ArticlePDF Available

Molecular Dynamics Simulations of Adsorption of Organic Compounds at the Clay Mineral/Aqueous Solution Interface

Authors:

Abstract and Figures

Computational studies of the sorption of organic compounds at clay mineral surfaces are described. Molecular dynamics simulations were performed with a recently developed empirical force field for dioctahedral clays. The studies allow the identification of three general mechanisms of adsorption. In the absence of water, organic compounds adsorb to mineral surfaces in such a way that contact area is maximized. In the presence of a sufficient amount of water, some molecules can adsorb via a single functional group, while the bulk of the molecular structure is immersed in the aqueous phase. When many water molecules are present, they form a structured layer, excluding organic adsorbates from the mineral basal plane. A detailed description is given of the characteristic structure found for water layers in the interlayer space of clays. The calculated trends are reasonable, but we also expect that current dynamics simulations may overestimate the extent of the structuring of water because of the absence of polarization terms in the available empirical force fields.
Content may be subject to copyright.
P1: GXB
Structural Chemistry (STUC) pp709-stuc-456242 January 14, 2003 17:16 Style file version Nov. 07, 2000
Structural Chemistry, Vol. 14, No. 2, April 2003 (C°2003)
Molecular Dynamics Simulations of Adsorption of Organic
Compounds at the Clay Mineral/Aqueous Solution Interface1
Ching-Hsing Yu,2Susan Q. Newton,2Mya A. Norman,2
Lothar Sch¨afer,2and David M. Miller3
Received November 29, 2001; revised January 4, 2002; accepted January 14, 2002
Computational studies of the sorption of organic compounds at clay mineral surfaces are described.
Molecular dynamics simulations were performed with a recently developed empirical force field for
dioctahedral clays. The studies allow the identification of three general mechanisms of adsorption. In
the absence of water, organic compounds adsorb to mineral surfaces in such a way that contact area is
maximized. In the presence of a sufficient amount of water, some molecules can adsorb via a single
functional group, while the bulk of the molecular structure is immersed in the aqueous phase. When
many water molecules are present, they form a structured layer, excluding organic adsorbates from
the mineral basal plane. A detailed description is given of the characteristic structure found for water
layers in the interlayer space of clays. The calculated trends are reasonable, but we also expect that
current dynamics simulations may overestimate the extent of the structuring of water because of the
absence of polarization terms in the available empirical force fields.
KEY WORDS: Molecular dynamics simulations; clay minerals; adsorption on clay surfaces; interactions of
organic compounds with clays; soil pollutants.
INTRODUCTION
In 1808 Gay-Lussac predicted [1] that “we are per-
haps not far removed from the time when we shall be able
to submit the bulk of chemical phenomena to calculation.
The great French chemist could not have anticipated the
computational revolutions that occurred in the mean time
but, nearly 200 years later, we have come very close to
realizing his vision.
In the current paper we will describe some fairly
novel applications of molecular dynamics simulations in
the area of soil chemistry. In soil chemistry, detailed de-
scriptions of the interactions of molecules with mineral
surfaces are of great general interest. During the past
decades soil chemists have increasingly employed spec-
troscopic techniques for this purpose. In support of these
1Dedicated to the memory of Barbara Starck.
2Department of Chemistry University of Arkansas, Fayetteville,
Arkansas 72701.
3Department of Crop, Soil, and Environmental Science, University of
Arkansas, Fayetteville, Arkansas 72701; email: dmmiller @uark.edu
efforts we are exploring the possibilities of atomic scale
computer simulations as a means to help interpret the em-
pirical data.
Withintheframework of aquasi-classical formalism,
molecular dynamics (MD) simulations are an effective
tool for providing information on the properties of large
molecularsystems. Simulationsof thiskind typicallygen-
erate collections of molecular configurations, which can
serve as approximate statistical ensembles. MD simula-
tions are characterized by the fact that the equilibrium
state of a system is not approached in a static way, but
all the atoms and molecules are allowed to move without
any restraints other than those imposed by the force field
employed and the state variables, such as temperature and
pressure. Standard procedures from statistical thermody-
namicscan beusedto calculateensembleaverages,which,
in turn, can be related to the macroscopic properties of
interest.
In the current paper we will describe some general
trends emerging from ongoing MD simulations of the
sorption of organic compounds, specifically pesticides, to
mineralsurfaces. Specialattention will begiven to therole
of water in the adsorption process.
175
1040-0400/03/0400-0175/0 C
°2003 Plenum Publishing Corporation
P1: GXB
Structural Chemistry (STUC) pp709-stuc-456242 January 14, 2003 17:16 Style file version Nov. 07, 2000
176 Yu, Newton, Norman, Sch¨afer, and Miller
COMPUTATIONAL PROCEDURES
Aluminosilicate clays typically consist of sheets of
aluminum oxide, in which the metal is octahedrally co-
ordinated, bonded to sheets of tetrahedral silicon oxide.
The sheets can be combined in various ways. Pure alu-
minum oxide sheets exist in gibbsite, whose crystal struc-
ture is given in Ref. [2]. Layers of aluminum oxide sheets
bonded to silicon oxide sheets exist in the 1:1 type clay
kaolinite, whose crystal structure is given in Ref. [3]. The
2:1 clay pyrophyllite contains layers of aluminum oxide
sandwiched between two silicon oxide sheets; its crystal
structure is given in Ref. [4].
In order to perform MD simulations of dioctahedral
clays, we have developed [5] potential energy functions
that can be used in combination with the force field sup-
plied with the former MSI/Insight/Discover (now Accel-
rys) software suite [6]. Among various options, the latter
contains the cff91 parameters [7] for standard organic and
inorganic compounds, augmented with parameters for sil-
icates and zeolites developed by Hill and Sauer. [8,9] For
octahedrally coordinated aluminum, it was necessary to
derive[5]additional potentialparameters thatmakeit pos-
sible to perform computations of phyllosilicates. Specifi-
cally,itwasnecessary toderiveanangle-bending potential
for octahedral O Al O angles, which displays dual min-
ima at 90 and 180. Furthermore, since the force field by
Hill and Sauer [8,9] was based in part on the results of
ab initio calculations that did not include electron correla-
tion,a new setofnonbonded parameters andpartialatomic
charges was derived from electron-correlated ab initio ge-
ometry optimizations of molecular structures represent-
ing fragments of phyllosilicates. The resulting potential
parameters [5] were further refined with the help of the
X-ray crystal structures of some oxides and phyllosilicate
minerals.
