ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/openmdDocs/openmdDoc.tex
(Generate patch)

Comparing trunk/openmdDocs/openmdDoc.tex (file contents):
Revision 4102 by gezelter, Fri Apr 18 18:39:38 2014 UTC vs.
Revision 4105 by gezelter, Mon Apr 28 21:41:01 2014 UTC

# Line 10 | Line 10
10   \pagestyle{plain}
11   \pagenumbering{arabic}
12   \usepackage{floatrow}
13 + \usepackage[margin=0.5cm,font=small,format=hang]{caption}
14 +
15   \oddsidemargin 0.0cm
16   \evensidemargin 0.0cm
17   \topmargin -21pt
# Line 29 | Line 31
31   \lstset{language=C,frame=TB,basicstyle=\footnotesize\ttfamily, %
32          xleftmargin=0.25in, xrightmargin=0.25in,captionpos=b, %
33          abovecaptionskip=0.5cm, belowcaptionskip=0.5cm, escapeinside={~}{~}}
34 < \renewcommand{\lstlistingname}{Scheme}
34 > \renewcommand{\lstlistingname}{Example}
35  
36 + \lstnewenvironment{code}[1][]%
37 +  {\noindent\minipage{\linewidth}\vspace{0.5\baselineskip}
38 +   \lstset{language=C,basicstyle=\footnotesize\ttfamily,%
39 +     captionpos=b,aboveskip=0.5cm,belowskip=0.5cm,abovecaptionskip=0.5cm,%
40 +     belowcaptionskip=0.5cm,%
41 +     escapeinside={~}{~},frame=single,#1}}
42 +  {\endminipage}
43 +
44 +
45 +
46   \begin{document}
47  
48   \newcolumntype{A}{p{1.5in}}
# Line 185 | Line 197 | formats are described in the following sections.
197   $<$Snapshot$>$} block.  Both the {\tt $<$MetaData$>$} and {\tt $<$Snapshot$>$}
198   formats are described in the following sections.
199  
200 < \begin{lstlisting}[float,caption={[The structure of an {\sc OpenMD} file]
200 > \begin{code}[caption={[The structure of an {\sc OpenMD} file]
201   The basic structure of an {\sc OpenMD} file contains HTML-like tags to
202   define simulation meta-data and subsequent instantaneous configuration
203 < information. A well-formed {\sc OpenMD} file must contain one $<$MetaData$>$
204 < block and {\it at least one} $<$Snapshot$>$ block.  Each
205 < $<$Snapshot$>$ is further divided into $<$FrameData$>$ and
206 < $<$StuntDoubles$>$ sections.},
195 < label=sch:mdFormat]
203 > information. A well-formed {\sc OpenMD} file must contain one {\tt <MetaData>}
204 > block and {\it at least one} {\tt <Snapshot>} block.  Each
205 > {\tt <Snapshot>} is further divided into {\tt <FrameData>} and
206 > {\tt <StuntDoubles>} sections.},label={sch:mdFormat}]
207   <OpenMD>
208    <MetaData>
209        // see section ~\ref{sec:miscConcepts}~ for details on the formatting
# Line 218 | Line 229 | label=sch:mdFormat]
229    <Snapshot>         // Further information on <Snapshot> blocks
230    </Snapshot>        // can be found in section ~\ref{section:coordFiles}~.
231   </OpenMD>
232 < \end{lstlisting}
232 > \end{code}
233  
234  
235   \section{OpenMD Files and $<$MetaData$>$ blocks}
# Line 234 | Line 245 | Scheme~\ref{sch:mdExample}.
245   shown in Scheme~\ref{sch:mdFormat} and example file is shown in
246   Scheme~\ref{sch:mdExample}.
247  
248 < \begin{lstlisting}[float,caption={[An example of a complete OpenMD
248 > \begin{code}[caption={[An example of a complete OpenMD
249   file] An example showing a complete OpenMD file.},
250   label={sch:mdExample}]
251   <OpenMD>
# Line 273 | Line 284 | statusTime = 50;  // statistics file frequency
284      </StuntDoubles>
285    </Snapshot>
286   </OpenMD>
287 < \end{lstlisting}
287 > \end{code}
288  
289   Within the {\tt $<$MetaData$>$} block it is necessary to provide a
290   complete description of the molecule before it is actually placed in
# Line 287 | Line 298 | become Scheme~\ref{sch:mdExPrime}.
298   Scheme~\ref{sch:mdIncludeExample}, and the new {\sc OpenMD} file would
299   become Scheme~\ref{sch:mdExPrime}.
