Anda di halaman 1dari 5

Fluid Phase Equilibria 409 (2016) 434e438

Contents lists available at ScienceDirect

Fluid Phase Equilibria


j o u r n a l h o m e p a g e : w w w . e l s e v i e r . c o m / l o c a t e / fl u i d

Short communication

Stability of the kaolinite-guest molecule intercalation system: A


molecular simulation study
bor Rutkai a, Zolta
n Hato
 b, Tama
s Kristo
 f b, *
Ga
a
b

Thermodynamics and Energy Technology, University of Paderborn, Warburger Str. 100, D-33098 Paderborn, Germany
Institute of Chemistry, Department of Physical Chemistry, University of Pannonia, P.O. Box 158, H-8201 Veszpr
em, Hungary

a r t i c l e i n f o

a b s t r a c t

Article history:
Received 30 July 2015
Received in revised form
30 October 2015
Accepted 31 October 2015
Available online 3 November 2015

Molecular dynamics and Monte Carlo simulations are used to determine the stable states of the
kaolinite-methanol intercalation system. Several evaluation scenarios are presented to describe the
stability of the system using single mSolventNClaypT ensemble simulations and a series of NpT ensemble
simulations coupled with thermodynamic integration.
2015 Elsevier B.V. All rights reserved.

Keywords:
Kaolinite
Intercalation
Methanol
Molecular simulation
Thermodynamic integration

1. Introduction
Clay minerals with layered structure on the molecular level are
capable of adsorbing organic and inorganic guest molecules in their
interlayer space. The phenomenon itself is referred as intercalation
and it is of interest due to its possible application in the design of
many clay mineral based nanomaterials. These materials are
promising candidates for catalysts, adsorbents, and composites [1].
When synthesizing nanomaterials from layered clays, the effect of
the intercalated molecules on the layers is diverse. They can change
the surface properties of the mineral layers due to chemical reactions, or simply make the interlayer space expand for other
molecules to enter that do not intercalate directly by themselves. At
some point, it also becomes essential that the guest molecules
completely push the molecular layers apart producing delaminated
layers that can form various nanostructures such as curled layers,
tubes, or scrolls. Although kaolinite is one of the most common
layered clay mineral found in nature and used routinely in the
paper, ceramic, medicine, and cosmetic industry, its role as building
block for the construction of nanomaterials, compared to other
expandable layered materials [2] has been less prominent, yet

* Corresponding author.
 f).
E-mail address: kristoft@almos.vein.hu (T. Kristo
http://dx.doi.org/10.1016/j.uid.2015.10.045
0378-3812/ 2015 Elsevier B.V. All rights reserved.

successful [3,4]. The reason behind this is that the clay layers in
kaolinite are held together by hydrogen bonds in its pure mineral
form and only a few highly polar molecules, such as hydrazine,
urea, potassium acetate, formamide, and dimethylsulfoxide, can be
used as a direct expansion compound [5,6]. On the other hand,
exactly the dipolar character of the interlayer spaces makes
kaolinite so appealing for materials design purposes. The range of
potential guest species can be considerably extended by applying
intermediate substances. These molecules displace the initially
intercalated highly polar ones. After that, they are replaced by longchain organic molecules in subsequent steps. This nally causes the
complete separation of the kaolinite layers. Methanol has proven to
be such an important intermediate [3,4,7] and it is the target of the
current investigation. In this work, we model the methanolkaolinite intercalation mixture without the possible chemical reactions between the surface of the clay layers and methanol
[3,4,7,8]. The results presented here serve as a clear staring point
and reference for further investigation purposes.
2. Methods of choice
Molecular simulation has evolved to point where it can effectively contribute to the understanding and detailed description of a
variety of engineering related processes including the swelling of
clay minerals [9,10]. It has an advantage over many experimental