Our MD simulations are characterized by the fact
that all the atoms of a given system, including those in the
mineral lattice, are allowed to move subject only to the
constraints of the force field. This is in contrast to those
previous procedures, which were only partially dynamic
in that only some of the atoms of a system were allowed
to move, while others were held rigid.
General information on the procedural details in-
volved in performing molecular dynamics simulations
can be found in a number of standard texts, such as the
one by Allen and Tildesley [10]. Some characteristic re-
sults of previous modeling efforts by others can be found
for simple oxides [11–14], mixed oxides [15–18], mica
[12,19–21], and smectite clay minerals [22]. References
[23–55]provide asurveyof therich spectrumof topics ad-
dressed and techniques applied, in computer simulations
of minerals. The excellent book by Cygan and Kubicky
[56] illustrate the current importance of this field.
Detailed descriptions of specific technical aspects of
ourcalculations havebeen given [57–60] andonly some of
the main points shall be repeated here. In all our studies,
we construct mineral surfaces from unit cells of repre-
sentative systems obtained from X-ray crystal structures.
For example, the experimental coordinates determined by
Bish [3] for the 1:1 clay mineral kaolinite can be used to
construct a surface for simulations by fusing several unit
cells together to form a supercell. A supercell with com-
position Al32Si32O80(OH)64, for example, offers a nearly
rectangular repeat unit base of 20.61 by 17.88 ˚
Ainthe
crystallographic ab-plane [57].
Unlike smectites, kaolinite does not typically swell
along the c-axis upon hydration. The layers can never-
theless be separated in the simulations, increasing the c
axis from 7.4 ˚
A, the crystal spacing [3], to some 20.0
˚
A or larger, in order to create an interlayer space of de-
sireddimension,where thebehaviorofwaterand adsorbed
molecules can be modeled. Kaolinite is an interesting
model mineral, since it presents two very different types
of surfaces to aqueous solutions. The artificial expansion
of the interlayer space creates a pore, which provides both
typesof basal externalsurfacefoundon akaolinitemineral
grain. Slit pores of this type are found in kaolinite books
[61] and perhaps at interfaces between silicate grains and
aluminum oxide coatings.
Systems with an artificially expanded interlayer
space are modeled under NVT (constant mass, volume
and temperature) conditions, if the intention is to keep
the d(001) spacing constant. In order to equilibrate an
expanded system, simulations are subjected to NPT (con-
stant mass, pressure, and temperature) conditions. During
that process, the separated layers spontaneously anneal,
restoring the equilibrium interlayer spacing characteristic
of a given system.
In other MD simulations [59], we have used the crys-
tallographic structure [4] of pyrophyllite, an uncharged,
2:1, dioctahedral phyllosilicate, to construct a model sys-
tem for studying adsorption. For example, fusing six unit
cells of pyrophyllite will yield an Al24Si48O120(OH)24 su-
percell of a neutral, idealized 2:1 clay. Pyrophyllite is well
suited for study because it has the same structure as the
smectites, but it is a neutral clay and the interlayer space
is thus devoid of any hydrated counter ions. When Si ions
in the tetrahedral sheet of a pyrophyllite supercell are iso-
morphically substituted with Al, an idealized beidellite
system results with a cation-exchange capacity dependent
on the chosen extent of substitution. Alternatively, model
montmorillonite systems can be constructed by isomor-
phic substitutions of Al in the octahedral sheet by Mg.
P1: GXB
Structural Chemistry (STUC) pp709-stuc-456242 January 14, 2003 17:16 Style file version Nov. 07, 2000
Molecular Dynamics Simulations of Adsorption of Organic Compounds 177
The mineral surfaces constructed in this way are usu-
ally hydrated at various levels, producing various layers
of water, as needed. Subsequently, the adsorbates to be
studied are added to the interlayer space or pore surface
and subjected to MD simulations for time periods of, typ-
ically, several hundred picoseconds, using a time step of
0.5 fs.
RESULTS AND DISCUSSION
Three Mechanisms of Adsorption of Organic
Compounds on Mineral Surfaces
The structures of two pesticides, 2-chloro-4-
ethylamino-6-isopropylamino-S-triazine, commonly
called atrazine (ATZ), and of 2-methyl-4,6-dinitro
phenol, commonly called DNOC, are shown in Fig. 1. We
are currently studying the adsorption of both systems on
a model montmorillonite with a cation-exchange capacity
of105 meq/100g, usingK+ascounter ions.Other counter
ions are also considered because they affect the transport
properties of pesticides on soils [62]. The behavior of
DNOC and ATZ on clays follows trends generally found
for the sorption of organic compounds [57–60].
When organic compounds like ATZ and DNOC are
placed with random orientations in the clay interlayer
space, as shown for DNOC in Fig. 2, they will move spon-
taneously during MD simulations to the mineral surface
and adsorb coplanar with the mineral basal plane (Fig. 2).
During this process, the counter ions will perform a simi-
lar movement, sometimes forming ion bridges, as seen for
Fig. 1. The molecular structures of 2-chloro-4-ethylamino-6-isopropylamino-S-triazine, commonly called atrazine (ATZ, on
the left), and of 2-methyl-4,6-dinitro phenol, commonly called DNOC (on the right).
DNOC in Fig. 2. Interestingly, no ion bridging was found
for ATZ.
The tendency to maximize contact area is a general
characteristic for organic compounds on dry mineral sur-
faces, when no water is present. In the case of peptides
and proteins [60] it is the cause of significant denaturing,
i.e., large changes in φ,ψ-torsional angles [60]. Details
of how adsorbed species in this state interact with silox-
ane surfaces are shown in Fig. 3. In a characteristic way,
adsorbing species tend to point a functional group to the
inside of a hexagonal siloxane cavity, but not to the ex-
act center. Off-center adsorption is particularly clear for
single ions (Fig. 3).
In the presence of water, the adsorption mechanism
can change significantly because water can displace or-
ganic compounds from surface sites, and because polar
compounds show an affinity of their own for aqueous so-
lutions. Basically, two different processes are possible in
the presence of water.
