9 |
|
\usepackage{longtable} |
10 |
|
\pagestyle{plain} |
11 |
|
\pagenumbering{arabic} |
12 |
+ |
\usepackage{floatrow} |
13 |
|
\oddsidemargin 0.0cm |
14 |
|
\evensidemargin 0.0cm |
15 |
|
\topmargin -21pt |
21 |
|
\usepackage[square, comma, sort&compress]{natbib} |
22 |
|
\bibpunct{[}{]}{,}{n}{}{;} |
23 |
|
|
24 |
+ |
\DeclareFloatFont{tiny}{\scriptsize}% "scriptsize" is defined by floatrow, "tiny" not |
25 |
+ |
\floatsetup[table]{font=tiny} |
26 |
|
|
27 |
+ |
|
28 |
|
%\renewcommand\citemid{\ } % no comma in optional reference note |
29 |
< |
\lstset{language=C,frame=TB,basicstyle=\footnotesize,basicstyle=\ttfamily, % |
29 |
> |
\lstset{language=C,frame=TB,basicstyle=\footnotesize\ttfamily, % |
30 |
|
xleftmargin=0.25in, xrightmargin=0.25in,captionpos=b, % |
31 |
|
abovecaptionskip=0.5cm, belowcaptionskip=0.5cm, escapeinside={~}{~}} |
32 |
|
\renewcommand{\lstlistingname}{Scheme} |
51 |
|
\newcolumntype{M}{p{1.55in}} |
52 |
|
|
53 |
|
|
54 |
< |
\title{{\sc OpenMD-2.1}: Molecular Dynamics in the Open} |
54 |
> |
\title{{\sc OpenMD-2.2}: Molecular Dynamics in the Open} |
55 |
|
|
56 |
|
\author{Joseph Michalka, James Marr, Kelsey Stocker, Madan Lamichhane, |
57 |
|
Patrick Louden, \\ |
843 |
|
|
844 |
|
\section{\label{section:frcFile}Force Field Files} |
845 |
|
|
846 |
< |
Force field files have a number of ``Blocks'' to demarkate different |
846 |
> |
Force field files have a number of ``Blocks'' to delineate different |
847 |
|
types of information. The blocks contain AtomType data, which provide |
848 |
|
properties belonging to a single AtomType, as well as interaction |
849 |
|
information which provides information about bonded or non-bonded |
850 |
|
interactions that cannot be deduced from AtomType information alone. |
851 |
|
A simple example of a forceField file is shown in scheme |
852 |
< |
\ref{sch:frcExample}. |
852 |
> |
\ref{sch:frcExample}. |
853 |
|
|
854 |
|
\begin{lstlisting}[float,caption={[An example of a complete OpenMD |
855 |
|
force field file for straight-chain united-atom alkanes.] An example |
856 |
|
showing a complete OpenMD force field for straight-chain united-atom |
857 |
|
alkanes.}, label={sch:frcExample}] |
858 |
|
begin Options |
859 |
< |
Name = "alkane" end |
860 |
< |
Options |
859 |
> |
Name = "alkane" |
860 |
> |
end Options |
861 |
|
|
862 |
|
begin BaseAtomTypes |
863 |
|
//name mass |
963 |
|
ability to print out the names of the base atom types for displaying |
964 |
|
simulations in Jmol or VMD. |
965 |
|
|
966 |
< |
\begin{lstlisting}[caption={[A simple example of a BaseAtomType |
967 |
< |
block.] A simple example of a BaseAtomType block.}, |
966 |
> |
\begin{lstlisting}[caption={[A simple example of a BaseAtomTypes |
967 |
> |
block.] A simple example of a BaseAtomTypes block.}, |
968 |
|
label={sch:baseAtomTypesBlock}] |
969 |
|
begin BaseAtomTypes |
970 |
|
//Name mass (amu) |
993 |
|
from the oxygen base type. |
994 |
|
|
995 |
|
\begin{lstlisting}[caption={[An example of a AtomTypes block.] A |
996 |
< |
simple example of an AtomType block which |
996 |
> |
simple example of an AtomTypes block which |
997 |
|
shows how multiple types can inherit from the same base type.}, |
998 |
|
label={sch:atomTypesBlock}] |
999 |
|
begin AtomTypes |
1022 |
|
\subsection{\label{section:ffDirectionalAtom}The DirectionalAtomTypes |
1023 |
|
block} |
1024 |
|
DirectionalAtoms have orientational degrees of freedom as well as |
1025 |
< |
translation, so they have moment of inertia tensors. |
1025 |
> |
translation, so moving these atoms requires information about the |
1026 |
> |
moments of inertias in the same way that translational motion requires |
1027 |
> |
mass. For DirectionalAtoms, OpenMD treats the mass distribution with |
1028 |
> |
higher priority than electrostatic distributions; the moment of |
1029 |
> |
inertia tensor, $\overleftrightarrow{\mathsf I}$, should be |
1030 |
> |
diagonalized to obtain body-fixed axes, and the three diagonal moments |
1031 |
> |
should correspond to rotational motion \textit{around} each of these |
