Vous êtes sur la page 1sur 8

Computational Materials Science 111 (2016) 269276

Contents lists available at ScienceDirect

Computational Materials Science


journal homepage: www.elsevier.com/locate/commatsci

Bond-order reactive force fields for molecular dynamics simulations of


crystalline silica
Benjamin J. Cowen a,b, Mohamed S. El-Genk a,b,c,d,
a

Institute for Space and Nuclear Power Studies, University of New Mexico, Albuquerque, NM, USA
Nuclear Engineering Department, University of New Mexico, Albuquerque, NM, USA
c
Mechanical Engineering Department, University of New Mexico, Albuquerque, NM, USA
d
Chemical and Biological Engineering Department, University of New Mexico, Albuquerque, NM, USA
b

a r t i c l e

i n f o

Article history:
Received 29 May 2015
Received in revised form 10 September
2015
Accepted 16 September 2015
Available online 18 October 2015
Keywords:
Bond-order
Variable-charge reactive force fields
MD simulation of silica polymorphs
Alphabeta transition
Accuracy and transferability

a b s t r a c t
This paper investigates the applicability of the bond-order, variable-charge (BOVC) force fields of the
H2 O
, and ReaxFFGSI
Charge-Optimized Many-Body (COMB10), ReaxFFSiO
SiO , for molecular dynamics (MD) simulations of crystalline SiO2. The calculated lattice constants and densities of the four SiO2 polymorphs,
quartz, cristobalite, coesite, and stishovite, are compared to experimental values. Additionally, the calculated pair distribution and bond-angle distribution functions and the ab transition for quartz, the most
stable low-energy polymorph, are compared to experimental results. The simulations with the COMB10
force field accurately predict the properties of the SiO2 polymorphs, except the a-cristobalite, and the
GSI
2O
quartz ab transition. The results with ReaxFFH
SiO and ReaxFFSiO accurately predict the properties of
2O
the SiO2 polymorphs, except the stishovite, but those with ReaxFFH
SiO inaccurately predict the quartz
ab transition.
2015 Elsevier B.V. All rights reserved.

1. Introduction
Characterizing the interface of silica (SiO2) with other materials,
such as silicon (Si), water (H2O), and oxygen (O), is important in a
wide range of applications. Examples are the interaction of SiO2
with biomolecules, such as H2O, in chromatography, geochemistry,
biomedical devices, and the encapsulation of biomolecules within
silica gels [13]. The Si/SiO2 interface is encountered in the manufacturing of semiconductor transistors [4,5]. The O/SiO2 interface is
encountered in the dry-etching and degradation of silica surfaces,
commonly used in the aerospace industry, due to the collision by
atomic oxygen [6,7]. The challenging characterization of these
interfaces experimentally has stimulated a growing interest in
using molecular dynamics (MD) simulations for that purpose.
The MD simulations employ force fields for modeling atom
atom interactions in the materials of interest and the force fields
are fitted to high-level ab initio or experimental results at specified
temperatures and pressures [6,7]. Some force fields are rather simple, such as the Lennard-Jones pair potential [8], originally developed to model noble gases, while others are parameterized for
modeling complex systems and interfaces. Examples are the twobody force fields with Coulomb interactions (e.g. BKS [6], Pedone
Corresponding author at: Institute for Space and Nuclear Power Studies,
University of New Mexico, Albuquerque, NM, USA.
http://dx.doi.org/10.1016/j.commatsci.2015.09.042
0927-0256/ 2015 Elsevier B.V. All rights reserved.

[7], TTAM [8], CHIK [9]), the three-body force fields with no Coulomb interactions (e.g. Munetoh [10]), and the bond-order,
variable-charge (BOVC) force fields [1113]. The BOVC force fields
are the most suitable for modeling complex chemical reactions and
interfaces.
Two of the most popular BOVC force fields are the ChargeOptimized Many-Body (COMB) and ReaxFF. Since the introduction
of the original COMB for simulating the Si/SiO2 interface [14], and
the original ReaxFF for simulating hydrocarbons [15], several
materials have been parameterized in forms of these powerful
force fields. For instance, the capabilities with COMB have been
successfully extended for modeling the interfaces of hafnium/
hafnium oxide [16] and Cu/SiO2 [17], as well as zirconium [18].
Similarly, the capabilities of ReaxFF have been extended for modeling the aluminummolybdenum alloys [19], zinc oxide [20], and
the silicon/silica system [21].
The COMB10 [11] force field is an extension of COMB07 [14],
developed as an extension of the Tersoff-based potential by Yasukuawa [22] for semiconductors. The COMB10 force field, parameterized for modeling the Si/SiO2 interface and amorphous silica,
accurately predicts the structural and elastic properties of crystalline SiO2 and the trend of defect formation in amorphous silica
[11].
The ReaxFF is the foundation of two other parameterizations
H2 O
involving silica. The first is that of Fogarty et al. [12] (ReaxFFSiO
),

270

B.J. Cowen, M.S. El-Genk / Computational Materials Science 111 (2016) 269276

who extended the parameterization by van Duin et al. [21]