When the amount of water in the interlayer space
is sufficiently high for forming several water layers (see
below), organic compounds can attach to a mineral sur-
face via a single functional group, while its bulk is im-
mersed in the aqueous phase. Such a case is shown in
Fig. 4, in which DNOC is seen to sorb via one of its NO2
groups, while the main part of its body remains immersed
in the aqueous phase. Similarly, ATZ was found to inter-
act with a mineral surface by penetrating a water layer
with its C-Cl bond. The same mechanism was previously
obtained for trichloroethene [57]. In the single-group ad-
sorption mode, the adsorbates are typically mobile and
P1: GXB
Structural Chemistry (STUC) pp709-stuc-456242 January 14, 2003 17:16 Style file version Nov. 07, 2000
178 Yu, Newton, Norman, Sch¨afer, and Miller
Fig. 2. When DNOC molecules are placed with random orientations in the interlayer space of a dry montmorillonite (top), they will move
spontaneously during MD simulations to the mineral surface plane and adsorb coplanar with the mineral basal plane (bottom).
P1: GXB
Structural Chemistry (STUC) pp709-stuc-456242 January 14, 2003 17:16 Style file version Nov. 07, 2000
Molecular Dynamics Simulations of Adsorption of Organic Compounds 179
Fig. 3. The positions of adsorbed ATZ (left), DNOC (right), and counter ions (K+, isolated spheres in both figures) relative to the hexagonal siloxane
cavity of montmorillonite.
Fig. 4. In the presence of several layers of water, an organic molecule can adsorb at a mineral surface via a single group, while
the bulk of its structure is immersed in the aqueous phase, as shown here for DNOC on montmorillonite (snapshot after 250 ps
of MD simulations). For graphic clarity, water molecules are rendered in the line-drawing mode, while the ball and stick mode
was chosen for DNOC and the mineral lattice.
P1: GXB
Structural Chemistry (STUC) pp709-stuc-456242 January 14, 2003 17:16 Style file version Nov. 07, 2000
180 Yu, Newton, Norman, Sch¨afer, and Miller
Fig. 5. Histogram constructed from 5000 time steps of MD simulations of an equilibrated ensemble of
DNOC, water, and K+ions on montmorillonite. The frequencies (vertical axis) are shown of distances (˚
A,
horizontal axis) between DNOC–oxygen atoms and K+ions.
Fig. 6. Rubredoxin with six DNOC molecules and 2070 water molecules in an artificially enlarged interlayer space of
pyrophyllite. DNOC can interact both with the mineral surface and the protein. After 150 ps of NPT dynamics, the protein
retained its globular shape and remained inside the aqueous solution phase. For reasons of graphic clarity, water molecules
are rendered in the line-drawing mode.
P1: GXB
Structural Chemistry (STUC) pp709-stuc-456242 January 14, 2003 17:16 Style file version Nov. 07, 2000
Molecular Dynamics Simulations of Adsorption of Organic Compounds 181
Fig. 7. When the interlayer space of a clay mineral is randomly soaked with water (top), MD simulations will lead spontaneously, within
a few picoseconds, to the formation of several structured water layers (bottom).
P1: GXB
Structural Chemistry (STUC) pp709-stuc-456242 January 14, 2003 17:16 Style file version Nov. 07, 2000
182 Yu, Newton, Norman, Sch¨afer, and Miller
able to diffuse significantly, jumping from one surface site
to another. In contrast, coplanar adsorbates are effectively
immobilized.
The ability of pesticides to form complexes with
counter ions is an important factor in their transport prop-
erties in soils [62]. To investigate whether such complexes
are being formed to a statistically significant extent be-
tween DNOC and K+, the histogram of Fig. 5 was gen-
erated, involving distances between DNOC–oxygen and
K+on montmorillonite. From that analysis (Fig. 5), short-
rangeinteractions betweenDNOC–NO2andK+appear as
relatively infrequent. Further analyses with different ions,
such as Ca2+, are currently under investigation.
Fig. 8. The structure of the aqueous phase (2000 water molecules) of Fig. 6, after 150 ps of MD simulations. Note the distinct layers of water
immediately adjacent to the mineral surfaces and separated from the bulk phase in the interior of the interlayer space.
When the interlayer space is soaked with a large
amount of water, DNOC and ATZ show a tendency to
remain completely in the aqueous phase without mak-
ing contact with the mineral surface. We have obtained
the same result before for TCE [57] and methylene blue
[59]. For proteins in proximity to mineral surfaces, large
amountsofwater allow for retainingtheglobularstructure,
which dry surfaces destroy. Figure 6 shows the protein,
rubredoxin, in the presence of DNOC and some 2000 wa-
ter molecules on pyrophyllite. It is seen that, after 150 ps
of dynamics, the essential globular structure has been pre-
served.In asystem of thiskind, agivenpesticide caninter-
actwith boththe surfacesofthe mineraland ofthe protein.
P1: GXB
Structural Chemistry (STUC) pp709-stuc-456242 January 14, 2003 17:16 Style file version Nov. 07, 2000
Molecular Dynamics Simulations of Adsorption of Organic Compounds 183
Fig. 9. Distances ( ˚
A) between water–oxygen atoms and the crystallo-
graphic ab plane (vertical axis), and the bc plane (horizontal axis) taken
from a snapshot of a random starting configuration (top) and of an MD
equilibrated configuration (bottom) of water in a montmorillonite inter-
layer space.
The Structure of Water in Mineral Interlayer Spaces
In our studies of hydrated clay surfaces, we noticed
thatdynamics relaxationwill leadina shorttime fromper-
fectly disordered starting configurations to highly struc-
tured arrangements of water. The phenomenon is illus-
trated in Fig. 7, where three distinct water layers are seen
in a typical snapshot of hydrated DNOC on montmoril-
lonite. When the loading is very high, as in the system
shown in Fig. 8, narrow water layers will form adjacent to
a surface and are visibly separated from the bulk phase in
the interior of the interlayer space.
To study the dependence of water structures on state
variables, a model beidellite system, with a 4 ×2×1 su-
percell, 178 water molecules, and six charges compen-
sated by Ca2+, was subjected to MD equilibrations at a
pressure of 1 atm and temperatures of 300, 350, 400, 450,
and 500 K; and at a temperature of 300 K and external
pressures of 1, 2, 3, 5, and 10 atm. In the first series,
d(001) spacings were found at 20.46, 20.66, 20.89, 21.20,
and 21.75 ˚
A, respectively (averages of batch averages of
50 steps taken during the last 5 ps of 30 ps runs). As the
temperature increased, the layer structure was destroyed
by the high kinetic energy. In contrast, in the second se-
ries (80 ps runs), changes in pressure had no effect on the
water layers, and also the d(001) spacing was practically
unaffected (at 20.46, 20.40, 20.47, 20.45, and 20.40 ˚
A,
respectively).