G. Rutkai et al. / Fluid Phase Equilibria 409 (2016) 434e438

methods that it offers immediate insight on the molecular level and


it is not limited by extreme conditions. Thanks to the rapid development of high performance computing, it is already possible to
simulate the initial phase of the delamination/exfoliation of
nanolayers with a realistic kaolinite layer size in real time [11].
Nevertheless, massively parallel supercomputers are still not
available for everyday use, not even in academic research. In terms
of evaluating the stability of a given kaolinite-solvent system, the
general intercalation phenomena of clay minerals can be targeted
with various simulation scenarios that involve considerably less
computational effort. Still, careful planning or preceding experience is often necessary to be fully aware of the pros and cons of the
selected approaches.
Generally, the most simple and convenient choice is to perceive
the kaolinite-solvent mixture as a result of a special phase equilibrium between the bulk phase of the solvent uid (gas or liquid
solvent) and that of the mixture. The obvious simulation setup for
this scenario consists of two simulation boxes: one box represents
the homogeneous bulk phase of the solvent molecules; the other
contains both the solvent and an arbitrary segment of the crystallographic structure of kaolinite (which is essentially molecular
layers of kaolinite stacked upon each other with solvent molecules
between these layers). This conguration is repeated in all three
spatial directions through periodic boundary conditions. To ensure
thermodynamic equilibrium, the pressure p and the temperature T
of the two systems are set to identical values. Whether the equality
of the chemical potential m of the solvent molecules with respect to
the two phases is ensured in separate constant chemical potential
ensemble simulations or using a Gibbs ensemble setup [12,13], it is
irrelevant for the current purpose. In any case, to capture the
swelling nature of the clay, it is sensible to model the phase containing it using a mSolventNClaypT ensemble. In such simulation, the
volume of the system V that corresponds to the external pressure p
is restricted to change only along the axis normal to the layers. The
chemical potential of the solvent molecules mSolvent, the pressure p,
and the temperature T are held constant and match those of the
bulk phase while the number of particles N for the clay mineral is
also constant. In a single classical molecular simulation, an
ensemble is guaranteed to sample the equilibrium state (or an
equilibrium state if there are more than one) of the model system.
In intercalation experiments, the kaolinite-solvent mixture is
traditionally characterized by the basal spacing d (distance between two consecutive layers) and the solvent content between
layers (often referred as loading). These properties can be conveniently obtained as simple ensemble averages using the mSolventNClaypT ensemble. This method was successfully applied to predict
the basal spacing and loading for variety of intercalation systems
for small polar solvent molecules and the obtained results showed
promising agreement with available experimental data [14]. Unfortunately, the efciency of a constant chemical potential
ensemble calculation is limited by the size of the solvent molecules,
because the number of successful particle insertions/deletions,
required for trustworthy statistical results, decreases drastically
with the increasing molecular size of the solvent in spatially
conned systems for a given runtime. Another drawback of the
mSolventNClaypT ensemble approach is that, although it nds equilibrium state(s), it cannot give further information around the
minimum about the shape of the energy surface that is dened by
other solvent loadings and corresponding basal spacings.
A simple method to obtain a general view with respect to stability of different compositions is to perform a series of NpT (NSolventNClaypT) ensemble simulations with increasing solvent content
NSolvent and plot NSolvent in function of the basal spacing d that is
obtained as an ensemble average (cf. Fig. 1 top). A basal spacing
range represents a stable state if d does not change substantially

435

Fig. 1. Molecular simulation results for the chemical potential m of methanol (MeOH)
calculated by thermodynamic integration and for basal spacing d in the kaolinitemethanol system. mMeOH denotes the methanol content in 1 kg of kaolinite. White
squares: series of NpT ensemble simulations with setup (I). Red triangles: mSolventNClaypT simulations with setup (I). Circles and crosses: series of NpT ensemble simulations with setup (II). Circles represent results, with setup (II), that do not consider
contributions from internal degrees of freedom as opposed to crosses. Grey circles
represent the approximation mconf z vHconf/vNMeOH. Both vHconf/vNMeOH and m0 vm/
vNMeOH were obtained by numerical differentiation. Note that the exclusively temperature dependent ideal contribution mid(T) (which includes internal degrees of
freedom) cancels out in this derivation step. The detailed description of setups (I) and
(II) can be found in the text. In the ideal contribution RTln(r/rreference), r NMeOH/
V nm3 and rreference was 1 nm3. For the black circles, dashed areas indicate the regions in which stable states are located based on the individual subplots; their joint
segments help to narrow down these regions. (For interpretation of the references to
colour in this gure legend, the reader is referred to the web version of this article).