(ReaxFFSiO) to modeling the H2O/SiO2 interface. The ReaxFFSiO,
based on the radical reactions described by quantum mechanical
data, was unable to describe the proton transfer reactions at the
interface of water with silica [12]. Fogarty et al. [12] replaced the
O/H parameters in the ReaxFFSiO with a set fitted to the protontransfer reactions and the water clusters from two quantum
mechanics datasets [24]. These datasets described the reaction
energies for the Si(OH)4 polymerization, and the binding and dissociation of the water molecules from Si(OH)4 [23]. Thus, the Si/H, Si/
Si, and Si/O parameters were refitted [24]. Fogarty et al. [12] provide detailed information on the refitting schemes.
The second parameterization of ReaxFF is that by Kulkarni et al.
[16] ReaxFFGSI
SiO . They extended ReaxFFSiO to modeling the gassurface interactions, specifically that of oxygen with silica. They
carried out electronic structure calculations to obtain a highquality dataset for describing the geometry and the density functional theory energies, and used the new quantummechanical
data to re-parameterize previous work [21]. The obtained parameters were optimized to reproduce the binding energy variations of
the gas-surface interactions.
Cowen and El-Genk [24] have reviewed and investigated four
fixed-charge, two-body force fields (BKS [6], Pedone [7], TTAM
[8], CHIK [9]) and one no-charge, three-body (Munetoh [10]) force
field for modeling silica, and compared the results of the computational cost, accuracy, and transferability. The results of the MD
simulations with the BKS force field accurately predicted the
experimental values of the density, cell volume, and radial and
bond-angle distribution functions for quartz, cristobalite, coesite,
and stishovite, to within 2%. In addition, the results of the ab transition and the quartz III transition were in good agreement with
the experimental results [24]. The results of the simulations with
both the CHIK and TTAM force fields, over the entire range of the
silica phase diagram, were slightly less accurate than with the
BKS potential. The simulations with the Munetoh force field were
the cheapest in terms of the short computational time, but also
the least accurate [24]. The predictions of the structural properties
sometimes differed from experimental values by more than 10%.
The simulations with the Munetoh force field did not produce
the ab and III phase transitions or the equation of state for
stishovite silica. The results of the MD simulations with the Pedone
potential of the material properties were within 5% of experimental values, except for the high-pressure polymorph of stishovite. The properties of stishovite were overestimated by 8% and
the quartz III phase transition was overestimated by as much as
60% [24].
This work extends the previous effort by Cowen & El-Genk [24]
to investigate the applicability of three variable-charge force fields
H2 O
ReaxFFSiO

ReaxFFGSI
SiO

(COMB10 [11],
[12],
[13]) for modeling crystalline silica, developed specifically for characterizing complex
interfaces and reactions. Investigated are the accuracy and transferability of these force fields for calculating the material properties, including the lattice constants, density, and the pair and
bond-angle distribution functions. The transferability is examined
by comparing the simulation results to the experimental values
for the various silica polymorphs; namely, quartz, cristobalite, coeH2 O
site, and stishovite. The COMB10 [11] and ReaxFFSiO
[12] are also
investigated for predicting the ab phase transition for quartz.
The MD simulations with the BOVC force fields require, as an
input, the frequency of updating the atom charges. Generally, it
is beneficial to update the atom charges as frequently as possible,
without sacrificing either accuracy or computational efficiency.
Updating the atom charges every time step, although computationally expensive, ensures that the charges are at their energy minima. Updating less frequently decreases the computational time,

but compromises the energy conservation. This work examined


the effect of changing the frequency of updating the atom charges
on the calculated results. Two updating options are considered:
every time step, and every 10-time steps. The number that follows
the abbreviation of the force field indicates the updating frequency
(e.g., COMB10-1 refers to COMB10, with atom charges updated
every time step).
2. Methodology
The crystallographic information files (CIF) for all four polymorphs of silica were obtained from the Crystallography Open
Database [2527]. The silica polymorphs are replicated in all three
spatial dimensions using symmetry operations to create a superGSI
2O
cell. The MD simulations with COMB10, ReaxFFH
SiO , and ReaxFFSiO ,

are carried out using the Large-Scale Atomic/Molecular Massively


Parallel Simulator (LAMMPS) code [28], with a time step of 0.2 fs.
Aktulga et al. [29] implemented the ReaxFF parameterization into
LAMMPS.
The performed MD simulations track the positions and velocities of the atoms by time-integrating the NoseHoover style,
non-Hamiltonian equations of motion, specifically by sampling
from the isothermalisobaric (NPT) ensemble [28]. The QEq [30]
scheme is used to calculate the charge of each atom, based on
the local bonding environment. The MD simulations of quartz,
cristobalite, coesite, and stishovite used 5184, 768, 768, and
864 atoms, respectively. The energy minimization is carried out
using the PolakRibiere version of the conjugate gradient (CG)
algorithm [31], as discussed by Cowen and El-Genk [24]. The simulated systems are equilibrated for 100 ps and the results of the
lattice constants, the average charge, and density, are collected
over the following 100 ps. The ranges of charges are obtained by
finding the minimum and maximum charges of the atoms following data collection.
2.1. Force fields
The COMB10 [11] and ReaxFF [12,13] force fields are expressed
somewhat differently. The COMB10 [11] force field is given as:

ET Eself ECoul EvdW Ebond Eothers

The ReaxFF [12,13] is expressed as:

ET Eself ECoul EvdW Ebond Eangle Etorsion Econjugation


EHbond Elonepair Eover Eunder Eothers

The first four terms in Eq. (2) are basically the same for the
COMB10
[11]
force
field,
Eq.
(1).
Other
terms:
Eangle ; Etorsion ; Eover ; Eunder ; Econjugation

in Eq. (2) are incorporated

implicitly in the Ebond term of the COMB10 force field (1). The
Eself term represents the energy difference in the various states of
the atom charges, while Eothers consists of several terms that correct
for allene-type terminal triple bond, and C2 species [32]. For more
details on the physics of the various terms in Eqs. (1) and (2), the
reader is referred to the original papers on the BOVC force fields
[1113].
2.2. Quartz ab transition
In the performed MD simulations of the ab transition for
H2 O
H2 O
quartz using COMB10-1, ReaxFFSiO
-1, and ReaxFFSiO
-10, the temperature in the ranges of 300700 K and 9001500 K is increased
in increments of 100 K. In the intermediate range of 700900 K,
the temperature increased in small increments of 10 K. The

271

B.J. Cowen, M.S. El-Genk / Computational Materials Science 111 (2016) 269276

COMB10-10 force field required a slightly different approach,


because the initial simulations predicted a structural transformation outside the 700900 K range. Thus, the range for using the
10 K increments is shifted to higher temperatures to accurately
capture the quartz ab transition. In the simulations with the
COMB10-10 force field, the temperature in the ranges of 300
1100 K and 14001500 K is increased in increments of 100 K.
Increments of 10 K are used in the intermediate range of 1100
1400 K. The used heating time between temperatures is 1.0 ps,
whereas equilibration times of 1, 2, 4, 8, and 16 ps and equal times
for data collection are used. For example, a 16-ps of data collection
time followed a 16-ps equilibration time at each temperature.
3. Results and discussion
This section compares the calculated properties of the crystalline silica polymorphs of quartz, cristobalite, coesite, and stisho-

distribution functions. The presented results demonstrate the


effect of changing the frequency of updating the atom charges
every time step, and every 10-time steps on the calculated materials properties. Only the MD simulations with the COMB10 and
H2 O
ReaxFFSiO
force fields are investigated for predicting the quartz
ab transition.

3.1. Quartz
Quartz, the most abundant and stable SiO2 polymorph at ambient conditions, is used in many applications due to its superior
optical and piezoelectric properties and the formation of SiO2 at
the SiO2/Si interface [3335]. Table 1 compares the calculated
properties (standard deviation shown in parenthesis) and the atom
charges for a-quartz, using MD simulations with the BOVC force
fields to experimental values. The simulations results with the
ReaxFFGSI
SiO -1 accurately predict the lattice constants and density of

H2 O
vite, using MD simulations with COMB10, ReaxFFSiO
; and

a-quartz, to within 2.1% of experimental values. The calculated

ReaxFFGSI
SiO , with experimental results. These material properties
are the lattice constants, density, and the pair and bond-angle

H2 O
material properties in the simulations with the ReaxFFSiO
-1 are
within 3.6%, and those with COMB10 are within 7.3%, of the

Table 1
Comparison of calculated properties (standard deviation shown in parenthesis) and atom charges for a-quartz, using MD simulations with the BOVC force fields to experimental
values.
Parameter

a ()
c ()
q (g/cm3)
qavg (O)
qavg (Si)
qrange (O)
qrange (Si)

Experimental/force field
Exp.[36]

COMB10-1

COMB10-10

H2 O
-1
ReaxFFSiO

H2 O
ReaxFFSiO
-10

ReaxFFGSI
SiO -1

4.916
5.405
2.646

4.79 (4e3)
5.30 (6e3)
2.84 (3e3)
1.45 (1e4)
2.91 (3e4)
1.38 to 1.54
2.882.93

4.79 (3e3)
5.28 (5e3)
2.85 (2e3)
1.45 (5e5)
2.91 (1e4)
1.39 to 1.53
2.892.92

4.95 (3e3)
5.52 (3e3)
2.55 (1e3)
0.81 (2e3)
1.62 (4e3)
0.69 to 0.91
1.401.80

4.93 (1e2)
5.48 (3e2)
2.60 (3e3)
0.81 (1e3)
1.62 (4e3)
0.69 to 0.91
1.581.64

4.95 (3e3)
5.46 (3e3)
2.59 (1e3)
0.46 (9e4)
0.92 (2e3)
0.41 to 0.50
0.880.96

H2 O
Fig. 1. Comparison of calculated pair distribution functions for a-quartz, at ambient conditions (300 K and 1 atm), in MD simulations with COMB10 and ReaxFFSiO
to
experimental results.