The ordering of water in interlayer spaces is an in-
teresting phenomenon [27–29]. Plots of the distances be-
tween water oxygen atoms and mineral surfaces (Fig. 9)
show that the water arranges not only in layers along the
c-axis, but also along the other axes. These trends are
apparent in single snapshots of equilibrated systems, as
well as in ensemble averages (Figs. 10 and 11). We think
that the extent of the ordering is perhaps exaggerated by
the calculations. Some ordering of the kind revealed by
Figs. 9–11 must be expected, [27–29], but the absence of
polarization functions in the current empirical force fields
may very well lead to a higher order than is realistic.
CONCLUSIONS
Molecules are more strongly adsorbed on dry sur-
faces than on hydrated ones. The siloxane cavity of clay
minerals plays a major role in neutral molecule adsorp-
tion as it does in the adsorption of ions. Ions as well
as polar functional groups (NO2in DNOC and C Cl
in ATZ) position themselves in a characteristically off-
center arrangement relative to the hexagonal siloxane
cavity.
In the presence of water, the organic compounds
can make contact with clay surfaces by single functional
Fig.10. Histogram constructed from 1000 time steps of MD simulations
of water equilibrated in a montmorillonite interlayer space. Frequencies
of occurrence are shown (vertical axis) of distances ( ˚
A, horizontal axis)
between water–oxygen atoms and the crystallographic ab plane.
P1: GXB
Structural Chemistry (STUC) pp709-stuc-456242 January 14, 2003 17:16 Style file version Nov. 07, 2000
184 Yu, Newton, Norman, Sch¨afer, and Miller
Fig.11. Histogram constructed from 1000 time steps of MD simulations
of water equilibrated in a montmorillonite interlayer space. Frequencies
of occurrence are shown (vertical axis) of distances ( ˚
A, horizontal axis)
between water–oxygen atoms and the crystallographic bc plane.
groups and can then be expected to move rapidly through
soils. When adsorbed by full molecular contact, copla-
nar with the basal surface, the adsorbates are essentially
immobilized. From that position desorption by water can
occur, with different rates, depending on surface type and
type of compound adsorbed.
The current work was based entirely on empirical
computational techniques. In the future, quantum molec-
ular dynamics simulations, using techniques such as those
implemented by program CASTEP [63] can be expected
to be of increasing importance.
ACKNOWLEDGMENTS
The authors gratefully acknowledge support by
USDA CSREES grant 99-35107-7782 and by the IBM
Shared University Research Program. Special thanks are
due to Prof. Collis Geren. Vice Chancellor for Research,
University of Arkansas, and Dr. Jamie Coffin, IBM.
REFERENCES
1. Gay-Lussac, J. L. Mem. Soc. d Arcueil 1808, 2, 207.
2. Saalfeld, H.; Wedde, M. Z. Kristallogr. 1974, 139, 129–135.
3. Bish, D. L. Clays Clay Miner. 1993, 41, 738–744.
4. Lee, J. H.; Guggenheim, S. Amer. Miner. 1981, 66, 350–357.
5. Teppen, B. J. et al. J. Phys. Chem. 1997, 101, 1579–158.
6. MSI, Discover User Guide (1997) MSI Corp. (now Accelrys), San
Diego, California.
7. Maple, J. R. et al. J. Comp. Chem. 1994, 15, 162–182.
8. Hill, J.-R.; Sauer, J. J. Phys. Chem. 1994, 98, 1238–1244.
9. Hill, J.-R.; Sauer, J. J. Phys. Chem. 1995, 99, 9536–9550.
10. Allen,M.P.;Tildesley, D. J. Computer Simulation of Liquids; Oxford
University Press, Oxford, 1987.
11. Sanders, M. J.; Leslie, M.; Catlow, C. R. A. J. Chem. Soc. Chem.
Commun. 1984, pp. 1271–1273.
12. Collins, D. R.; Catlow, C. R. A. Molec. Simul. 1990, 4, 341–346.
13. Gale, J. D.; Catlow, C. R. A.; Mackrodt, W. C. Modeling Simul.
Mater. Sci. Eng. 1992, 1, 73–81.
14. Purton, J. et al. Phys. Chem. Miner. 1993, 19, 392–400.
15. Bush, T. S. et al. J. Mater. Chem. 1992, 2, 309–316.
16. Bush, T. S. et al. J. Mater. Chem. 1994, 4, 831–837.
17. Battle, P. D.; Bush, T. S.; Catlow, C. R. A. J. Amer. Chem. Soc. 1995,
117, 6292–6296.
18. Lewis, D. W. et al. J. Phys. Chem. 1995, 99, 2377–2383.
19. Collins, D. R.; Catlow, C. R. A. Acta Crystall. 1991, B 47, 678–682.
20. Collins, D. R.; Catlow, C. R. A. Amer. Miner. 1992, 77, 1172–1181.
21. Collins, D. R. et al. Phys. Chem. Miner. 1993, 19, 520–527.
22. Breu, J.; Catlow, C. R. A. Inorg. Chem.,1995, 34, 4504–4510.
23. Rustad,J. R.; Hay,B.P.Geochim. Cosmochim. Acta 1995,59,1251–
1257.
24. Rustad, J. R.; Hay, B. P.; Halley, J. W. J. Chem. Phys. 1995, 102,
427–431.
25. Rustad, J. R.; Felmy, A. R.; Hay, B. P. Geochim. Cosmochim. Acta
1996, 60, 1553–1562.
26. Rustad, J. R.; Felmy, A. R.; Hay, B. P. Geochim. Cosmochim. Acta
1996, 60, 1563–1576.
27. Delville, A. Langmuir 1991, 7, 547–555.
28. Delville, A. Langmuir,1992, 8, 1796–1805.
29. Delville, A. J. Phys. Chem. 1993, 97, 9703–9712.
30. Delville, A.; Sokolowski, S. J. Phys. Chem. 1993, 97, 6261–6271.
31. Skipper, N. T. Mineralogical Mag. 1998, 62, 657–667.
32. Skipper,N.T.;Refson,K.;McConnell, J. D. C.J.Chem.Phys. 1991,
94, 7434–7445.
33. Skipper, N. T.; Chang, F.-R. C.; Sposito, G. Clays Clay Miner. 1995,
43, 285–293.
34. Skipper, N. T.; Sposito, G.; Chang, F.-R. C. Clays Clay Miner. 1995,
43, 294–303.
35. Chang, F.-R. C. et al. In Mineral-Water Interfacial Reactions: Ki-
netics and Mechanisms, (Sparks, D. L. and Grunell, T. J., American
Chemical Society, Washington, D.C., 1999) pp. 88–106.