with solvent content. Due to its simplicity, this approach is a


straightforward choice of simulation works [15e18]. However, in
order to be able to properly comment on the stability of states at
constant N, p and T, the corresponding Gibbs free energy prole or
equivalent information is required. The local minima of the system's Gibbs free energy G are indicated by the rst and second
derivative tests. Namely, mSolvent vG/vNSolvent 0 (for which
mSolvent must be smaller than zero left and larger than zero right
from the local minimum) and m0 Solvent vmSolvent/vNSolvent > 0
around the minimum while NClay, p, and T are constant [19]. In
molecular simulations, the residual chemical potential can be
calculated, in principle, directly by Widom's particle insertion
method [20]. However, the method requires sufcient amount of
particle insertions without overlaps, which practically cannot be
achieved in spatially conned systems such as the one containing

436

G. Rutkai et al. / Fluid Phase Equilibria 409 (2016) 434e438

the kaolinite-solvent phase. One immediate solution is to approximate the chemical potential by

vH
vNSolvent

p;T; NClay

vG

vNSolvent


T
p;T; NClay

vmSolvent
mSolvent  T
vT

vS
vNSolvent

p; NClay


p;T; NClay

zmSolvent
(1)

(S is the entropy) using the system's enthalpy H [16,17], which is


almost always a standard output for any classical simulations.
Nevertheless, independently of which approach was chosen
earlier, the main conclusion for each and every solvent molecule in
previous works was that more than one stable kaolinite-solvent
mixture can be detected in simulations [14,16,17,21] for a given
mSolvent, p, and T. These stable mixtures have quite different basal
spacings and always one stable mixture corresponds to a single
layered arrangement of solvent molecules between the kaolinite
layers, while the other corresponds to a double layered structure.
The existence of the second stable mixture was never before
conrmed in experiments to our knowledge until recently [17,22].
Here, we aim to extend the arsenal of previously applied
methods [14,16,17,21] that did not use true chemical potential
calculation in the kaolinite-solvent phase. For the determination of
the chemical potential, we apply the thermodynamic integration
method [13]. The idea behind calculating m using thermodynamic
integration is to avoid insertion of particle i in system A and
perform it in system B for which it is not problematic or practically
not required at all:

+
vUi l
dl
vl

3. Results
Intercalation experiments involving clay minerals are mainly
performed at room temperature and atmospheric pressure.
Therefore, the pressure in our calculations was set to p 1 bar and
the temperature to T 298 K. Figs. 1 and 2 summarize the results
obtained by the previously described methods. The mMeOH vs.
d plots indicate two stable d regions (narrow ranges, identied
where the loading exhibits signicant changes) and the two setups
using considerably different models predict the equilibrium basal
spacings with relatively good agreement: 1.09 0.02 nm and
1.12 0.01 nm with system (I) and (II), respectively, for the rst
stable state. The calculations yielded 1.41 0.09 nm and
1.46 0.08 nm with system (I) and (II), respectively, for the second
stable state. For system (I), the mSolventNClaypT ensemble simulations
pinpoint the equilibrium states at 1.091 0.004 nm and
1.444 0.009 nm. The experimental value for the rst stable state is
1.11e1.12 nm [4,7,24] (although it is generally accepted in experiments that methanol can even react with the eAleOH groups of the
kaolinite surface forming eAleOeCH3 groups and water as a
byproduct, the basal spacing of this methanol-wet state is
unambiguously established). Once the regions in which mconf shows
increasing tendency which means that m0 conf > 0 are identied, it
can be concluded that the mconf vs. d and m0 conf vs. d plots are fully
consistent with the previously described stable states. Even the
simple approximations mconf z vH/vNMeOH and m0 conf z v2H/

A;l
Z max *

mi;A  mi;B
B;lmax

(2)

