsorption and diffusion in compacted smectite. Macroscopic scale measurements of the adsorption and diffusion of calcium, bromide, and tritiated water in Na-montmorillonite were modeled using a multi-component reactive transport approach while testing a variety of conceptual models of pore scale properties (adsorption and diffusion in individual pores). Molecular dynamics (MD) simulations were carried out under conditions similar to those of our macroscopic scale diffusion experiments to help constrain the pore scale models. Our results indicate that single porosity models cannot be simultaneously consistent with our MD simulation results and our macroscopic scale diffusion data. A dual porosity model, which allows for the existence of a significant fraction of bulk liquid water—even at conditions where the average pore width is only a few nanometers—may be required to describe both pore scale and macroscopic scale data.
Article
Full-text available
GROMACS is one of the most widely used open-source and free software codes in chemistry, used primarily for dynamical simulations of biomolecules. It provides a rich set of calculation types, preparation and analysis tools. Several advanced techniques for free-energy calculations are supported. In version 5, it reaches new performance heights, through several new and enhanced parallelization algorithms. These work on every level; SIMD registers inside cores, multithreading, heterogeneous CPU-GPU acceleration, state-of-the-art 3D domain decomposition, and ensemble-level parallelization through built-in replica exchange and the separate Copernicus framework. The latest best-in-class compressed trajectory storage format is supported.
Article
We investigate the individual activity coefficients of ions in LaCl3 using our theory that is based on the competition of ion–ion (II) and ion–water (IW) interactions. The II term is computed from Grand Canonical Monte Carlo simulations on the basis of the implicit solvent model of electrolytes using hard sphere ions with Pauling radii. The IW term is computed on the basis of Born’s treatment of solvation using experimental hydration-free energies. The results show good agreement with experimental data for La3+. This agreement is remarkable considering the facts that (i) the result is the balance of two terms that are large in absolute value (up to 20 kT) but opposite in sign, and (ii) that our model does not contain any adjustable parameter. All the parameters used in the model are taken fromexperiments: concentration-dependent dielectric constant, hydration free energies and Pauling radii.
Article
The density functional theory is applied to a study of the structure and differential capacitance of a planar electric double layer formed by a valency asymmetric mixture of charged dimers and monomers. The dimer consists of two tangentially tethered hard spheres of equal diameters of which one is charged and the other is neutral, while the monomer is a charged hard sphere of the same size. The dimer electrolyte is next to a uniformly charged, smooth planar electrode. The electrode-particle singlet distributions, the mean electrostatic potential, and the differential capacitance for the model double layer are evaluated for a 2:1/1:2 valency electrolyte at a given concentration. Important consequences of asymmetry in charges and in ion shapes are (i) a finite, non-zero potential of zero charge, and (ii) asymmetric shaped 2:1 and 1:2 capacitance curves which are not mirror images of each other. Comparisons of the density functional results with the corresponding Monte Carlo simulations show the theoretical predictions to be in good agreement with the simulations overall except near zero surface charge.
Article
Canonical Monte Carlo simulation results are obtained for the structure and differential capacitance of a planar electric double layer consisting of fused dimer cations and monomer anions in a dielectric continuum (solvent). The physical double layer model is mimicked by a uniformly charged, non-polarisable, planar electrode next to an electrolyte consisting of positively charged fused dimers and negatively charged monomer rigid spheres in a continuum dielectric. The fused dimer comprises two same sized hard spheres that are fused into one another, one of the hard spheres being charged and the other neutral. The various electrode-particle species singlet distribution functions and the mean electrostatic potential profiles are obtained for 1:1, 2:2, 2:1, and 1:2 valency systems for different degrees of fusion. The differential capacitance as a function of the electrode surface charge density is evaluated for a 1:1 system. Comparison of these results are made with those corresponding to the situation when the dimer spheres are simply tangentially tethered. It is seen that the fused dimer influences the double layer structure and capacitance substantially for a negatively charged electrode but less so for a positively charged electrode.
Article
Monte Carlo simulations of a model ionic liquid show that if ions have charged heads and neutral counterparts, the latter give rise to the camel shape of the voltage dependence of the double layer capacitance. Neutral 'tails' of ions play the role of latent voids that can be replaced by charged groups via rotations and translations of ions. This provides extra degrees of freedom for the field-induced charge rearrangements in the double layer which results in the peculiar double-hump capacitance profile.
Article
Canonical and grand canonical Monte Carlo simulation results are reported for the structure of a planar electric double layer containing anisotropic-shaped ions having asymmetric valencies. The model double layer consists of an electrolyte formed by a mixture of charged dimers and charged hard spheres in a dielectric continuum next to a uniformly charged, non-polarisable, planar hard electrode. A dimer is made of two tangentially touching equi-sized hard spheres one of which is positively charged and the other neutral, while the monomer rigid ion of the same size is negatively charged. Results for the electrode-ion, electrode-neutral sphere singlet distributions, and the mean electrostatic potential are obtained for 2:1/1:2 valency asymmetry in the electrolyte solution regime at room temperature, and at a given electrolyte concentration for a series of electrode surface charge densities. Valency asymmetry coupled with anisotropic ion shape leads to a richer double-layer structure than seen previously with dimers in 1:1 symmetric valency situations. Comparisons are also made with the corresponding results for a 2:1/1:2 restricted primitive model electrolyte. The overall asymmetry also leads to a finite, non-zero potential of zero charge indicating a charge separation at the interface.