1032 |
> |
body-fixed axes. Charge distributions may then result in dipole |
1033 |
> |
vectors that are oriented along a linear combination of the body-axes, |
1034 |
> |
and in quadrupole tensors that are not necessarily diagonal in the |
1035 |
> |
body frame. |
1036 |
|
|
1037 |
|
\begin{lstlisting}[caption={[An example of a DirectionalAtomTypes block.] A |
1038 |
|
simple example of a DirectionalAtomTypes block.}, |
1040 |
|
begin DirectionalAtomTypes |
1041 |
|
//Name I_xx I_yy I_zz (All moments in (amu*Ang^2) |
1042 |
|
SSD 1.7696 0.6145 1.1550 |
1029 |
– |
SSD_E 1.7696 0.6145 1.1550 |
1043 |
|
GBC6H6 88.781 88.781 177.561 |
1044 |
|
GBCH3OH 4.056 20.258 20.999 |
1045 |
|
GBH2O 1.777 0.581 1.196 |
1046 |
+ |
CO2 43.06 43.06 0.0 // single-site model for CO2 |
1047 |
|
end DirectionalAtomTypes |
1048 |
|
|
1049 |
|
\end{lstlisting} |
1050 |
|
|
1051 |
+ |
For a DirectionalAtom that represents a linear object, it is |
1052 |
+ |
appropriate for one of the moments of inertia to be zero. In this |
1053 |
+ |
case, OpenMD identifies that DirectionalAtom as having only 5 degrees |
1054 |
+ |
of freedom (three translations and two rotations), and will alter |
1055 |
+ |
calculation of temperatures to reflect this. |
1056 |
|
|
1057 |
|
\subsection{\label{section::ffAtomProperties}AtomType properties} |
1058 |
|
\subsubsection{\label{section:ffLJ}The LennardJonesAtomTypes block} |
1059 |
< |
The most basic interatomic interaction implemented in {\sc OpenMD} is |
1060 |
< |
the Lennard-Jones potential, which mimics the van der Waals |
1061 |
< |
interaction at long distances and uses an empirical repulsion at short |
1062 |
< |
distances. The Lennard-Jones potential is given by: |
1059 |
> |
One of the most basic interatomic interactions implemented in {\sc |
1060 |
> |
OpenMD} is the Lennard-Jones potential, which mimics the van der |
1061 |
> |
Waals interaction at long distances and uses an empirical repulsion at |
1062 |
> |
short distances. The Lennard-Jones potential is given by: |
1063 |
|
\begin{equation} |
1064 |
|
V_{\text{LJ}}(r_{ij}) = |
1065 |
|
4\epsilon_{ij} \biggl[ |
1074 |
|
|
1075 |
|
Interactions between dissimilar particles requires the generation of |
1076 |
|
cross term parameters for $\sigma$ and $\epsilon$. These parameters |
1077 |
< |
are determined using the Lorentz-Berthelot mixing |
1077 |
> |
are usually determined using the Lorentz-Berthelot mixing |
1078 |
|
rules:\cite{Allen87} |
1079 |
|
\begin{equation} |
1080 |
|
\sigma_{ij} = \frac{1}{2}[\sigma_{ii} + \sigma_{jj}], |
1086 |
|
\label{eq:epsilonMix} |
1087 |
|
\end{equation} |
1088 |
|
|
1089 |
< |
\subsubsection{\label{section:ffCharge}The ChargeAtomTypes block} |
1090 |
< |
\subsubsection{\label{section:ffMultipole}The MultipoleAtomTypes block} |
1091 |
< |
The dipole-dipole potential has the following form: |
1089 |
> |
The values of $\sigma_{ii}$ and $\epsilon_{ii}$ are properties of atom |
1090 |
> |
type $i$, and must be specified in a section of the force field file |
1091 |
> |
called the {\tt LennardJonesAtomTypes} block (see listing |
1092 |
> |
\ref{sch:LJatomTypesBlock}). Separate Lennard-Jones interactions |
1093 |
> |
which are not determined by the mixing rules can also be specified in |
1094 |
> |
the {\tt NonbondedInteractionTypes} block (see section |
1095 |
> |
\ref{section:ffNBinteraction}). |
1096 |
> |
|
1097 |
> |
\begin{lstlisting}[caption={[An example of a LennardJonesAtomTypes block.] A |
1098 |
> |
simple example of a LennardJonesAtomTypee block. Units for |
1099 |
> |
$\epsilon$ are kcal / mol and for $\sigma$ are \AA\ .}, |
1100 |
> |
label={sch:LJatomTypesBlock}] |
1101 |
> |
begin LennardJonesAtomTypes |
1102 |
> |
//Name epsilon sigma |
1103 |
> |
O_TIP4P 0.1550 3.15365 |
1104 |
> |
O_TIP4P-Ew 0.16275 3.16435 |
1105 |
> |
O_TIP5P 0.16 3.12 |
1106 |
> |
O_TIP5P-E 0.178 3.097 |
1107 |
> |
O_SPCE 0.15532 3.16549 |
1108 |
> |
O_SPC 0.15532 3.16549 |
1109 |
> |
CH4 0.279 3.73 |
1110 |
> |
CH3 0.185 3.75 |
1111 |
> |
CH2 0.0866 3.95 |
1112 |
> |
CH 0.0189 4.68 |
1113 |
> |