300  
301 < \begin{lstlisting}[float,caption={An example molecule definition in an
301 > \begin{code}[caption={An example molecule definition in an
302   include file.},label={sch:mdIncludeExample}]
303   molecule{
304    name = "Ar";
# Line 296 | Line 307 | molecule{
307      position( 0.0, 0.0, 0.0 );
308    }
309   }
310 < \end{lstlisting}
310 > \end{code}
311  
312 < \begin{lstlisting}[float,caption={Revised OpenMD input file
312 > \begin{code}[caption={Revised OpenMD input file
313   example.},label={sch:mdExPrime}]
314   <OpenMD>
315    <MetaData>
# Line 331 | Line 342 | statusTime = 50;
342      </StuntDoubles>
343    </Snapshot>
344   </OpenMD>
345 < \end{lstlisting}
345 > \end{code}
346  
347   \section{\label{section:atomsMolecules}Atoms, Molecules, and other
348   ways of grouping atoms}
# Line 402 | Line 413 | rigid body can be seen in Scheme
413   rigid body can be seen in Scheme
414   \ref{sch:rigidBody}.
415  
416 < \begin{lstlisting}[float,caption={[Defining rigid bodies]A sample
416 > \begin{code}[caption={[Defining rigid bodies]A sample
417   definition of a molecule containing a rigid body and a cutoff
418   group},label={sch:rigidBody}]
419   molecule{
# Line 428 | Line 439 | molecule{
439      members(0, 1, 2);
440    }
441   }
442 < \end{lstlisting}
442 > \end{code}
443  
444   \section{\label{sec:miscConcepts}Creating a $<$MetaData$>$ block}
445  
# Line 636 | Line 647 | quaternions to save space in the output files.
647   complete rotation matrix, directional entities are written out using
648   quaternions to save space in the output files.
649  
650 < \begin{lstlisting}[float,caption={[The format of the {\tt $<$Snapshot$>$} block]
650 > \begin{code}[caption={[The format of the {\tt $<$Snapshot$>$} block]
651   An example of the format of the {\tt $<$Snapshot$>$} block.  There is an
652   initial sub-block called {\tt $<$FrameData$>$} which contains the time
653   stamp, the three column vectors of $\mathsf{H}$, and optional extra
# Line 645 | Line 656 | additional information is present on the line.  Atoms
656   configuration of each integrable object.  For each integrable object,
657   the global index is followed by a short string describing what
658   additional information is present on the line.  Atoms with only
659 < position and velocity information use the ``pv'' string which must
659 > position and velocity information use the {\tt pv} string which must
660   then be followed by the position and velocity vectors for that atom.
661 < Directional atoms and Rigid Bodies typically use the ``pvqj'' string
661 > Directional atoms and Rigid Bodies typically use the {\tt pvqj} string
662   which is followed by position, velocity, quaternions, and
663 < lastly, body fixed angular momentum for that integrable object.},
653 < label=sch:dumpFormat]
663 > lastly, body fixed angular momentum for that integrable object.},label={sch:dumpFormat}]