272

B.J. Cowen, M.S. El-Genk / Computational Materials Science 111 (2016) 269276

experimental values. In the MD simulations of a-quartz with the


COMB10 force field, the charge-equilibration frequency negligibly
2O
affects the results. In the simulations with the ReaxFFH
SiO , there is
a larger disparity in the results, predicting a-quartz densities of
2.55 g/cm3 and 2.60 g/cm3 with a charge equilibration frequency
of every time step and every 10-time steps, respectively.
Fig. 1 compares the calculated pair distribution functions for
a-quartz at ambient conditions, in the MD simulations with the

H2 O
H2 O
COMB10-1, COMB10-10, ReaxFFSiO
-1, and ReaxFFSiO
-10 force
fields, with experimental values. The pair distribution functions
are important for showing the structural order of crystalline materials. There are differences in the calculated distributions with
updating the atom charges every time step and every 10-time
steps. The distribution functions curves, produced by evaluating
the atom charges every time step are smoother and more accurate,
compared to experimental values. The most accurate curves for the
OSi and OO pairs are those calculated in the simulations with the
COMB10-1 force field. They are accurate to within 1% of the

H2 O
the calculated curves with ReaxFFSiO
-10, which disappear when
updating the atom charges every time step. The MD simulations
with the COMB10 force field underestimate the experimental
results of the SiOSi BAD by 4%, while the simulations with
H2 O
ReaxFFSiO
over-predict the experimental results by as much as
10% (Fig. 2). For judicious comparison to experimental results,
Table 2 lists the ranges of the calculated bond-angles at the base
and half-width, the average values of the bond-angles, and the
bond-angles at the peaks of the BAD functions. The results in this
table show close agreements with the experimental values of the
average bond angles.
The equilibration frequency of the atom charges clearly affects
the simulations results, as demonstrated in Figs. 1 and 2. In
addition to not conserving energy, when updating the atom
charges each 10 time steps, the peaks of the calculated pair and

H2 O
experimental results. The simulation results with ReaxFFSiO
-1 for
the SiSi pairs are the most accurate; they are within 1% of experimental values. The MD simulations with the COMB10 force field
underestimate the SiSi distribution by 3%, and that with the
H2 O
ReaxFFSiO
underestimate both the OSi and OO pairs by 3%,
compared to the experimental values of the pair distribution functions (Fig. 1). For an accurate characterization, Table 2 lists the
ranges of the calculated bond-lengths at the base and half-width,
the average values of the bond-lengths, and the bond-lengths at
the peaks of the pair distribution functions.
Fig. 2 compares the calculated OSiO and SiOSi bond-angle
distribution (BAD) functions for a-quartz at ambient conditions,
using MD simulations with the COMB10-1, COMB10-10,
H2 O
H2 O
ReaxFFSiO
-1, and ReaxFFSiO
-10 to the experimental values. The
results with all BOVC force fields predict the OSiO BAD to within
1% of the experimental values. However, there are slight kinks in

Fig. 2. Comparison of calculated BAD functions for a-quartz, at ambient conditions


H2 O
(300 K and 1 atm), in MD simulations with COMB10 and ReaxFFSiO
to experimental
results.

Table 2
Comparison of the bond-lengths at the 1st peak of the pair distribution function, and the bond-angles for a-quartz, using MD simulations with the BOVC force fields, to
experimental values.
Force field

Atom Pairs

Bond lengths for 1st peak in RDF ()


MinMax @ Base

MinMax @ Half-Width

@ Peak

Average value
Calculated

Exp. [36]

COMB10-1

OSi
OO
SiSi

1.461.86
2.292.87
2.763.41

1.571.64
2.542.72
2.953.05

1.61
2.63
2.98

1.62
2.63
3.01

1.61
2.612.64
3.06

COMB10-10

OSi
OO
SiSi

1.511.80
2.332.82
2.853.38

1.591.64
2.562.70
2.953.02

1.61
2.61
2.99

1.62
2.62
3.02

1.61
2.612.64
3.06

H2 O
ReaxFFSiO
-1

OSi
OO
SiSi

1.461.78
2.192.94
2.943.31

1.531.61
2.472.66
3.043.13

1.57
2.55
3.08

1.57
2.56
3.09

1.61
2.612.64
3.06

H2 O
ReaxFFSiO
-10

OSi
OO
SiSi

1.521.65
2.162.90
3.023.23

1.561.59
2.442.67
3.063.10

1.57
2.52
3.08

1.57
2.57
3.09

1.61
2.612.64
3.06

Interface

Bond angles ()