36. Chang, F.-R. C.; Skipper, N. T.; Sposito, G. Langmuir 1995, 11,
2734–2741.
37. Chang, F.-R. C.; Skipper, N. T.; Sposito, G. Langmuir,1997, 13,
2074–2082.
38. Boek, E. S.; Coveney, P. V.; Skipper, N. T. Langmuir 1995, 11,
4629–4631.
39. Boek, E. S.; Coveney, P. V.; Skipper, N. T. J. Amer. Chem. Soc. 1995,
117, 12608–12617.
40. Karaborni, S. et al. Science 1996, 271, 1102–1104.
41. Greathouse, J.; Sposito, G. J. Phys. Chem. B 1998, 102, 2406–2414.
42. Cygan, R. T. In Molecular Modeling of Clays and Mineral Surfaces;
Kubicki, J. D.; Bleam, W. F. Ed. Clay Minerals Society: Boulder,
CO, 1998, pp. 127–176.
43. Cygan, R. T.; Nagy, K. L.; Brady, P. V. In Adsorption of Metals by
Geomedia;Jenne,E. A., Ed.; Academic Press,SanDiego,CA,1998;
pp. 383–399.
44. Smith, D. E. Langmuir 1998, 14, 5959–5967.
45. Capkov´a, P. et al. Clays Clay Miner. 1998, 46, 232–239.
46. Capkov´a, P. et al. Clays Clay Miner. 1998, 46, 240–244.
47. Capkov´a, P. et al. J. Molec. Model. 1997, 3, 467–472.
48. Janeba, D.; Capkov´a, P.; Weiss, Z. J. Molec. Model. 1998, 4, 176–
182.
49. Sato, H.; Yamagishi, A.; Kato, S. J. Phys. Chem. 1992, 96, 9377–
9382.
50. Sato, H.; Yamagishi, A.; Kato, S. J. Phys. Chem. 1992, 96, 9382–
9387.
51. Sato, H. et al. J. Phys. Chem. 1996, 100, 1711–1717.
52. Breu, J.; Raj, N.; Catlow, C. R. A. J. Chem. Soc. Dalton 1999,
pp. 835–845.
53. Park,S.;Fitch,A.;Wang,Y.J.Phys.Chem.B1997,101,4889–4896.
54. Skipper,N.T.InMolecularModelingofClays and Mineral Surfaces.
Kubicki, J. D.; Bleam, W. F. Eds.; Clay Minerals Society, Boulder,
Colorado, 1998, pp. 90–126.
55. Keldsen, G. L. et al. J. Phys. Chem. 1994, 98, 279–284.
P1: GXB
Structural Chemistry (STUC) pp709-stuc-456242 January 14, 2003 17:16 Style file version Nov. 07, 2000
Molecular Dynamics Simulations of Adsorption of Organic Compounds 185
56. Cygan, R. T.; Kubicky, J. D. Eds., In Reviews in Mineralogy and
Geochemistry, Vol. 42; Geochemical Society, Rosso, J. J., Series
Ed.; Mineralogical Society of America: Ribbe, P. H., Series Editor;
Washington, 2001.
57. Teppen, B. J. et al. J. Comp. Chem. 1998, 19, 144–153.
58. Teppen, B. J. et al. Procedures of ECCOMAS, European Congress
on Computational Methods in Applied Sciences and Engineering,
Barcelona, 2000.
59. Yu, C.-H. et al. Clay Clay Miner. 2000, 48, 665–681.
60. Yu, C.-H. et al. J. Mol. Struct. 2000, 556, 95–103.
61. Dixon, J. B. In Minerals in Soil Environments, 2nd ed Dixon, J.
B.; Weed, S. B., Eds.; Soil Science Society of America: Madison,
Wisconsin, 1989, pp. 468–526.
62. Sheng, G. et al. J. Agr. Food Chem. 2001, 49, 2899–2907.
63. Payne, M. C. et al. Rev. Mod. Phys.1992, 64, 1045–
1097.
... -17 of 28 that organic inhibitors typically have long-chain structures containing multiple binding sites, it could be difficult for them to reach equilibrium on the calcite surface within the limited simulation time. Besides, Yu et al. [264] discovered that organic molecules often bind the mineral surfaces with a single functional group in aqueous solutions with the rest being immersed in the water. Therefore, investigating the adsorption properties of key functional groups seems to be a more practical approach. ...
Article
Full-text available
Calcium carbonate (CaCO3) is a crucial mineral with great scientific relevance in biomineralization and geoscience. However, excessive precipitation of CaCO3 is posing a threat to industrial production and the aquatic environment. The utilization of chemical inhibitors is typically considered an economical and successful route for addressing the scaling issues, while the underlying mechanism is still debated and needs to be further investigated. In this context, a deep understanding of the crystallization process of CaCO3 and how the inhibitors interact with CaCO3 nuclei and crystals are of great significance in evaluating the performance of scale inhibitors. In recent years, with the rapid development of computing facilities, computer simulations have provided an atomic‐level perspective on the kinetics and thermodynamics of possible association events in CaCO3 solutions as well as the predictions of nucleation pathway and growth mechanism of CaCO3 crystals as a complement to experiment. This review surveys several computational methods and their achievements in this field with a focus on analyzing the functional mechanisms of different types of inhibitors. A general discussion of the current challenges and future directions in applying atomistic simulations to the discovery, design, and development of more effective water‐scale inhibitors is also discussed.
... The number of load step in each calculated run is 6:0 × 10 6 , in which the balance step number is 3:0 × 10 6 and the process step number is 3:0 × 10 6 . This GCMC method was successfully used to calculate the absorption of CO 2 in magnesite in our previous works [35], and its validity also was proved by different works [31,[36][37][38][39][40]. ...
Article
Full-text available
CO2 transports in the Earth’s interior play a crucial role in understanding the deep carbon cycle and the global climate changes. Currently, CO2 transports inside of the Earth under extreme condition of pressure and temperature have not been understood well. In this study, the molecular dynamics (MD) calculations were performed to study CO2 transports under different CO2 pressures in slit-like magnesite pores with different pore sizes at 350~2500 K and 3~50 GPa are presented. Diffusion of CO2 in magnesite was improved as the temperature increases but showed the different features as a function of pressure. The diffusion coefficients of CO2 in magnesite were found in the range of 9×10−12 m2 s−1~28000×10−12 m2 s−1. Magnesite with the pore size of 20~25 Å corresponds to the highest transports. Anisotropic diffusion of CO2 in magnesite may help to understand the inhomogeneous distribution of carbon in the upper mantle. The time of CO2 diffusion from the mantle to Earth surface was estimated to be around several tens of Ma and has an important effect on deep carbon cycle. The simulation of CO2 transports based on the Earth condition provides new insights to revealing the deep carbon cycle in the Earth’s interiors.