end LennardJonesAtomTypes |
1114 |
> |
\end{lstlisting} |
1115 |
> |
|
1116 |
> |
\subsubsection{\label{section:ffCharge}The ChargeAtomTypes block} |
1117 |
> |
|
1118 |
> |
In molecular simulations, proper accumulation of the electrostatic |
1119 |
> |
interactions is essential and is one of the most |
1120 |
> |
computationally-demanding tasks. Most common molecular mechanics |
1121 |
> |
force fields represent atomic sites with full or partial charges |
1122 |
> |
protected by Lennard-Jones (short range) interactions. Partial charge |
1123 |
> |
values, $q_i$ are empirical representations of the distribution of |
1124 |
> |
electronic charge in a molecule. This means that nearly every pair |
1125 |
> |
interaction involves a calculation of charge-charge forces. Coupled |
1126 |
> |
with relatively long-ranged $r^{-1}$ decay, the monopole interactions |
1127 |
> |
quickly become the most expensive part of molecular simulations. The |
1128 |
> |
interactions between point charges can be handled via a number of |
1129 |
> |
different algorithms, but Coulomb's law is the fundamental physical |
1130 |
> |
principle governing these interactions, |
1131 |
|
\begin{equation} |
1132 |
+ |
V_{\text{charge}}(r_{ij}) = \sum_{ij}\frac{q_iq_je^2}{4 \pi \epsilon_0 |
1133 |
+ |
r_{ij}}, |
1134 |
+ |
\end{equation} |
1135 |
+ |
where $q$ represents the charge on particle $i$ or $j$, and $e$ is the |
1136 |
+ |
charge of an electron in Coulombs. $\epsilon_0$ is the permittivity |
1137 |
+ |
of free space. |
1138 |
+ |
|
1139 |
+ |
\begin{lstlisting}[caption={[An example of a ChargeAtomTypes block.] A |
1140 |
+ |
simple example of a ChargeAtomTypes block. Units for |
1141 |
+ |
charge are in multiples of electron charge.}, |
1142 |
+ |
label={sch:ChargeAtomTypesBlock}] |
1143 |
+ |
begin ChargeAtomTypes |
1144 |
+ |
// Name charge |
1145 |
+ |
O_TIP3P -0.834 |
1146 |
+ |
O_SPCE -0.8476 |
1147 |
+ |
H_TIP3P 0.417 |
1148 |
+ |
H_TIP4P 0.520 |
1149 |
+ |
H_SPCE 0.4238 |
1150 |
+ |
EP_TIP4P -1.040 |
1151 |
+ |
Na+ 1.0 |
1152 |
+ |
Cl- -1.0 |
1153 |
+ |
end ChargeAtomTypes |
1154 |
+ |
\end{lstlisting} |
1155 |
+ |
|
1156 |
+ |
\subsubsection{\label{section:ffMultipole}The MultipoleAtomTypes |
1157 |
+ |
block} |
1158 |
+ |
For complex charge distributions that are centered on single sites, it |
1159 |
+ |
is convenient to write the total electrostatic potential in terms of |
1160 |
+ |
multipole moments, |
1161 |
+ |
\begin{equation} |
1162 |
+ |
U_{\bf{ab}}(r)=\hat{M}_{\bf a} \hat{M}_{\bf b} \frac{1}{r} \label{kernel}. |
1163 |
+ |
\end{equation} |
1164 |
+ |
where the multipole operator on site $\bf a$, |
1165 |
+ |
\begin{equation} |
1166 |
+ |
\hat{M}_{\bf a} = C_{\bf a} - D_{{\bf a}\alpha} \frac{\partial}{\partial r_{\alpha}} |
1167 |
+ |
+ Q_{{\bf a}\alpha\beta} |
1168 |
+ |
\frac{\partial^2}{\partial r_{\alpha} \partial r_{\beta}} + \dots |
1169 |
+ |
\end{equation} |
1170 |
+ |
Here, the point charge, dipole, and quadrupole for site $\bf a$ are |
1171 |
+ |
given by $C_{\bf a}$, $D_{{\bf a}\alpha}$, and $Q_{{\bf |
1172 |
+ |
a}\alpha\beta}$, respectively. These are the primitive |
1173 |
+ |
multipoles. If the site is representing a distribution of charges, |
1174 |
+ |
these can be expressed as, |
1175 |
+ |
\begin{align} |
1176 |
+ |
C_{\bf a} =&\sum_{k \, \text{in \bf a}} q_k , \label{eq:charge} \\ |
1177 |
+ |
D_{{\bf a}\alpha} =&\sum_{k \, \text{in \bf a}} q_k r_{k\alpha}, \label{eq:dipole}\\ |
1178 |
+ |
Q_{{\bf a}\alpha\beta} =& \frac{1}{2} \sum_{k \, \text{in \bf a}} q_k |
1179 |
+ |
r_{k\alpha} r_{k\beta} . \label{eq:quadrupole} |
1180 |
+ |
\end{align} |
1181 |
+ |
Note that the definition of the primitive quadrupole here differs from |
1182 |
+ |
the standard traceless form, and contains an additional Taylor-series |
1183 |
+ |
based factor of $1/2$. |
1184 |
+ |
|
1185 |
+ |
The details of the multipolar interactions will be given later, but |
1186 |
+ |
many readers are familiar with the dipole-dipole potential: |
1187 |
+ |
\begin{equation} |
1188 |
|
V_{\text{dipole}}(\mathbf{r}_{ij},\boldsymbol{\Omega}_{i}, |