The bracket in this equation denotes NVT or NpT ensemble


average and Ui is the potential energy of particle i that must be a
part of the system the same way any other particle is. The only
difference between i and regular particles is that the interaction
energy Ui(l) is scaled between states A and B with respect to a
parameter l. As long as vUi(l)/vl can be calculated analytically and
sampled during simulation, the actual way of scaling Ui(l) with l is
completely arbitrary since m is a state variable. Finally, Eq. (2) is
evaluated by numerical integration. Assuming that mi,B is practically
zero or at least can be successfully calculated by Widom's particle
insertion method [20] at the end point B, mi,A yields the residual part
of the total chemical potential mtotal(T, r) mres(T, r) mid(T, r). The
ideal part mid(T, r) mid(T) RTln(r/rreference) is dened as the value
of mtotal(T, r) when no intermolecular interactions are at work. mid(T)
has non-trivial but exclusive temperature dependence and it is a
sum of contributions due to translation, rotation, and internal degrees of freedom. The sum mres(T, r) RTln(r/rreference) is often
referred as the congurational chemical potential mconf(T, r).
Two different system setups are employed in this work in terms
of interaction models, systems size, and numerical evaluation.
Setup (I) considers a smaller system sampled with Monte Carlo
(MC) calculations using potential models without internal degrees
of freedom (rigid). For setup (II), the system is six times larger, using
exible models in molecular dynamics (MD) calculations. A
detailed description can be found in the Appendix. Extension of the
model system, by considering other factors like chemically modied kaolinite surface, is investigated in a work that has just been
published [23].

Fig. 2. Description and notations are identical that of Fig. 1.

G. Rutkai et al. / Fluid Phase Equilibria 409 (2016) 434e438

vN2MeOH can capture the general trend. This suggests that the
predominant contributions in this system with such small guest
molecules are the energetic ones and the entropic term in Eq. (1)
can be neglected in certain cases. In Fig. 2, the mconf vs. mMeOH
and m0 conf vs. mMeOH plots imply that the minima of Gibbs free
energy of the systems are broad and have indeed a low gradient in
function of the loading (despite the generally larger uncertainties of
the numerical derivatives, especially at higher loadings). This
observation in simulation is again in accordance with the current
practice that a basal spacing range represents stability if d does not
change substantially with solvent content. It can also explain why
independent laboratory measurements for the same solvent in
kaolinite tend to nd considerably different loadings (even up to
200% deviation) [25e27]. Nevertheless, it is also true that it is
difcult to separate the actual interlayer loading of the kaolinite
particles and the solvent content adsorbed on surface of the grains
in experiments. Despite the considerably different simulation
setups, qualitatively similar conclusions can be derived from the
analyses with the two (a rigid and a fully exible) interaction
models. The existence of two stable states can be clearly observed
in the simulations, which is in agreement with previous ndings for
other solvents [14,16,17,21]. Thus, the present work offers a
consistent set of test scenarios to determine the stability of intercalate complexes that serves as a solid ground for further investigation purposes.
Acknowledgement
We gratefully acknowledge the nancial support of the Hungarian State and the European Union under TAMOP-4.2.2.B-15/1/
KONV-2015-0004 and the computational support of the Paderborn Center for Parallel Computing (PC2) for providing access to the
OCuLUS cluster.
Appendix. Simulation details
The composition of the kaolinite unit cell which has space group
symmetry C1 is Al2Si2O5(OH)4 and the lattice parameters are
a 0.5154 nm, b 0.8942 nm, c 0.7391 nm, a 91.93 ,
b 105.05 , and g 89.80 [28]. We constructed two systems
setups that consist of clay layers stacked upon each other forming
interlayer regions with the periodic boundary conditions applied in
all three spatial directions. Setup (I) is built by multiplying a double
unit cell 4  2  2 times resulting in two clay layers with 272 atoms
(32 Al, 32 Si, 144 O, and 64 H) per kaolinite layers with two interlayer regions. Setup (II) is six times larger and it has a 6  4  4
double cell arrangement forming four clay layers with 816 atoms
(96 Al, 96 Si, 432 O, and 192 H) per layers with four interlayer
regions.
MC simulations of setup (I) were carried out with our own code.
The total length of the simulations was 107 MC moves for the
equilibration plus 108 MC moves for the production period. One MC
move was dened as either particle translation, particle insertions/
deletions (70e80% of the total number of MC moves for the two
mSolventNClaypT ensemble simulations), or volume change attempts
(1.0e3.0% of the total number of MC moves). The sampling efciency of particle insertions/deletions was enhanced using the
congurational-bias technique [29]. Using the assumption that the
models of the solvent molecules are rigid, this technique reduces to
the selection of trial orientations before insertion/deletion.
Thermodynamic integration was carried out applying nonlinear scaling Ui(l) l12$Ui (0.01  l  1) with an adaptive sampling technique [30]. This technique allows the sampling of the
entire range of 0.01  l  1 in a single MC simulation run with
arbitrary resolution by treating particle i as uctuating particle.