664    <Snapshot>
665      <FrameData>
666          Time: 0
# Line 665 | Line 675 | label=sch:dumpFormat]
675           3      pvqj        x y z vx vy vz  qw qx qy qz jx jy jz
676      </StuntDoubles>
677    </Snapshot>
678 < \end{lstlisting}
678 > \end{code}
679  
680   There are three {\sc OpenMD} files that are written using the combined
681   format.  They are: the initial startup file (\texttt{.md}), the
# Line 716 | Line 726 | An example is given in the {\sc OpenMD} file in Scheme
726   \end{enumerate}
727   An example is given in the {\sc OpenMD} file in Scheme~\ref{sch:initEx1}.
728  
729 < \begin{lstlisting}[float,caption={Example declaration of the
730 < $\text{I}_2$ molecule and the HCl molecule in $<$MetaData$>$ and
731 < $<$Snapshot$>$ blocks.  Note that even though $\text{I}_2$ is
732 < declared before HCl, the $<$Snapshot$>$ block follows the order {\it in
729 > \begin{code}[caption={Example declaration of the
730 > $\text{I}_2$ molecule and the HCl molecule in {\tt <MetaData>} and
731 > {\tt <Snapshot>} blocks.  Note that even though $\text{I}_2$ is
732 > declared before HCl, the {\tt <Snapshot>} block follows the order {\it in
733   which the components were included}.}, label=sch:initEx1]
734   <OpenMD>
735    <MetaData>
# Line 763 | Line 773 | component{
773      </StuntDoubles>
774    </Snapshot>
775   </OpenMD>
776 < \end{lstlisting}
776 > \end{code}
777  
778   \section{The Statistics File}
779  
# Line 779 | Line 789 | statistics file is denoted with the \texttt{.stat} fil
789   allowing the user to gauge the stability of the integrator. The
790   statistics file is denoted with the \texttt{.stat} file extension.
791  
792 < \chapter{\label{section:forceFields}Force Fields}
792 > \chapter{\label{chapter:forceFields}Force Fields}
793  
794   Like many molecular simulation packages, {\sc OpenMD} splits the
795   potential energy into the short-ranged (bonded) portion and a
# Line 793 | Line 803 | section.
803   are defined by the force field which is selected in the MetaData
804   section.
805  
806 < \section{\label{section:shortRange}The basic interactions}
806 > \section{\label{section:divisionOfLabor}Separation into Internal and
807 >  Cross interactions}
808  
809 < The energy function for a system composed of $N$ molecules is
810 < traditionally written
809 > The classical potential energy function for a system composed of $N$
810 > molecules is traditionally written
811   \begin{equation}
812   V = \sum^{N}_{I=1} V^{I}_{\text{Internal}}
813          + \sum^{N-1}_{I=1} \sum_{J>I} V^{IJ}_{\text{Cross}},
814   \label{eq:totalPotential}
815   \end{equation}
816 < where $V^{IJ}_{\text{Cross}}$ contains all intermolecular interactions
817 < between molecules $I$ and $J$, and $V^{I}_{\text{Internal}}$ is the
818 < internal potential of molecule $I$:
819 < \begin{align*}
820 < V^{I}_{\text{Internal}} =  &
821 < \sum_{r_{ij} \in I} V_{\text{bond}}(r_{ij})
811 < + \sum_{\theta_{ijk} \in I} V_{\text{bend}}(\theta_{ijk})
812 < + \sum_{\phi_{ijkl} \in I} V_{\text{torsion}}(\phi_{ijkl})
813 < + \sum_{\omega_{ijkl} \in I} V_{\text{inversion}}(\omega_{ijkl}) \\
814 < & + \sum_{i \in I} \sum_{(j>i+4) \in I}
815 < \biggl[ V_{\text{dispersion}}(r_{ij}) +  V_{\text{electrostatic}}
816 < (\mathbf{r}_{ij},\boldsymbol{\Omega}_{i},\boldsymbol{\Omega}_{j})
817 < \biggr].
818 < \label{eq:internalPotential}
819 < \end{align*}
820 < Here $V_{\text{bond}}, V_{\text{bend}},
821 < V_{\text{torsion}},\mathrm{~and~} V_{\text{inversion}}$ represent the
822 < bond, bend, torsion, and inversion potentials for all
823 < topologically-connected sets of atoms within the molecule.  Bonds are
824 < the primary way of specifying how the atoms are connected together to
825 < form the molecule (i.e. they define the molecular topology).  The
826 < other short-range interactions may be specified explicitly in the
827 < molecule definition, or they may be deduced from bonding information.
828 < For example, bends can be implicitly deduced from two bonds which
829 < share an atom.  Torsions can be deduced from two bends that share a
830 < bond.  Inversion potentials are utilized primarily to enforce
831 < planarity around $sp^2$-hybridized sites, and these are specified with
832 < central atoms and satellites (or an atom with bonds to exactly three
833 < satellites).  The pairwise portions of the non-bonded interactions are
834 < usually excluded for atom pairs that are involved in the same bond,
835 < bend, or torsion. All other atom pairs within a molecule are subject
836 < to non-bonded pair potentials.