1189 |
< |
\boldsymbol{\Omega}_{j}) = \frac{|\mu_i||\mu_j|}{4\pi\epsilon_{0}r_{ij}^{3}} \biggl[ |
1189 |
> |
\boldsymbol{\Omega}_{j}) = \frac{|{\bf D}_i||{\bf D}_j|}{4\pi\epsilon_{0}r_{ij}^{3}} \biggl[ |
1190 |
|
\boldsymbol{\hat{u}}_{i} \cdot \boldsymbol{\hat{u}}_{j} |
1191 |
|
- |
1192 |
|
3(\boldsymbol{\hat{u}}_i \cdot \hat{\mathbf{r}}_{ij}) % |
1196 |
|
Here $\mathbf{r}_{ij}$ is the vector starting at atom $i$ pointing |
1197 |
|
towards $j$, and $\boldsymbol{\Omega}_i$ and $\boldsymbol{\Omega}_j$ |
1198 |
|
are the orientational degrees of freedom for atoms $i$ and $j$ |
1199 |
< |
respectively. The magnitude of the dipole moment of atom $i$ is |
1200 |
< |
$|\mu_i|$, $\boldsymbol{\hat{u}}_i$ is the standard unit orientation |
1199 |
> |
respectively. The magnitude of the dipole moment of atom $i$ is $|{\bf |
1200 |
> |
D}_i|$, $\boldsymbol{\hat{u}}_i$ is the standard unit orientation |
1201 |
|
vector of $\boldsymbol{\Omega}_i$, and $\boldsymbol{\hat{r}}_{ij}$ is |
1202 |
|
the unit vector pointing along $\mathbf{r}_{ij}$ |
1203 |
|
($\boldsymbol{\hat{r}}_{ij}=\mathbf{r}_{ij}/|\mathbf{r}_{ij}|$). |
1204 |
|
|
1205 |
+ |
|
1206 |
+ |
\begin{lstlisting}[caption={[An example of a MultipoleAtomTypes block.] A |
1207 |
+ |
simple example of a MultipoleAtomTypes block. Dipoles are given in |
1208 |
+ |
units of Debyes, and Quadrupole moments are given in units of Debye |
1209 |
+ |
\AA~(or $10^{-26} \mathrm{~esu~cm}^2$)}, |
1210 |
+ |
label={sch:MultipoleAtomTypesBlock}] |
1211 |
+ |
begin MultipoleAtomTypes |
1212 |
+ |
// Euler angles are given in zxz convention in units of degrees. |
1213 |
+ |
// |
1214 |
+ |
// point dipoles: |
1215 |
+ |
// name d phi theta psi dipole_moment |
1216 |
+ |
DIP d 0.0 0.0 0.0 1.91 // dipole points along z-body axis |
1217 |
+ |
// |
1218 |
+ |
// point quadrupoles: |
1219 |
+ |
// name q phi theta psi Qxx Qyy Qzz |
1220 |
+ |
CO2 q 0.0 0.0 0.0 0.0 0.0 -0.430592 //quadrupole tensor has zz element |
1221 |
+ |
// |
1222 |
+ |
// Atoms with both dipole and quadrupole moments: |
1223 |
+ |
// name dq phi theta psi dipole_moment Qxx Qyy Qzz |
1224 |
+ |
SSD dq 0.0 0.0 0.0 2.35 -1.682 1.762 -0.08 |
1225 |
+ |
end MultipoleAtomTypes |
1226 |
+ |
\end{lstlisting} |
1227 |
+ |
|
1228 |
+ |
Specifying a MultipoleAtomType requires declaring how the |
1229 |
+ |
electrostatic frame for the site is rotated relative to the body-fixed |
1230 |
+ |
axes for that atom. The Euler angles $(\phi, \theta, \psi)$ for this |
1231 |
+ |
rotation must be given, and then the dipole, quadrupole, or all of |
1232 |
+ |
these moments are specified in the electrostatic frame. In OpenMD, |
1233 |
+ |
the Euler angles are specified in the $zxz$ convention and are entered |
1234 |
+ |
in units of degrees. Dipole moments are entered in units of Debye, |
1235 |
+ |
and Quadrupole moments in units of Debye \AA. |
1236 |
+ |
|
1237 |
|
\subsubsection{\label{section:ffGB}The FluctuatingChargeAtomTypes block} |
1238 |
|
\subsubsection{\label{section:ffPol}The PolarizableAtomTypes block} |
1239 |
|
\subsubsection{\label{section:ffGB}The GayBerneAtomTypes block} |
1240 |
+ |
|
1241 |
+ |
The Gay-Berne potential has been widely used in the liquid crystal |
1242 |
+ |
community to describe this anisotropic phase |
1243 |
+ |
behavior.~\cite{Gay:1981yu,Berne:1972pb,Kushick:1976xy,Luckhurst:1990fy,Cleaver:1996rt} |
1244 |
+ |
The form of the Gay-Berne potential implemented in OpenMD was |
1245 |
+ |
generalized by Cleaver {\it et al.} and is appropriate for dissimilar |
1246 |
+ |
uniaxial ellipsoids.\cite{Cleaver:1996rt} The potential is constructed in the |
1247 |
+ |
familiar form of the Lennard-Jones function using |
1248 |
+ |
orientation-dependent $\sigma$ and $\epsilon$ parameters, |
1249 |
+ |
\begin{equation*} |
1250 |
+ |
V_{ij}({{\bf \hat u}_i}, {{\bf \hat u}_j}, {{\bf \hat |
1251 |
+ |
r}_{ij}}) = 4\epsilon ({{\bf \hat u}_i}, {{\bf \hat u}_j}, |
1252 |
+ |
{{\bf \hat r}_{ij}})\left[\left(\frac{\sigma_0}{r_{ij}-\sigma({{\bf \hat u |
1253 |
+ |
}_i}, |
1254 |
+ |