... With the current high-performance computational resources, molecular simulations have become powerful tools for understanding the molecular-scale structural 18,19 , thermodynamic 20 , mechanical 21 and dynamic [22][23][24][25] properties of clay. Grand-canonical Monte Carlo (GCMC) method was used to examine the adsorption of CO 2 with H 2 O 23 , CH 4 26-28 and organic molecules 25,29 in clay. Yang et al. used molecular dynamics (MD) to study the structure and self-diffusion coefficient (SDC) of CO 2 in uncharged clay-like slit pores 24 . ...
Article
Full-text available
Storing CO2 in underground saline aquifers is an important way to reduce CO2 emission in atmosphere, where gas/fluid diffusion in clay plays a key role in CO2 leakage and migration. Various diffusivities, self-diffusivity, Maxwell–Stefan (M–S) diffusivity and Fick diffusivity, in clay interlayer are investigated by molecular dynamics (MD). Self-diffusivity varies with CO2 concentration, and reaches the maximum value at 2 molecules/unit-cell. High fluid concentration leads to clay swelling, thereby increasing self-diffusivity. However, the fractional free volume of clay explains the trend of CO2 self-diffusivity, which does not decrease with CO2 concentration monotonously but reaches the maximum when CO2 concentration reaches 2. Displacement distribution of CO2 molecules is analysed to explore the microscopic diffusion mechanism, which is characterised by logarithmic normal distribution. The mean value of such distribution further explains the self-diffusivity dependence on CO2 concentration. M–S and Fick diffusivities of CO2 are calculated by MD for the first time, both of which increase with increasing CO2 and H2O concentration and temperature. Based on self-diffusivity and M–S diffusivity, a quantity representing the coupling strength between CO2 molecules is presented; it increases firstly with CO2 concentration but begins to decrease when CO2 concentration is beyond 2.
... Several studies, summarized by Kalbitz [21], have shown that after heavy rainfall events and during snowmelt in the early stages, preferential flow through macropores in structure soils can affect DOC output by reducing contact time between the solid and solution phase. According to the laboratory experiments of Yu et al. [38], in the absence of water, organic compounds adsorb on the mineral surfaces in such a way that contact area is maximized. ...
Article
Full-text available
This study demonstrates the potential of the electrochemical methods for the characterization of dissolved organic matter (DOM) in the drainage water of hydroameliorated agricultural areas. A study of drainage water could lead to a better understanding of the distribution and fate of terrestrial DOM in the freshwater systems. We have applied the voltammetric techniques which were developed by our group for the characterization of organic matter in the natural waters in general. Studied samples were collected in the experimental amelioration fields in the Sava river valley (45° 33′ 52″ N/16° 31′ 33″ E, 100 m above sea level), in the hydroameliorated agricultural areas in Croatia. The rough characterization of the type, nature and reactivity of DOM was done through the study of surface activity (SA) of dissolved organic carbon (DOC), copper complexing capacity (CuCC) and apperent stability constants, and measurements of organic and inorganic reduced sulfur species (RSS) fractions. The results confirm that the electrochemical approach gives a valuable and comprehensive insight into physicochemical characteristics of DOM in the drainage water and could be successfully applied to temporal studies in different terrestrial ecosystem.
... Dans le premier cas, ces études explorent des stratégies de modélisation, le positionnement des atomes dans la structure cristalline , influence de la taille et la charge du contre-ion (Na + , K + , Li + , Cs + , Sr + ou Ca + ), la distribution des molécules d'eau autour du contre-ion, etc. Dans le deuxième cas, la modélisation est utilisée comme un outil pour étudier les matériaux composites (aussi dénommés « organoclays »), l'argile la plus suivant utilisée est la montmorillonite sodique avec des polymères de synthèse comme le polypropylène [170], la poly(ε-caprolactone) [162][163][164], la εcaprolactame [165], la rhodamine B [158,175], ou le nylon [169,171] ; dans ces structures, les cations sodium sont souvent échangés par des cations alkylammonium ou alkylphosphonium [167,168,174,176], ici une attention particulière est portée sur la détermination des distances interfoliaires et la comparaison des résultats avec mesures expérimentales (obtenues par Infrarouge ou par diffraction de rayons X). D'autre part, quelques études concernent la modélisation de l'adsorption de composés organiques [177,178], d'aminoacides [179], de protéines [180] ou des substances humiques [181] sur diverses surfaces minérales comme la beidellite, la muscovite, la hydrotalcite, la pyrophyllite ou la montmorillonite. Néanmoins, aucune étude par modélisation moléculaire n'a été reportée précédemment pour l'étude de l'adsorption des exopolysaccharides bactériens sur la montmorillonite sodique. ...
Article
The bacterial exopolysaccharides (EPS) play an important role in soil aggregate stabilisation during their association with mineral clays. This study uses molecular simulation methods to investigate the clay/EPS interfaces at the molecular level. Sodium montmorillonite (Na-MMT) was chosen as a representative model of mineral clay along with several EPS: dextran, MWAP71, xanthan, KYGT207, rhamsan, YAS34 and succinoglycan. Two aspects were explored: the adsorption of constitutive units of EPS onto basal surface of Na-MMT, and the relationships established between the different molecules onto clay/EPS/water complexes. All EPS tested exhibit a positive affinity for the mineral surface and multiple interactions are established at the interfaces, that can be ionic, hydrophobic or of hydrogen bond type. We found a linear correlation between values of adsorption enthalpy with the effective contact surface. Structural analysis indicates that the key factor determining the adsorption is the ability of the EPS to unfold in order to maximize its contact surface. The most flexible structures were the most favourable candidates for adsorption. Concerning the clay/EPS/water complexes, for the established models the interaction with water is stronger at lower hydration rates, indicating a resistance of the system to desiccation. Also, the interaction of EPS with mineral clay avoids the dispersion of the layers for water saturated systems. The results obtained by molecular modelling have good correlation with experimental observations.