437

Besides the standard MC moves, the simulation includes l change


(was here set to 2.5% of the total number of MC moves) that is
controlled by a proper MC acceptance criterion ensuring visits in
the entire lmin  l  lmax region. Chemical potential results were
veried using a fourth dimension based thermodynamic integration [31] that also uses the same adaptive sampling framework. In
this 4D method, l serves as an extra spatial dimension for the
uctuating particle. For this scenario, l can change within the
boundaries lmin m  l  lmax 0, where m is some upper limiting
value for the extra spatial dimension. The chemical potentials of the
methanol bulk phase required by the mSolventNClaypT ensemble
simulations were also calculated by the 4D method. These results
were veried using an independent code [32].
For the rigid kaolinite framework, the model of Fang et al. [15]
was used that has proven to yield consistently good results for
various solvents in earlier works [14,16,17]. Methanol was represented by an OPLS type model [33].
MD simulations of setup (II) were performed using the GROMACS [34,35] program suite. For equilibrating the system, the leapfrog integrator was used and the integral time step was set to 1 fs.
During the initial 10 ns long equilibration period, the temperature
and pressure were scaled using velocity rescaling with a modied
version of the Berendsen thermostat [36] and Berendsen barostat
[37]. The pressure coupling was semi-isotropic: isotropic in the xand y-directions, but different in the z-direction, which is perpendicular to the kaolinite layers. Right after this another equilibration
period was performed for 10 ns, but this time with NoseeHoover
thermostat and ParrinelloeRahman barostat [38,39]. These are
known to provide correct NpT ensemble uctuations in simulations.
For obtaining the chemical potential of methanol, the GROMACS
built-in thermodynamic integration method was used that employs an essentially linear scaling Ui(l) l$Ui in Eq. (2). In cases
where intramolecular contributions were considered in the calculations, the bond, angle, or dihedral potentials were modied
through scaling the corresponding force constants. Flat-bottomed
position restraints were applied in the direction perpendicular to
the kaolinite layers to create excluded volumes in the middle of the
kaolinite layers for the methanol test particle. Thus, the sampling
was restricted to the interlayer space and the outer atoms of the
kaolinite layers. To obtain vUi(l)/vl, individual simulations were
performed for each l value (from 0 < l  1 with an increment of 0.1)
with the leap-frog stochastic dynamics integrator with the same
parameters described above (1 fs time-step, 10 ns time, ParrinelloeRahman barostat, NoseeHoover thermostat). To obtain
adequate results, statistical data collected in the rst third of the
simulation were excluded from the averaging process.
The simulations employed the fully exible all-atom force eld,
CHARMM [40] and a recently proposed thermodynamically
consistent force eld, INTERFACE [41]. The parameters for kaolinite
were taken from the INTERFACE force eld (v1.3). This package
operates as an extension of common harmonic force elds such as
CHARMM for inorganic compounds and it is the only available force
eld to our knowledge that describes intra- and intermolecular
interactions and contains dedicated parameters for kaolinite. The
CHARMM27 force eld variant implemented in GROMACS [42] was
used for methanol. The structure and the atomic charges of the
methanol model were adapted from the Topology for the Charmm
General Force Field v. 2b7 for Small Molecule Drug Design. The
periodic electrostatic interactions were calculated by the particle
mesh Ewald (PME) [43] algorithm with 1.2 nm cutoff. The van der
Waals nonbonded interactions were truncated at 1.2 nm and were
treated with the standard 12-6 LennardeJones potential using the
LorentzeBerthelot combination rule. Only the H-bond vibrations
(which are generally beyond the classical limit in these conditions)
were constrained with the LINCS algorithm [44].