{{\bf \hat u}_j}, {{\bf \hat r}_{ij}})+\sigma_0}\right)^{12} |
1255 |
+ |
-\left(\frac{\sigma_0}{r_{ij}-\sigma({{\bf \hat u}_i}, {{\bf \hat u}_j}, |
1256 |
+ |
{{\bf \hat r}_{ij}})+\sigma_0}\right)^6\right] |
1257 |
+ |
\label{eq:gb} |
1258 |
+ |
\end{equation*} |
1259 |
+ |
|
1260 |
+ |
The range $(\sigma({\bf \hat{u}}_{i},{\bf \hat{u}}_{j},{\bf |
1261 |
+ |
\hat{r}}_{ij}))$, and strength $(\epsilon({\bf \hat{u}}_{i},{\bf |
1262 |
+ |
\hat{u}}_{j},{\bf \hat{r}}_{ij}))$ parameters |
1263 |
+ |
are dependent on the relative orientations of the two ellipsoids (${\bf |
1264 |
+ |
\hat{u}}_{i},{\bf \hat{u}}_{j}$) as well as the direction of the |
1265 |
+ |
inter-ellipsoid separation (${\bf \hat{r}}_{ij}$). The shape and |
1266 |
+ |
attractiveness of each ellipsoid is governed by a relatively small set |
1267 |
+ |
of parameters: |
1268 |
+ |
\begin{itemize} |
1269 |
+ |
\item $d$: range parameter for the side-by-side (S) and cross (X) configurations |
1270 |
+ |
\item $l$: range parameter for the end-to-end (E) configuration |
1271 |
+ |
\item $\epsilon_X$: well-depth parameter for the cross (X) configuration |
1272 |
+ |
\item $\epsilon_S$: well-depth parameter for the side-by-side (S) configuration |
1273 |
+ |
\item $\epsilon_E$: well depth parameter for the end-to-end (E) configuration |
1274 |
+ |
\item $dw$: The ``softness'' of the potential |
1275 |
+ |
\end{itemize} |
1276 |
+ |
Additionally, there are two universal paramters to govern the overall |
1277 |
+ |
importance of the purely orientational ($\nu$) and the mixed |
1278 |
+ |
orientational / translational ($\mu$) parts of strength of the |
1279 |
+ |
interactions. These parameters have default or ``canonical'' values, |
1280 |
+ |
but may be changed as a force field option: |
1281 |
+ |
\begin{itemize} |
1282 |
+ |
\item $\nu$: purely orientational part : defaults to 1 |
1283 |
+ |
\item $\mu$: mixed orientational / translational part : defaults to |
1284 |
+ |
2 |
1285 |
+ |
\end{itemize} |
1286 |
+ |
Further details of the potential are given |
1287 |
+ |
elsewhere,\cite{Luckhurst:1990fy,Golubkov06,SunX._jp0762020} and an |
1288 |
+ |
excellent overview of the computational methods that can be used to |
1289 |
+ |
efficiently compute forces and torques for this potential can be found |
1290 |
+ |
in Ref. \citealp{Golubkov06} |
1291 |
+ |
|
1292 |
+ |
\begin{lstlisting}[caption={[An example of a GayBerneAtomTypes block.] A |
1293 |
+ |
simple example of a GayBerneAtomTypes block. Distances ($d$ and $l$) |
1294 |
+ |
are given in \AA\ and energies ($\epsilon_X, \epsilon_S, \epsilon_E$) |
1295 |
+ |
are in units of kcal/mol. $dw$ is unitless.}, |
1296 |
+ |
label={sch:GayBerneAtomTypes}] |
1297 |
+ |
begin GayBerneAtomTypes |
1298 |
+ |
//Name d l eps_X eps_S eps_E dw |
1299 |
+ |
GBlinear 2.8104 9.993 0.774729 0.774729 0.116839 1.0 |
1300 |
+ |
GBC6H6 4.65 2.03 0.540 0.540 1.9818 0.6 |
1301 |
+ |
GBCH3OH 2.55 3.18 0.542 0.542 0.55826 1.0 |
1302 |
+ |
end GayBerneAtomTypes |
1303 |
+ |
\end{lstlisting} |
1304 |
+ |
|
1305 |
|
\subsubsection{\label{section:ffSticky}The StickyAtomTypes block} |
1306 |
|
|
1307 |
< |
One of the solvents used by {\sc OpenMD} is the extended Soft Sticky |
1308 |
< |
Dipole (SSD/E) water model.\cite{fennell04} The original SSD was |
1309 |
< |
developed by Ichiye \emph{et al.}\cite{liu96:new_model} as a modified |
1310 |
< |
form of the hard-sphere water model proposed by Bratko, Blum, and |
1307 |
> |
One of the solvents that can be simulated by {\sc OpenMD} is the |
1308 |
> |
extended Soft Sticky Dipole (SSD/E) water model.\cite{fennell04} The |
1309 |
> |
original SSD was developed by Ichiye \emph{et |
1310 |
> |
al.}\cite{liu96:new_model} as a modified form of the hard-sphere |
1311 |
> |
water model proposed by Bratko, Blum, and |
1312 |
|
Luzar.\cite{Bratko85,Bratko95} It consists of a single point dipole |
1313 |
|
with a Lennard-Jones core and a sticky potential that directs the |
1314 |
|
particles to assume the proper hydrogen bond orientation in the first |
1391 |
|
|
1392 |
|