816 > where $V^{I}_{\text{Internal}}$ contains all of the terms internal to
817 > molecule $I$ (e.g. bonding, bending, torsional, and inversion terms)
818 > and $V^{IJ}_{\text{Cross}}$ contains all intermolecular interactions
819 > between molecules $I$ and $J$.  For large molecules, the internal
820 > potential may also include some non-bonded terms like electrostatic or
821 > van der Waals interactions.
822  
823   The types of atoms being simulated, as well as the specific functional
824   forms and parameters of the intra-molecular functions and the
# Line 851 | Line 836 | A simple example of a forceField file is shown in sche
836   A simple example of a forceField file is shown in scheme
837   \ref{sch:frcExample}.
838  
839 < \begin{lstlisting}[float,caption={[An example of a complete OpenMD
839 > \begin{code}[caption={[An example of a complete OpenMD
840   force field file for straight-chain united-atom alkanes.] An example
841   showing a complete OpenMD force field for straight-chain united-atom
842   alkanes.}, label={sch:frcExample}]
# Line 898 | Line 883 | end TorsionTypes
883   CH3   CH2  CH2  CH2  Trappe  0.0  0.70544  -0.13549  1.5723  
884   CH2   CH2  CH2  CH2  Trappe  0.0  0.70544  -0.13549  1.5723  
885   end TorsionTypes
886 < \end{lstlisting}
886 > \end{code}
887  
888 < \subsection{\label{section:ffOptions}The Options block}
888 > \section{\label{section:ffOptions}The Options block}
889  
890   The Options block defines properties governing how the force field
891   interactions are carried out.  This block is delineated with the text
# Line 909 | Line 894 | the various keywords and their possible values are giv
894   the various keywords and their possible values are given in Scheme
895   \ref{sch:optionsBlock}.
896  
897 < \begin{lstlisting}[caption={[A force field Options block showing default values
897 > \begin{code}[caption={[A force field Options block showing default values
898   for many force field options.] A force field Options block showing default values
899   for many force field options.  Most of these options do not need to be
900   specified if the default values are working.},
# Line 918 | Line 903 | begin Options
903   Name                      = "alkane"       // any string
904   vdWtype                   = "Lennard-Jones"
905   DistanceMixingRule        = "arithmetic"   // can also be "geometric" or "cubic"
906 < DistanceType              = "sigma"        // can also be Rmin
906 > DistanceType              = "sigma"        // can also be "Rmin"
907   EnergyMixingRule          = "geometric"    // can also be "arithmetic" or "hhg"
908   EnergyUnitScaling         = 1.0
909   MetallicEnergyUnitScaling = 1.0
# Line 935 | Line 920 | end Options
920   GayBerneNu                = 1.0
921   EAMMixingMethod           = "Johnson"      // can also be "Daw"
922   end Options
923 < \end{lstlisting}
923 > \end{code}
924  
925 < \subsection{\label{section:ffBase}The BaseAtomTypes block}
925 > \section{\label{section:ffBase}The BaseAtomTypes block}
926  
927   An AtomType the primary data structure that OpenMD uses to store
928   static data about an atom.  Things that belong to AtomType are
# Line 963 | Line 948 | simulations in Jmol or VMD.
948   ability to print out the names of the base atom types for displaying
949   simulations in Jmol or VMD.
950  
951 < \begin{lstlisting}[caption={[A simple example of a BaseAtomTypes
951 > \begin{code}[caption={[A simple example of a BaseAtomTypes
952   block.] A simple example of a BaseAtomTypes block.},
953   label={sch:baseAtomTypesBlock}]
954   begin BaseAtomTypes
# Line 983 | Line 968 | end BaseAtomTypes
968   Ba      137.327
969   Cl      35.453
970   end BaseAtomTypes
971 < \end{lstlisting}
971 > \end{code}
972  
973 < \subsection{\label{section:ffAtom}The AtomTypes block}
973 > \section{\label{section:ffAtom}The AtomTypes block}
974  
975   AtomTypes inherit most properties from BaseAtomTypes, but can override
976   their lower-level properties as well.  Scheme \ref{sch:atomTypesBlock}
977   shows an example where multiple types of oxygen atoms can inherit mass
978   from the oxygen base type.