COMB10-1

OSiO
SiOSi

89.9130.5
99.2153.5

103.6115.2
133.6141.5

108.9
136.7

109.3
137.7

109.5
143.7

COMB10-10

OSiO
SiOSi

94.4128.5
128.6155.3

104.4115.1
134.1140.1

108.7
137.2

109.3
138.0

109.5
143.7

H2 O
ReaxFFSiO
-1

OSiO
SiOSi

91.2132.6
137.5179.6

103.6114.6
151.9163.5

108.3
156.6

109.4
158.0

109.5
143.7

H2 O
ReaxFFSiO
-10

OSiO
SiOSi

89.6134.2
146.6172.0

102.6115.5
154.2160.8

107.0
157.0

109.3
157.7

109.5
143.7

B.J. Cowen, M.S. El-Genk / Computational Materials Science 111 (2016) 269276

bond-angle distribution functions are not smooth, unlike those


with updating the atom charges each time step. Notwithstanding
that, the average values are still remarkably close to the experimental results for both charge-equilibration frequencies. At ambient conditions, the impact of the charge-equilibration frequency on
the MD simulation results is small; however, predicting the complex phase transition is a challenge.
3.1.1. ab transition
The ab phase transition for quartz has been shown experimentally to occur at 846 K [37]. A good test of the transferability of a
force field in MD simulations is to characterize this transition. This
section compares the calculated cell volume and c/a ratio at the a
b transition in MD simulations with the different BOVC force fields,
as a function of temperature, to reported experimental values.
3.1.1.1. Cell volume. Fig. 3 shows that in the performed MD simulations with the COMB10-1 force field for durations of 1, 2, and 4 ps,
the calculated cell volume steadily increases with increasing temperature, with no structural transformation. The simulation results
with the COMB10-1 force field and with equilibration and data collection times of 8 ps and 16 ps show a sharp drop in the calculated
cell volume at the ab transition. Conversely, the experimental
results indicate a plateau in the cell volume expansion at the ab
transition, due to the rotation of the SiO4 tetrahedra.
The MD simulations with the COMB10-10 force field (Fig. 3) and
different equilibration times reveal a structural transformation, but
at a much higher temperature than predicted with the COMB10-1
potential. As with the COMB10-1 potential, the not fully converged
results show a continuous decrease in the cell volume at the ab
transition, rather than reaching a plateau, as indicated experimentally (Fig. 3).
2O
ReaxFFH
SiO -1

The MD simulation results with the


converged for
all cases (Fig. 3). However, a plateau is not predicted over the entire
temperature range (3001500 K) of interest, indicating the inabilH2 O
ity of the ReaxFFSiO
-1 to model the ab transition for quartz. Only

273

the simulation results of the cell volume expansion using the


H2 O
ReaxFFSiO
-10, with 16 ps of equilibration and data collection time,
predict a structural transformation. However, the transformation
does not predict the experimentally measured plateau at the ab
transition (Fig. 3).
The simulations with the BOVC force fields and with updating
the atom charges every 10 time steps give drastically different
results than when updating the charges every time step. Indeed,
the period of the vibrational modes for the high frequency asymmetric SiO bond stretching is approximately an order of magnitude larger than the 2 fs time interval for the lower chargeequilibration frequency [38]. However, the charge-equilibration
frequency is uncoupled to the vibrational modes. Thus, in the MD
simulations with the force fields of COMB and ReaxFF, the atom
charges need to be updated every time step in order to find the
local energy minima. As indicated earlier, updating the atom
charges less frequently no longer conserves energy.
At ambient conditions, the effect of the charge-equilibration
frequency on the calculated results is small, but more pronounced
over the quartz ab transition, as the results using chargeequilibration frequencies of each time step and each 10 time steps
diverge. Therefore, for the complex ab structural phase transition,
it is important to update atom charges as often as possible to compensate for the change in the bonding environment. Perhaps there
is a more optimal value between the two investigated in this work,
namely each time step and 10-time steps. Even when the atom
charges are updated every time step, the MD simulations results
H2 O
with the COMB10 and the ReaxFFSiO
inaccurately predict the
quartz ab transition.

3.1.1.2. c/a. The c/a ratio at the ab transition for quartz is difficult
to model and has eluded many force fields. Herzbach et al. [39]
indicated that the polarizable TS [40] is the only force field, among
those they investigated, that is capable of modeling the sudden
drop in c/a ratio at the ab transition for quartz. The present MD
simulation results with the COMB10-1 and COMB10-10 force fields

H2 O
Fig. 3. Comparison of calculated cell volume expansion at the ab transition of quartz in MD simulations with COMB10 and ReaxFFSiO
to experimental results.

274

B.J. Cowen, M.S. El-Genk / Computational Materials Science 111 (2016) 269276

H2 O
Fig. 4. Comparison of calculated c/a ratio at the ab transition of quartz, in MD simulations with COMB10 and ReaxFFSiO
to experimental results.

inaccurately predict the c/a ratio at the ab transition (Fig. 4). At


the ab transition, the calculated c/a ratio experiences a sudden
increase, rather than a decrease, as obtained experimentally. The

a-quartz of 2.84 g/cm3. Thus, it is unclear how reported work

H2 O
performed simulations with the ReaxFFSiO
-10 show a similar

H2 O
simulations with both ReaxFFSiO
and ReaxFFGSI
SiO are within 3%
and 5% of the lattice constants and density, respectively, compared
to experimental values (Table 3). These results demonstrate the
transferability of these force fields to the a-cristobalite bonding
environment.

increase in the c/a ratio near the ab transition. The


is the only force field that does not show any structural transformation in the calculated c/a ratio.
H2 O
ReaxFFSiO
-1

[11] could have predicted values closer to experimental results.