Recent constant pressure simulations revealed issues in the original |
1393 |
|
SSD model that led to lower than expected densities at all target |
1394 |
< |
pressures.\cite{Ichiye03,fennell04} The default model in {\sc OpenMD} |
1395 |
< |
is therefore SSD/E, a density corrected derivative of SSD that |
1396 |
< |
exhibits improved liquid structure and transport behavior. If the use |
1397 |
< |
of a reaction field long-range interaction correction is desired, it |
1398 |
< |
is recommended that the parameters be modified to those of the SSD/RF |
1399 |
< |
model (an SSD variant parameterized for reaction field). These solvent |
1188 |
< |
parameters are listed and can be easily modified in the {\sc duff} |
1189 |
< |
force field file ({\tt DUFF.frc}). A table of the parameter values |
1190 |
< |
and the drawbacks and benefits of the different density corrected SSD |
1191 |
< |
models can be found in reference~\cite{fennell04}. |
1394 |
> |
pressures,\cite{Ichiye03,fennell04} so variants on the sticky |
1395 |
> |
potential can be specified by using one of a number of substitute atom |
1396 |
> |
types (see listing \ref{sch:StickyAtomTypes}). A table of the |
1397 |
> |
parameter values and the drawbacks and benefits of the different |
1398 |
> |
density corrected SSD models can be found in |
1399 |
> |
reference~\citealp{fennell04}. |
1400 |
|
|
1401 |
< |
\subsection{\label{section::ffMetals}Metallic Atom Types} |
1402 |
< |
\subsubsection{\label{section:ffEAM}The EAMAtomTypes block} |
1403 |
< |
{\sc OpenMD} implements a potential that describes bonding in |
1404 |
< |
transition metal |
1405 |
< |
systems.~\cite{Finnis84,Ercolessi88,Chen90,Qi99,Ercolessi02} This |
1406 |
< |
potential has an attractive interaction which models ``Embedding'' a |
1407 |
< |
positively charged pseudo-atom core in the electron density due to the |
1408 |
< |
free valance ``sea'' of electrons created by the surrounding atoms in |
1409 |
< |
the system. A pairwise part of the potential (which is primarily |
1410 |
< |
repulsive) describes the interaction of the positively charged metal |
1411 |
< |
core ions with one another. The Embedded Atom Method ({\sc |
1412 |
< |
eam})~\cite{Daw84,FBD86,johnson89,Lu97} has been widely adopted in the |
1413 |
< |
materials science community and has been included in {\sc OpenMD}. A |
1206 |
< |
good review of {\sc eam} and other formulations of metallic potentials |
1207 |
< |
was given by Voter.\cite{Voter:95} |
1401 |
> |
\begin{lstlisting}[caption={[An example of a StickyAtomTypes block.] A |
1402 |
> |
simple example of a StickyAtomTypes block. Distances ($r_l$, $r_u$, |
1403 |
> |
$r_{l}'$ and $r_{u}'$) are given in \AA\ and energies ($v_0, v_{0}'$) |
1404 |
> |
are in units of kcal/mol. $w_0$ is unitless.}, |
1405 |
> |
label={sch:StickyAtomTypes}] |
1406 |
> |
begin StickyAtomTypes |
1407 |
> |
//name w0 v0 (kcal/mol) v0p rl (Ang) ru rlp rup |
1408 |
> |
SSD_E 0.07715 3.90 3.90 2.40 3.80 2.75 3.35 |
1409 |
> |
SSD_RF 0.07715 3.90 3.90 2.40 3.80 2.75 3.35 |
1410 |
> |
SSD 0.07715 3.7284 3.7284 2.75 3.35 2.75 4.0 |
1411 |
> |
SSD1 0.07715 3.6613 3.6613 2.75 3.35 2.75 4.0 |
1412 |
> |
end StickyAtomTypes |
1413 |
> |
\end{lstlisting} |
1414 |
|
|
1415 |
< |
The {\sc eam} potential has the form: |
1415 |
> |
\subsection{\label{section::ffMetals}Metallic Atom Types} |
1416 |
> |
|
1417 |
> |
{\sc OpenMD} implements a number of related potentials that describe |
1418 |
> |
bonding in transition metals. These potentials have an attractive |
1419 |
> |
interaction which models ``Embedding'' a positively charged |
1420 |
> |
pseudo-atom core in the electron density due to the free valance |
1421 |
> |
``sea'' of electrons created by the surrounding atoms in the system. |
1422 |
> |
A pairwise part of the potential (which is primarily repulsive) |
1423 |
> |
describes the interaction of the positively charged metal core ions |
1424 |
> |
with one another. These potentials have the form: |
1425 |
|
\begin{equation} |
1426 |
|
V = \sum_{i} F_{i}\left[\rho_{i}\right] + \sum_{i} \sum_{j \neq i} |