... No attempt has been made to quantify the effect of methylgroup orientation because these rearrangements in general have a minor role in adsorption structure or energetics. At higher concentrations, aggregates of adsorbed MB + consist mainly of MB + oriented perpendicular to the surface ( Figure 4b), in agreement with MD simulations of MB + trimers and pentamers adsorbed on a dry mica surface ( Yu et al., 2003). Details of this surface structure will be explored in more detail below. ...
Article
Full-text available
Organic dyes such as methylene blue (MB) are often used in the characterization of clays and related minerals, but details of the adsorption mechanisms of such dyes are only partially understood from spectroscopic data, which indicate the presence of monomers, dimers, and higher aggregates for varying mineral surfaces. A combination of quantum (density functional theory) and classical molecular simulation methods was used to provide molecular detail of such adsorption processes, specifically the adsorption of MB onto kaolinite basal surfaces. Slab models with vacuum-terminated surfaces were used to obtain detailed structural properties and binding energies at both levels of theory, while classical molecular dynamics simulations of aqueous pores were used to characterize MB adsorption at infinite dilution and at higher concentration in which MB dimers and one-dimensional chains formed. Results for the neutral MB molecules are compared with those for the corresponding cation. Simulations of the aqueous pore indicate preferred adsorption on the hydrophobic siloxane surface, while charge-balancing chloride ions adsorb at the aluminol surface. At infinite dilution and in the gas-phase models, MB adsorbs with its primary molecular plane parallel to the siloxane surface to enhance hydrophobic interactions. Sandwiched dimers and chains are oriented perpendicular to the surface to facilitate the strong hydrophobic intermolecular interactions. Compared with quantum results, the hybrid force field predicts a weaker MB adsorption energy but a stronger dimerization energy. The structure and energetics of adsorbed MB at infinite dilution are consistent with the gas-phase binding results, which indicate that monomer adsorption is driven by strong interfacial forces rather than by the hydration properties of the dye. These results inform spectroscopic studies of MB adsorption on mineral surfaces while also revealing critical areas for development of improved hybrid force fields.
Article
Recent advancements in agriculture, industry, and pharmaceutical formulations have increased the presence of organic contaminants in the environment. It is important and necessary to study and understand the processes which control the environmental fate and transformation of contaminants and improve removal and remediation techniques. The use of advanced quantum mechanical modeling is a promising technique to better understand the mechanisms of adsorption within the environment. Relative Gibbs free energy values of adsorption have been calculated using such modeling for selected organic acids sorption to iron oxides, revealing the thermodynamic favorability of each of the reactions, except one involving bidentate mononuclear iron oxide. Theoretical spectra were constructed and evaluated with previous IR spectra studies to confirm the appropriate reliance on the modeling to accurately predict adsorption mechanisms.
Conference Paper
The purpose of this paper is to obtain the elasticity constants (including the volume and shear modulus, Young modulus, and Poisson’s ratio) of clay minerals, or montmorillonite, kaolinite and illite, using the molecular mechanics simulations. The Lennard-Jones potential function and Ewald summation method were respectively used to compute the Van der Waals’ and Coulomb forces. The integrated approach was made to perform the energy optimization of the mineral crystals. The unit and super cells of each mineral were then established to compute the elasticity constants of clay minerals. The substitutions, structural waters and anisotropy were also considered during the simulations. It shows that the montmorillonite easily move laterally while kaolinite is reasonably stable under vertical load; the substitution plays an important role for the elasticity constants of illite; the bulk modulus, shear modulus, Young’s modulus decrease if the number of waters increases; the intrinsical anisotropy occurred in the clay minerals; montmorillonite and illite may expand under external load.
Chapter
Environmental pollution is a major problem of our civilization mainly due to urban, industrial and agricultural human activities. This chapter summarizes molecular modeling methods, tools, and techniques relevant for organic and contaminant geochemistry. It presents the results of the authors&;#x00027; research group achieved approximately in the last decade. The examples represent mainly molecular modeling studies of interactions and mechanisms of binding several polar and nonpolar organic contaminants with typical minerals and natural organic matter (NOM). Special attention is devoted to developing force field (FF) models for water. The chapter emphasizes on the density functional theory (DFT)&;#x02010;based methods as they are by far the most frequently used methods to describe large molecular systems of the size of several hundred atoms. Coarse&;#x02010;grained force fields (CG&;#x02010;FF) simulations are frequently used in the simulations of processes such as self&;#x02010;assembling and folding of macromolecules.
Article
Cetyl trimethylammonium bromide (CTAB), a cationic surfactant, sodium dodecyl sulfate (SDS), an anionic surfactant, and hydrogenated castor oil (HCO), a nonionic surfactant, were sequentially intercalated into sodium montmorillonite (Mt). The basal spacing of Mt increased to 5.68. nm after the intercalation of CTAB and SDS. In addition, both FTIR analyses and molecular simulation confirmed the presence of these different types of surfactants in the Mt interlayer. Although the basal spacing decreased to 4.14. nm with the additional intercalation of HCO, the order and stability of the surfactants in the Mt interlayer increased. Thus, a high contact angle and good rheological property were achieved by the highly ordered surfactant arrangements.
Article
Full-text available
The crystal structure of Keokuk kaolinite, including all H atoms, was refined in space group C1 using low-temperature (1.5 K) neutron powder diffraction data (~ = 1.9102 A) and Rietveld refinement/ difference-Fourier methods to R,,,p = 1.78%, reduced x 2 = 3.32. Unit-cell parameters are: a = 5,1535(3) /~, b = 8.9419(5) &, c = 7.3906(4) &, a = 91.926(2) ~ = 105.046(2) ~ 7 = 89-797(2) ~ and V= 328.70(5) ik ~. Unit-cell parameters show that most of the thermal contraction occurred along the (001) direction, apparently due to a decrease in the interlayer distance. The non-H structure is very similar to published C1 structures, considering the low temperature of data collection, but the H atom positions are distinct. The inner OH group is essentially in the plane of the layers, and the inner-surface OH groups make angles of 600-73 ~ with the (001) plane. Difference-Fourier maps show minor anisotropy of the inner-OH group in the (001) direction, but the inner-surface OH groups appear to have their largest vibrational (or positional disorder) component parallel to the layers. Although no data indicate a split position of any of the H sites in kaolinite, there is support for limited random positional disorder of the H atoms. However, these data provided no support for a space group symmetry lower than C1.