The calculated properties for a-cristobalite in the performed MD

3.2. Cristobalite

3.3. Coesite

Similar to quartz, cristobalite is formed at higher pressures, and


is a four-coordinated polymorph. However, the density is lower,
comparable to that of amorphous silica. Based on the previous
results showing that evaluating the atom charges every 10
time steps is not quite accurate and does not conserve energy
(Figs. 14), the results of the performed MD simulations discussed
in remainder of this paper, are with updating the atom charges
every time step.
In Table 3, the calculated structural properties and charges of
the oxygen and silicon atoms of a-cristobalite, in MD simulations
with the three BOVC force fields investigated in this work, are compared to experimental values. Despite the fact that a-quartz and acristobalite have similar crystal structures, with the silicon atoms
tetrahedrally coordinated, the results of the MD simulations with
the COMB10 force field of the material properties are not as accurate for this polymorph, as they are for a-quartz (Table 3). Despite
the extensive simulations with different time steps, frequencies of
updating the atom charges of the atoms, equilibration and data collection times, and number of atoms, the calculated results for this
polymorph are significantly different from those previously
reported with the COMB10 force field [11]. Energy minimization
of the force field produces results consistent with the experimental
values, but equilibration at ambient conditions produces a 21%
higher density. The calculated density in this work for the
a-cristobalite of 2.81 g/cm3 is similar to that calculated for

Coesite is another four-coordinated silica polymorph, which


forms at high pressures due to events such as meteor impacts
[42]. The bonding environment is thus much different from those
for quartz and cristobalite. The monoclinic crystal structure of coesite comprises four silicon dioxidetetrahedra in Si4O8 and Si8O16
rings [43]. Table 4 compares the calculated properties and the
charges of the oxygen and silicon atoms for coesite with experimental values. The results of the performed MD simulations using
all three BOVC force fields accurately predict the properties of
coesite.
The monoclinic cell of coesite is characterized as a b c and
b 90. The b angle is shown experimentally to be 120.34 [43].
All three force fields predict this angle to within 1 of the experimental value. The MD simulation results of the coesite properties,
H2 O
with the COMB10, ReaxFFSiO
, and ReaxFFGSI
SiO force fields, are within
1.7%, 2.2%, and 2.3% of the experimental values, respectively. Such
good accuracy attests to the transferability of these force fields to
the coesite bonding environment.

3.4. Stishovite
The crystal structure of stishovite resembles that of rutile
(TiO2), with each of the octahedrally coordinated silicon atoms
bound to six oxygen atoms [44]. Therefore, the transferability of

275

B.J. Cowen, M.S. El-Genk / Computational Materials Science 111 (2016) 269276

Table 3
Comparison of calculated properties (standard deviation shown in parenthesis) and atom charges for a-cristobalite in MD simulations with the BOVC force fields, to experimental
values.
Parameter

a ()
c ()
q (g/cm3)
qavg (O)
qavg (Si)
qrange (O)
qrange (Si)

Experimental/force field
Exp. [41]

COMB10 [11]

H2 O
[12]
ReaxFFSiO

ReaxFFGSI
SiO [13]

4.978
6.948
2.318

4.81 (1e2)
6.13 (1e2)
2.81 (7e3)
1.45 (1e4)
2.90 (3e4)
1.39 to 1.51
2.882.92

5.03 (7e3)
7.11 (1e2)
2.22 (3e3)
0.87 (4e3)
1.73 (7e3)
0.76 to 0.97
1.581.90

5.04 (9e3)
7.12 (1e2)
2.20 (4e3)
0.47 (1e3)
0.94 (3e3)
0.43 to 0.51
0.900.98

Table 4
Comparison of calculated properties (standard deviation shown in parenthesis) and atom charges for coesite, in MD simulations with the BOVC force fields, to experimental
values.
Parameter

a ()
b ()
c ()
q (g/cm3)
b ()
qavg (O)
qavg (Si)
qrange (O)
qrange (Si)

Exp. [43]

7.136
12.369
7.174
2.921
120.34

Force fields
COMB10 [11]

H2 O
[12]
ReaxFFSiO

ReaxFFGSI
SiO [13]

7.04 (1e2)
12.40 (2e2)
7.22 (9e3)
2.97 (6e3)
121.12 (1e1)
1.46 (2e4)
2.91 (4e4)
1.37 to 1.54
2.892.93

7.29 (2e2)
12.33 (1e2)
7.12 (7e3)
2.88 (3e3)
119.75 (8e2)
0.79 (3e3)
1.59 (5e3)
0.71 to 0.88
1.481.76

7.20 (1e2)
12.19 (2e2)
7.01 (1e2)
2.96 (4e3)
119.41 (1e1)
0.46 (1e3)
0.91 (2e3)
0.41 to 0.50
0.870.95

Table 5
Comparison of calculated properties (standard deviation shown in parenthesis) and atom charges for stishovite, in MD simulations with BOVC force fields to experimental values.
Parameter

a ()
c ()
q (g/cm3)
qavg (O)
qavg (Si)
qrange (O)
qrange (Si)

Exp. [44]

4.180
2.666
4.283

Force field
COMB10 [11]

H2 O
[12]
ReaxFFSiO

ReaxFFGSI
SiO [13]