1427 |
|
\phi_{ij}({\bf r}_{ij}) |
1438 |
|
transition metal potentials require two loops through the atom pairs |
1439 |
|
to compute the inter-atomic forces. |
1440 |
|
|
1441 |
< |
The pairwise portion of the potential, $\phi_{ij}$, is a primarily |
1442 |
< |
repulsive interaction between atoms $i$ and $j$. In the original |
1228 |
< |
formulation of {\sc eam}\cite{Daw84}, $\phi_{ij}$ was an entirely |
1229 |
< |
repulsive term; however later refinements to {\sc eam} allowed for |
1230 |
< |
more general forms for $\phi$.\cite{Daw89} The effective cutoff |
1231 |
< |
distance, $r_{{\text cut}}$ is the distance at which the values of |
1232 |
< |
$f(r)$ and $\phi(r)$ drop to zero for all atoms present in the |
1233 |
< |
simulation. In practice, this distance is fairly small, limiting the |
1234 |
< |
summations in the {\sc eam} equation to the few dozen atoms |
1235 |
< |
surrounding atom $i$ for both the density $\rho$ and pairwise $\phi$ |
1236 |
< |
interactions. |
1441 |
> |
The pairwise portion of the potential, $\phi_{ij}$, is usually a |
1442 |
> |
repulsive interaction between atoms $i$ and $j$. |
1443 |
|
|
1444 |
< |
In computing forces for alloys, mixing rules as outlined by |
1445 |
< |
Johnson~\cite{johnson89} are used to compute the heterogenous pair |
1446 |
< |
potential, |
1444 |
> |
\subsubsection{\label{section:ffEAM}The EAMAtomTypes block} |
1445 |
> |
The Embedded Atom Method ({\sc eam}) is one of the most widely-used |
1446 |
> |
potentials for transition |
1447 |
> |
metals.~\cite{Finnis84,Ercolessi88,Chen90,Qi99,Ercolessi02,Daw84,FBD86,johnson89,Lu97} |
1448 |
> |
It has been widely adopted in the materials science community and a |
1449 |
> |
good review of {\sc eam} and other formulations of metallic potentials |
1450 |
> |
was given by Voter.\cite{Voter:95} |
1451 |
> |
|
1452 |
> |
In the original formulation of {\sc eam}\cite{Daw84}, the pair |
1453 |
> |
potential, $\phi_{ij}$ was an entirely repulsive term; however later |
1454 |
> |
refinements to {\sc eam} allowed for more general forms for |
1455 |
> |
$\phi$.\cite{Daw89} The effective cutoff distance, $r_{{\text cut}}$ |
1456 |
> |
is the distance at which the values of $f(r)$ and $\phi(r)$ drop to |
1457 |
> |
zero for all atoms present in the simulation. In practice, this |
1458 |
> |
distance is fairly small, limiting the summations in the {\sc eam} |
1459 |
> |
equation to the few dozen atoms surrounding atom $i$ for both the |
1460 |
> |
density $\rho$ and pairwise $\phi$ interactions. |
1461 |
> |
|
1462 |
> |
In computing forces for alloys, OpenMD uses mixing rules outlined by |
1463 |
> |
Johnson~\cite{johnson89} to compute the heterogenous pair potential, |
1464 |
|
\begin{equation} |
1465 |
|
\label{eq:johnson} |
1466 |
|
\phi_{ab}(r)=\frac{1}{2}\left( |
1491 |
|
should be noted that the energy units in these files are in eV, not |
1492 |
|
$\mbox{kcal mol}^{-1}$ as in the rest of the {\sc OpenMD} force field |
1493 |
|
files. |
1494 |
+ |
|
1495 |
+ |
\begin{lstlisting}[caption={[An example of a EAMAtomTypes block.] A |
1496 |
+ |
simple example of a EAMAtomTypes block. Here the only data provided is |
1497 |
+ |
the name of a {\tt funcfl} file which contains the raw data for spline |
1498 |
+ |
interpolations for the density, functional, and pair potential.}, |
1499 |
+ |
label={sch:EAMAtomTypes}] |
1500 |
+ |
begin EAMAtomTypes |
1501 |
+ |
Au Au.u3.funcfl |
1502 |
+ |
Ag Ag.u3.funcfl |
1503 |
+ |
Cu Cu.u3.funcfl |
1504 |
+ |
Ni Ni.u3.funcfl |
1505 |
+ |
Pd Pd.u3.funcfl |
1506 |
+ |
Pt Pt.u3.funcfl |
1507 |
+ |
end EAMAtomTypes |
1508 |
+ |
\end{lstlisting} |
1509 |
|
|
1510 |
+ |
|
1511 |
|
\subsubsection{\label{section:ffSC}The SuttonChenAtomTypes block} |
1512 |
|
|
1513 |
|
The Sutton-Chen ({\sc sc})~\cite{Chen90} potential has been used to |
1514 |
< |
study a wide range of phenomena in metals. Although it is similar in |
1515 |
< |
form to the {\sc eam} potential, the Sutton-Chen model takes on a |
1516 |
< |
simpler form, |
1514 |
> |