438

G. Rutkai et al. / Fluid Phase Equilibria 409 (2016) 434e438

References
[1] F. Annabi-Bergaya, Microporous Mesoporous Mater. 107 (2008) 141e148.
http://dx.doi.org/10.1016/j.micromeso.2007.05.064.
ve, L. Le Pluart, P.[2] N. Sibold, C. Dufour, F. Gourbilleau, M.-N. Metzner, C. Lagre
J. Madec, T.-N. Pham, Appl. Clay Sci. 38 (2007) 130e138. http://dx.doi.org/10.
1016/j.clay.2007.02.006.
[3] Y. Kuroda, K. Ito, K. Itabashi, K. Kuroda, Langmuir 27 (2011) 2028e2035.
http://dx.doi.org/10.1021/la1047134.
[4] P. Yuan, D. Tan, F. Annabi-Bergaya, W. Yan, D. Liu, Z. Liu, Appl. Clay Sci. 83e84
(2013) 68e76. http://dx.doi.org/10.1016/j.clay.2013.08.027.
[5] H. Cheng, Q. Liu, J. Yang, S. Ma, R.L. Frost, Thermochim. Acta 545 (2012) 1e13.
http://dx.doi.org/10.1016/j.tca.2012.04.005.
[6] R.L. Ledoux, J.L. White, J. Colloid Interface Sci. 21 (1966) 127e152. http://dx.
doi.org/10.1016/0095-8522(66)90029-8.
[7] Y. Komori, Y. Sugahara, K. Kuroda, J. Mater. Res. 13 (1998) 930e934. http://dx.
doi.org/10.1557/JMR.1998.0128.
[8] J.J. Tunney, C. Detellier, J. Mater. Chem. 6 (1996) 1679e1685. http://dx.doi.org/
10.1039/JM9960601679.
[9] R.M. Shroll, D.E. Smith, J. Chem. Phys. 111 (1999) 9025. http://dx.doi.org/10.
1063/1.480245.
ez, K. Van Workum, L. de Pablo, J.J. de Pablo, J. Chem. Phys. 114
[10] M. Ch
avez-Pa
(2001) 1405. http://dx.doi.org/10.1063/1.1322639.
, G. Rutkai, J. Vrabec, T. Kristo
f, J. Chem. Phys. 141 (091102) (2014).
[11] Z. Hato
http://dx.doi.org/10.1063/1.4894756.
[12] A.Z. Panagiotopoulos, Mol. Phys. 61 (1987) 813e826. http://dx.doi.org/10.
1080/00268978700101491.
[13] D. Frenkel, B. Smit, Understanding Molecular Simulation: from Algorithms to
Applications, Academic Press, Elsevier, San Diego, CA, 2002.
f, Chem. Phys. Lett. 462 (2008) 269e274. http://dx.doi.org/
[14] G. Rutkai, T. Kristo
10.1016/j.cplett.2008.07.092.
[15] Q. Fang, S. Huang, W. Wang, Chem. Phys. Lett. 411 (2005) 233e237. http://dx.
doi.org/10.1016/j.cplett.2005.06.052.
 Mako
, T. Kristo
f, J. Colloid Interface Sci. 334 (2009) 65e69. http://
[16] G. Rutkai, E.
dx.doi.org/10.1016/j.jcis.2009.03.022.

, G. Rutkai, T. Kristo
f, J. Colloid Interface Sci. 349 (2010) 442e445.
[17] E. Mako
http://dx.doi.org/10.1016/j.jcis.2010.05.021.
[18] M. Szczerba, Z. Kapyta, A. Kalinichev, Appl. Clay Sci. 91 (2014) 87e91. http://
dx.doi.org/10.1016/j.clay.2014.02.014.
[19] J.S. Rowlinson, F.L. Swinton, Liquids and Liquid Mixtures, third ed., Butterworths, 1982.
[20] B. Widom, J. Chem. Phys. 39 (1963) 2808. http://dx.doi.org/10.1063/1.
1734110.
 Mako
 , E.
, T. Kristo
 f, J. Mol. Model. 20 (2014) 2140. http://dx.doi.org/10.
[21] Z. Hato
1007/s00894-014-2140-9.

, A. Kov
, B. Zsirka, T. Kristo
f, J. Colloid Interface Sci. 431
[22] E. Mako
acs, Z. Hato
(2014) 125e131. http://dx.doi.org/10.1016/j.jcis.2014.06.006.