Article
Full-text available
The structure of montmorillonite intercalated with (Al1304(OH)24+x(H20)l 2 x) (7 ~1* cations (AIt 7 ~ ~ for short), where x = 0, 2 and 4, has been studied using the Cerius z modeling environment. The Crystal Packer module used in the present study takes into account only the nonbonded interactions between the silicate layer and the Keggin cations. Minimization of the total sublimation energy led to the following conclusions: the structure of the interlayer (that is, the orientation of Keggin cations and the basal spacing) depends on the charge of cations (that is, on the degree of hydrolysis, x). The values of basal spacings in the range 19.38-20.27 ,~ have been obtained, depending on the charge and arrange- ment of cations in the interlayer. The dominating contribution to the total sublimation energy comes from the electrostatic interactions. Translations of All 7 ~ cations along the 2:1 layers give only small fluctu- ations of the total sublimation energy and basal spacings. No preference for the position of All 7'~+ cations in the interlayer of montmorillonite was found during translation along the 2:1 layers. This result con- firmed the inhomogeneous distribution of cations in the interlayer and turbostratic stacking of layers.
Article
Molecular dynamics simulations were performed of the adsorption of methylene blue (MB) on model beidellite, montmorillonite, and muscovite mica surfaces, using a previously determined empirical force field developed for dioctahedral clays. The simulations show that the adsorption of MB on mineral surfaces can result in a variety of configurations, including single and double layers of MB parallel to the basal surface, and irregular clusters. The d(001) values of approx. 12.3 and approx. 15.7 angstrom are assigned to dry phases with parallel single and double layers of MB, respectively, in agreement with X-ray studies. At intermediate MB loadings, stacks inclined to basal surfaces are formed. The stacks of MB ions inclined by 65-70° relative to the (001) plane of muscovite are not found on dry surfaces, in contrast to previous studies. Configurations similar to those proposed by others form spontaneously in the presence of H2O, but the ions in the model systems are not quite as ordered and not ordered in exactly the same way as the ones previously described, and they display a mobility that is not compatible with strict atomic order. The formation of a triple layer of H2O interspersed with ions may occur in the interlayer. Overall, the results of the simulations confirm that the MB-ion method must be used with great caution in surface-area determinations, because of the multiplicity of possible configurations. At the same time, the ability for adsorption to occur as either single or multiple MB layers is useful to determine cation-exchange capacity over a wide range of surface-charge densities.
Article
A new approach to estimating stability constants for proton binding in multisite surface complexation models is presented. The method is based on molecular statics computation of energies for the formation of proton vacancies and interstitials in ideal periodic slabs representing the (100), (110), (010), (001), and (021) surfaces of goethite. Gas-phase energies of clusters representing the hydrolysis products of ferric iron are calculated using the same potential energy functions used for the surface. These energies are linearly related to the hydrolysis constants for ferric iron in aqueous solution. Stability constants for proton binding at goethite surfaces are estimated by assuming the same log K-{Delta}E relationship for goethite surface protonation reactions. These stability constants predict a pH of zero charge of 8.9, in adequate agreement with measurements on CO{sub 2}-free goethite. The estimated stability constants differ significantly from previous estimations based on Pauling bond strength. We find that nearly all the surface oxide ions are reactive; nineteen of the twenty-six surface sites investigated have log K{sup int} between 7.7 and 9.4. This implies a site density between fifteen and sixteen reactive sites/nm for crystals dominated by (110) and (021) crystal faces. 39 refs., 8 figs., 4 tabs.
Article
The Crystal Packer module in the Cerius 2 modeling environment has been used to study the structure of montmorillonite intercalated with Al(OH)3-fragment (gibbsite-like) polymers. Basal spacings in gibbsite-like polymers arranged in 2 layers in the interlayer of montmorillonite varied in the range 19.54-20.13 A, depending on the type and arrangement of AI(OH) 3 fragments. The inhomogeneous distribution of intercalating species in the interlayer and, consequently, the turbostratic stacking of layers has been found for gibbsite-like polymers as well as in the case of Keggin cations (Capkov~i et al. 1998). The dominating contribution to the total sublimation energy comes from electrostatic interactions for both intercalating species, gibbsite-like polymers and Keggin cations.
Article
Molecular dynamics simulations were performed of the adsorption of methylene blue (MB) on model beidellite, montmorillonite, and muscovite mica surfaces, using a previously determined empirical force field developed for dioctahedral clays. The simulations show that the adsorption of MB on mineral surfaces can result in a variety of configurations, including single and double layers of MB parallel to the basal surface, and irregular clusters. The d (001) values of ~12.3 and ~15.7 A are assigned to dry phases with parallel single and double layers of MB, respectively, in agreement with X-ray studies. At intermediate MB loadings, stacks inclined to basal surfaces are formed. The stacks of MB ions inclined by 65–70° relative to the (001) plane of muscovite are not found on dry surfaces, in contrast to previous studies. Configurations similar to those proposed by others form spontaneously in the presence of H2O, but the ions in the model systems are not quite as ordered and not ordered in exactly the same way as the ones previously described, and they display a mobility that is not compatible with strict atomic order. The formation of a triple layer of H2O interspersed with ions may occur in the interlayer. Overall, the results of the simulations confirm that the MB-ion method must be used with great caution in surface-area determinations, because of the multiplicity of possible configurations. At the same time, the ability for adsorption to occur as either single or multiple MB layers is useful to determine cation-exchange capacity over a wide range of surface-charge densities.
Article
Monte Carlo (MC) simulations of molecular structure in the interlayers of 2:1 Na-saturated clay minerals were performed to address several important simulation methodological issues. Investigation was focused on monolayer hydrates of the clay minerals because these systems provide a severe test of the quality and sensitivity of MC interlayer simulations. Comparisons were made between two leading models of the water-water interaction in condensed phases, and the sensitivity of the simulations to the size or shape of the periodically-repeated simulation cell was determined. The results indicated that model potential functions permitting significant deviations from the molecular environment in bulk liquid water are superior to those calibrated to mimic the bulk water structure closely. Increasing the simulation cell size or altering its shape from a rectangular 21.12 Angstrom x 18.28 Angstrom x 6.54 Angstrom cell (about eight clay mineral unit cells) had no significant effect on the calculated interlayer properties.
Article
Monte Carlo and molecular dynamics computer simulations are now able to provide detailed information concerning the structure, dynamics, and thermodynamics of pore fluids in 2:1 clays. This article will discuss interparticle interaction potentials currently available for atomistic simulations of clay-water systems, and will describe how computational techniques can be applied to modelling of clay systems. Some recent simulation studies of 2:1 clay hydration will then be reviewed. Comparison with experimental data promotes confidence in the molecular models and simulation techniques, and points to exciting future prospects.