study a wide range of phenomena in metals. Although it has the same |
1515 |
> |
basic form as the {\sc eam} potential, the Sutton-Chen model takes on |
1516 |
> |
a simpler form, |
1517 |
|
\begin{equation} |
1518 |
|
\label{eq:SCP1} |
1519 |
|
U_{tot}=\sum _{i}\left[ \frac{1}{2}\sum _{j\neq |
1520 |
< |
i}D_{ij}V^{pair}_{ij}(r_{ij})-c_{i}D_{ii}\sqrt{\rho_{i}}\right] , |
1520 |
> |
i}\epsilon_{ij}V^{pair}_{ij}(r_{ij})-c_{i}\epsilon_{ii}\sqrt{\rho_{i}}\right] , |
1521 |
|
\end{equation} |
1522 |
|
where $V^{pair}_{ij}$ and $\rho_{i}$ are given by |
1523 |
|
\begin{equation} |
1524 |
|
\label{eq:SCP2} |
1525 |
|
V^{pair}_{ij}(r)=\left( |
1526 |
< |
\frac{\alpha_{ij}}{r_{ij}}\right)^{n_{ij}}, \rho_{i}=\sum_{j\neq i}\left( |
1526 |
> |
\frac{\alpha_{ij}}{r_{ij}}\right)^{n_{ij}} \hspace{1in} \rho_{i}=\sum_{j\neq i}\left( |
1527 |
|
\frac{\alpha_{ij}}{r_{ij}}\right) ^{m_{ij}} |
1528 |
|
\end{equation} |
1529 |
|
|
1531 |
|
interactions of the pseudo-atom cores. The $\sqrt{\rho_i}$ term in |
1532 |
|
Eq. (\ref{eq:SCP1}) is an attractive many-body potential that models |
1533 |
|
the interactions between the valence electrons and the cores of the |
1534 |
< |
pseudo-atoms. $D_{ij}$, $D_{ii}$, $c_i$ and $\alpha_{ij}$ are |
1535 |
< |
parameters used to tune the potential for different transition |
1536 |
< |
metals. |
1534 |
> |
pseudo-atoms. $\epsilon_{ij}$, $\epsilon_{ii}$, $c_i$ and |
1535 |
> |
$\alpha_{ij}$ are parameters used to tune the potential for different |
1536 |
> |
transition metals. |
1537 |
|
|
1538 |
|
The {\sc sc} potential form has also been parameterized by Qi {\it et |
1539 |
|
al.}\cite{Qi99} These parameters were obtained via empirical and {\it |
1540 |
|
ab initio} calculations to match structural features of the FCC |
1541 |
< |
crystal. To specify the original Sutton-Chen variant of the {\sc sc} |
1542 |
< |
force field, the user would add the {\tt forceFieldVariant = "SC";} |
1304 |
< |
line to the meta-data file, while specification of the Qi {\it et al.} |
1305 |
< |
quantum-adapted variant of the {\sc sc} potential, the user would add |
1306 |
< |
the {\tt forceFieldVariant = "QSC";} line to the meta-data file. |
1541 |
> |
crystal. Interested readers are encouraged to consult reference |
1542 |
> |
\citealp{Qi99} for further details. |
1543 |
|
|
1544 |
+ |
\begin{lstlisting}[caption={[An example of a SCAtomTypes block.] A |
1545 |
+ |
simple example of a SCAtomTypes block. Distances ($\alpha$) |
1546 |
+ |
are given in \AA\ and energies ($\epsilon$) are (by convention) given in |
1547 |
+ |
units of eV. These units must be specified in the {\tt Options} block |
1548 |
+ |
using the keyword {\tt MetallicEnergyUnitScaling}. Without this {\tt |
1549 |
+ |
Options} keyword, the default units for $\epsilon$ are kcal/mol. The |
1550 |
+ |
other parameters, $m$, $n$, and $c$ are unitless.}, |
1551 |
+ |
label={sch:SCAtomTypes}] |
1552 |
+ |
begin SCAtomTypes |
1553 |
+ |
// Name epsilon(eV) c m n alpha(angstroms) |
1554 |
+ |
Ni 0.0073767 84.745 5.0 10.0 3.5157 |
1555 |
+ |
Cu 0.0057921 84.843 5.0 10.0 3.6030 |
1556 |
+ |
Rh 0.0024612 305.499 5.0 13.0 3.7984 |
1557 |
+ |
Pd 0.0032864 148.205 6.0 12.0 3.8813 |
1558 |
+ |
Ag 0.0039450 96.524 6.0 11.0 4.0691 |
1559 |
+ |
Ir 0.0037674 224.815 6.0 13.0 3.8344 |
1560 |
+ |
Pt 0.0097894 71.336 7.0 11.0 3.9163 |
1561 |
+ |
Au 0.0078052 53.581 8.0 11.0 4.0651 |
1562 |
+ |
Au2 0.0078052 53.581 8.0 11.0 4.0651 |
1563 |
+ |
end SCAtomTypes |
1564 |
+ |
\end{lstlisting} |
1565 |
+ |
|
1566 |
|
\subsection{\label{section::ffShortRange}Short Range Interactions} |
1567 |
|
\subsubsection{\label{section:ffBond}The BondTypes block} |
1568 |
|
\subsubsection{\label{section:ffBend}The BendTypes block} |
1613 |
|
|
1614 |
|
\subsubsection{\label{section:ffInversion}The InversionTypes block} |
1615 |
|
\subsection{\label{section::ffLongRange}Long Range Interactions} |
1616 |
< |
\subsubsection{\label{section:ffInversion}The NonBondedInteraction block} |
1616 |
> |
\subsubsection{\label{section:ffNBinteraction}The NonBondedInteraction block} |
1617 |
|
|
1618 |
|
|
1619 |
|
|