 Mako
 , A. Kova
cs, Z. Hato
, T. Kristo
f, Appl. Surf. Sci. 357 (2015) 626e634.
[23] E.
http://dx.doi.org/10.1016/j.apsusc.2015.09.081.
[24] J. Matusik, E. Scholtzova, D. Tunega, Clays Clay Miner. 60 (2012) 227e239.
http://dx.doi.org/10.1346/CCMN.2012.0600301.
[25] J.J. Tunney, C. Detellier, Chem. Mater. 5 (1993) 747e748. http://dx.doi.org/10.
1021/cm00030a002.
[26] S. Hayashi, J. Phys. Chem. 99 (1995) 7120e7129. http://dx.doi.org/10.1021/
j100018a053.
[27] J.E. Gardolinski, P. Peralta-Zamora, F. Wypych, B. Smit, J. Colloid Interface Sci.
211 (1999) 137e141. http://dx.doi.org/10.1006/jcis.1998.5982.
[28] D.L. Bish, Clays Clay Miner. 41 (1993) 738e744. http://dx.doi.org/10.1346/
CCMN.1993.0410613.
[29] J.I. Siepmann, D. Frenkel, Mol. Phys. 75 (1992) 59e70. http://dx.doi.org/10.
1080/00268979200100061.
[30] M. Fasnacht, R.H. Swendsen, J.M. Rosenberg, Phys. Rev. E 69 (056704) (2004).
http://dx.doi.org/10.1103/PhysRevE.69.056704.
f, G. Rutkai, Chem. Phys. Lett. 445 (2007) 74e78. http://dx.doi.org/10.
[31] T. Kristo
1016/j.cplett.2007.07.054.
[32] C.W. Glass, S. Reiser, G. Rutkai, S. Deublein, A. Koster, G. Guevara-Carrion,
A. Wafai, M. Horsch, M. Bernreuther, T. Windmann, H. Hasse, J. Vrabec,
Comput. Phys. Commun. 185 (2014) 3302e3306. http://dx.doi.org/10.1016/j.
cpc.2014.07.012.
[33] E. van Leeuwen, B. Smit, J. Phys. Chem. 99 (1995) 1831e1833. http://dx.doi.
org/10.1021/j100007a006.
[34] H.J.C. Berendsen, D. van der Spoel, R. van Drunen, Comput. Phys. Commun. 91
(1995) 43e56. http://dx.doi.org/10.1016/0010-4655(95)00042-E.
[35] B. Hess, C. Kutzner, D. van der Spoel, E. Lindahl, J. Chem. Theory Comput. 4
(2008) 435e447. http://dx.doi.org/10.1021/ct700301q.
[36] G. Bussi, D. Donadio, M. Parrinello, J. Chem. Phys. 126 (2007) 014101. http://
dx.doi.org/10.1063/1.2408420.
[37] H.J.C. Berendsen, J.P.M. Postma, A. DiNola, J.R. Haak, J. Chem. Phys. 81 (1984)
3684. http://dx.doi.org/10.1063/1.448118.
[38] M. Parrinello, A. Rahman, J. Appl. Phys. 52 (1981) 7182e7190. http://dx.doi.
org/10.1063/1.328693.
, M.L. Klein, Mol. Phys. 50 (1983) 1055e1076. http://dx.doi.org/10.
[39] S. Nose
1080/00268978300102851.
[40] A.D. MacKerell Jr., Atomistic models and force elds, in: O.M. Becker,
A.D. MacKerell Jr., B. Roux, M. Watanabe (Eds.), Computational Biochemistry
and Biophysics, Marcel Dekker, Inc, New York, 2001, pp. 7e38.
[41] H. Heinz, T.J. Lin, R.K. Mishra, F.S. Emami, Langmuir 29 (2013) 1754e1765.
http://dx.doi.org/10.1021/la3038846.
[42] P. Bjelkmar, P. Larsson, M.A. Cuendet, B. Bess, E. Lindahl, J. Chem. Theory
Comput. 6 (2010) 459e466. http://dx.doi.org/10.1021/ct900549r.
[43] T. Darden, D. York, L. Pedersen, J. Chem. Phys. 98 (1993) 10089. http://dx.doi.
org/10.1063/1.464397.
[44] B. Hess, J. Chem. Theory Comput. 4 (2008) 116e122. http://dx.doi.org/10.
1021/ct700200b.

Anda mungkin juga menyukai