4.09 (7e3)
2.97 (2e3)
4.02 (4e3)
1.47 (1e4)
2.94 (3e4)
1.43 to 1.51
2.932.96

4.56 (4e3)
3.04 (2e3)
3.16 (3e3)
0.53 (2e3)
1.07 (3e3)
0.48 to 0.59
1.021.11

5.68 (4e2)
2.54 (2e2)
2.38 (2e2)
0.41 (3e3)
0.82 (6e3)
0.23 to 0.51
0.600.96

the force fields can be critically evaluated based on their ability to


change from a tetrahedra bonding environment (quartz, cristobalite, coesite) to an octahedra bonding environment (stishovite).
Table 5 compares the calculated properties and charges of the oxygen and silicon atoms of stishovite in the MD simulations with the
three BOVC force fields, with experimental values.
H2 O
The performed MD simulations with ReaxFFSiO
and ReaxFFGSI
SiO
inaccurately predict the properties of the stishovite under equilibration in the NPT ensemble. The calculated properties are within
H2 O
26% and 44% of experimental results for the ReaxFFSiO
and

ReaxFFGSI
SiO , respectively. This could be because the density and the
cohesive energies of various polymorphs were determined with
single point energies at different densities, which is more of an
evaluation than an optimization. Thus, despite the fact that MD
H2 O
simulations with both ReaxFFSiO
and ReaxFFGSI
SiO accurately predict
the properties of the other silica polymorphs, these force fields
are not necessarily suitable for all phases of silica, and need to be
optimized more effectively for stishovite. The MD simulation
results with the COMB10 force field are accurate, to within 11%
of the experimental values for the c lattice parameter, and
2.2% for the a lattice parameter (Table 5). Thus, the COMB10
force field is a more appropriate choice for modeling stishovite.

4. Summary & conclusions


The accuracy and transferability of the three BOVC force fields:
H2 O
COMB10, ReaxFFSiO
, and ReaxFFGSI
SiO , are investigated for modeling
the four SiO2 polymorphs, quartz, cristobalite, coesite, and stishovite. These force fields have been designed for modeling interfaces
and complex chemical reactions. MD simulation results of the
structural properties of crystalline silica, using these force fields,
are compared to experimental values. In these simulations, the
atom charges are updated each time step. The results with updating the atom charges every 10-time steps are inaccurate, since
energy is not conserved.
The MD simulation results with the COMB10 force field accurately predict the properties of the quartz and coesite polymorphs
at ambient conditions, to within 7.3% and 1.7% of the experimental
values, respectively. The simulation results for the a-cristobalite
and stishovite are not as accurate; they are within 21% and 11%
of the experimental values, respectively.
H2 O
The simulation results with ReaxFFSiO
and ReaxFFGSI
SiO , of the
properties for quartz, cristobalite, and coesite at ambient conditions, except the BAD function, are accurate to within 5% of the
H2 O
experimental values. The simulation results with ReaxFFSiO
of

276

B.J. Cowen, M.S. El-Genk / Computational Materials Science 111 (2016) 269276

the BAD function for quartz are within 10% of the experimental
values.
H2 O
The simulation results with ReaxFFSiO
and ReaxFFGSI
SiO
inaccurately predict the properties of the high-pressure stishovite
polymorph; they are within 26% and 44% of the experimental
values, respectively. In addition, the results of the MD simulations
H2 O
with the COMB10 and the ReaxFFSiO
inaccurately model the ab
transition for quartz. The results of the cell volume expansion at
the ab transition are quite different from experimental values
and do not predict the experimentally measured plateau, caused
by the rotation of the SiO4 tetrahedra.

Acknowledgements
The financial support for this research is provided by the Institute for Space and Nuclear Power Studies at the University of New
Mexico (ISNPS-UNM) and by a U.S. Nuclear Regulatory Commission
(NRC) Graduate Fellowship under Grant # NRC-38-09-931 to
ISNPS-UNM.
The authors thank Dr. Susan Atlas of the Center for Advanced
Research Computing (CARC), at the University of New Mexico for
her help and suggestions, acknowledge the valuable support
provided by the CARC staff and are grateful for using the
supercomputer clusters at the center.
References
[1] J. Livage, T. Coradin, C. Roux, J. Phys.-Condens. Matter 13 (2001) R673.
[2] A. Cimas, F. Tielens, M. Sulpizi, M.P. Gaigeot, D. Costa, J. Phys.-Condens. Matter
26 (2014) 244106.
[3] A. Rimola, D. Costa, M. Sodupe, J.F. Lambert, P. Ugliengo, Chem. Rev. 113 (2013)
4216.
[4] M. Schulz, Nature 399 (1999) 729.
[5] C.J. Cochrane, P.M. Lenahan, Appl. Phys. Lett. 103 (2013) 053506.
[6] B.W. van Beest, G.J. Kramer, R.A. van Santen, Phys. Rev. Lett. 64 (1990) 1955.
[7] A. Pedone, G. Malavasi, M.C. Menziani, A.N. Cormack, U. Segre, J. Phys. Chem. B
110 (2006) 11780.
[8] S. Tsuneyuki, M. Tsukada, H. Aoki, Y. Matsui, Phys. Rev. Lett. 61 (1988) 869.
[9] A. Carre, J. Horbach, S. Ispas, W. Kob, EPL-Europhys. Lett. 82 (2008) 17001.
[10] S. Munetoh, T. Motooka, K. Moriguchi, A. Shintani, Comput. Mater. Sci. 39
(2007) 334.

[11] T.R. Shan, B.D. Devine, J.M. Hawkins, A. Asthagiri, S.R. Phillpot, S.B. Sinnott,
Phys. Rev. B 82 (2010) 255.
[12] J.C. Fogarty, H.M. Aktulga, A.Y. Grama, A.C.T. van Duin, S.A. Pandit, J. Chem.
Phys. 132 (2010) 174704.
[13] A.D. Kulkarni, D.G. Truhlar, S.G. Srinivasan, A.C.T. van Duin, P. Norman, T.E.
Schwartzentruber, J. Phys. Chem. C 117 (2013) 258.
[14] J.G. Yu, S.B. Sinnott, S.R. Phillpot, Phys. Rev. B 75 (2007) 085311.
[15] A.C.T. van Duin, S. Dasgupta, F. Lorant, W.A. Goddard, J. Phys. Chem. A 105
(2001) 9396.
[16] T.R. Shan, B.D. Devine, T.W. Kemper, S.B. Sinnott, S.R. Phillpot, Phys. Rev. B 81
(2010) 125328.
[17] T.R. Shan, B.D. Devine, S.R. Phillpot, S.B. Sinnott, Phys. Rev. B 83 (2011) 115327.
[18] M.J. Noordhoek, T. Liang, Z.Z. Lu, T.R. Shan, S.B. Sinnott, S.R. Phillpot, J. Nucl.
Mater. 441 (2013) 274.
[19] W.X. Song, S.J. Zhao, J. Mater. Res. 28 (2013) 1155.
[20] D. Raymand, A.C.T. van Duin, M. Baudin, K. Hermansson, Surf. Sci. 602 (2008)
1020.
[21] A.C.T. van Duin, A. Strachan, S. Stewman, Q.S. Zhang, X. Xu, W.A. Goddard, J.
Phys. Chem. A 107 (2003) 3803.
[22] A. Yasukawa, JSME Int. J. A-Mech. M 39 (1996) 313.
[23] T.T. Trinh, A.P.J. Jansen, R.A. van Santen, J. Phys. Chem. B 110 (2006) 23099.
[24] B.J. Cowen, M.S. El-Genk, Comput. Mater. Sci. 107 (2015) 88.
[25] S. Grazulis, D. Chateigner, R.T. Downs, A.F.T. Yokochi, M. Quiros, L. Lutterotti, E.
Manakova, J. Butkus, P. Moeck, A. Le Bail, J. Appl. Crystallogr. 42 (2009) 726.
[26] S. Grazulis, A. Daskevic, A. Merkys, D. Chateigner, L. Lutterotti, M. Quiros, N.R.
Serebryanaya, P. Moeck, R.T. Downs, A. Le Bail, Nucleic Acids Res. 40 (2012)
D420.
[27] R.T. Downs, M. Hall-Wallace, Am. Mineral. 88 (2003) 247.
[28] S. Plimpton, J. Comput. Phys. 117 (1995) 1.
[29] H.M. Aktulga, J.C. Fogarty, S.A. Pandit, A.Y. Grama, Parallel Comput. 38 (2012)
245.
[30] A.K. Rapp, W.A. Goddard, J. Phys. Chem.-Us 95 (1991) 3358.
[31] E. Polak, G. Ribiere, Rev. Fr. Inform. Rech. Oper. 3 (1969) 35.
[32] Y.K. Shin, T.R. Shan, T. Liang, M.J. Noordhoek, S.B. Sinnott, A.C.T. van Duin, S.R.
Phillpot, MRS Bull. 37 (2012) 504.
[33] I. Tajima, O. Asami, E. Sugiura, Anal. Chim. Acta 365 (1998) 147.
[34] J. Fontanel, C. Andeen, D. Schuele, J. Appl. Phys. 45 (1974) 2852.
[35] E.K. Chang, M. Rohlfing, S.G. Louie, Phys. Rev. Lett. 85 (2000) 2613.
[36] L. Levien, C.T. Prewitt, D.J. Weidner, Am. Mineral. 65 (1980) 920.
[37] M.A. Carpenter, E.K.H. Salje, A. Graeme-Barber, B. Wruck, M.T. Dove, K.S.
Knight, Am. Mineral. 83 (1998) 2.
[38] Y.H. Kim, M.S. Hwang, H.J. Kim, J.Y. Kim, Y. Lee, J. Appl. Phys. 90 (2001) 3367.
[39] D. Herzbach, K. Binder, M.H. Mser, J. Chem. Phys. 123 (2005) 124711.
[40] P. Tangney, S. Scandolo, J. Chem. Phys. 117 (2002) 8898.
[41] D.R. Peacor, Z. Kristallogr. 138 (1973) 274.
[42] E.C.T. Chao, E.M. Shoemaker, B.M. Madsen, Science 132 (1960) 220.
[43] L. Levien, C.T. Prewitt, Am. Mineral. 66 (1981) 324.
[44] L.G. Liu, W.A. Bassett, T. Takahashi, J. Geophys. Res. 79 (1974) 1160.

Vous aimerez peut-être aussi