979  
980 < \begin{lstlisting}[caption={[An example of a AtomTypes block.] A
980 > \begin{code}[caption={[An example of a AtomTypes block.] A
981   simple example of an AtomTypes block which
982   shows how multiple types can inherit from the same base type.},
983   label={sch:atomTypesBlock}]
# Line 1017 | Line 1002 | end AtomTypes
1002   feo     Fe
1003   lio     Li
1004   end AtomTypes
1005 < \end{lstlisting}
1005 > \end{code}
1006  
1007 < \subsection{\label{section:ffDirectionalAtom}The DirectionalAtomTypes
1007 > \section{\label{section:ffDirectionalAtom}The DirectionalAtomTypes
1008    block}
1009   DirectionalAtoms have orientational degrees of freedom as well as
1010   translation, so moving these atoms requires information about the
# Line 1034 | Line 1019 | body frame.
1019   and in quadrupole tensors that are not necessarily diagonal in the
1020   body frame.
1021  
1022 < \begin{lstlisting}[caption={[An example of a DirectionalAtomTypes block.] A
1022 > \begin{code}[caption={[An example of a DirectionalAtomTypes block.] A
1023   simple example of a DirectionalAtomTypes block.},
1024   label={sch:datomTypesBlock}]
1025   begin DirectionalAtomTypes
# Line 1046 | Line 1031 | end DirectionalAtomTypes                    
1031   CO2             43.06   43.06   0.0    // single-site model for CO2
1032   end DirectionalAtomTypes                    
1033  
1034 < \end{lstlisting}
1034 > \end{code}
1035  
1036   For a DirectionalAtom that represents a linear object, it is
1037   appropriate for one of the moments of inertia to be zero.  In this
# Line 1054 | Line 1039 | calculation of temperatures to reflect this.
1039   of freedom (three translations and two rotations), and will alter
1040   calculation of temperatures to reflect this.
1041  
1042 < \subsection{\label{section::ffAtomProperties}AtomType properties}
1043 < \subsubsection{\label{section:ffLJ}The LennardJonesAtomTypes block}
1042 > \section{\label{section::ffAtomProperties}AtomType properties}
1043 > \subsection{\label{section:ffLJ}The LennardJonesAtomTypes block}
1044   One of the most basic interatomic interactions implemented in {\sc
1045    OpenMD} is the Lennard-Jones potential, which mimics the van der
1046   Waals interaction at long distances and uses an empirical repulsion at
# Line 1094 | Line 1079 | the {\tt NonbondedInteractionTypes} block (see section
1079   the {\tt NonbondedInteractionTypes} block (see section
1080   \ref{section:ffNBinteraction}).
1081  
1082 < \begin{lstlisting}[caption={[An example of a LennardJonesAtomTypes block.] A
1082 > \begin{code}[caption={[An example of a LennardJonesAtomTypes block.] A
1083   simple example of a LennardJonesAtomTypee block.   Units for
1084   $\epsilon$ are kcal / mol and for $\sigma$ are \AA\ .},
1085   label={sch:LJatomTypesBlock}]
# Line 1111 | Line 1096 | end LennardJonesAtomTypes
1096   CH2             0.0866          3.95
1097   CH              0.0189          4.68
1098   end LennardJonesAtomTypes
1099 < \end{lstlisting}
1099 > \end{code}
1100  
1101 < \subsubsection{\label{section:ffCharge}The ChargeAtomTypes block}
1101 > \subsection{\label{section:ffCharge}The ChargeAtomTypes block}
1102  
1103   In molecular simulations, proper accumulation of the electrostatic
1104   interactions is essential and is one of the most
# Line 1136 | Line 1121 | of free space.
1121   charge of an electron in Coulombs.  $\epsilon_0$ is the permittivity
1122   of free space.
1123  
1124 < \begin{lstlisting}[caption={[An example of a ChargeAtomTypes block.] A
1124 > \begin{code}[caption={[An example of a ChargeAtomTypes block.] A
1125   simple example of a ChargeAtomTypes block.   Units for
1126   charge are in multiples of electron charge.},
1127   label={sch:ChargeAtomTypesBlock}]
# Line 1151 | Line 1136 | end ChargeAtomTypes
1136   Na+             1.0
1137   Cl-            -1.0
1138   end ChargeAtomTypes
1139 < \end{lstlisting}
1139 > \end{code}
1140  
1141 < \subsubsection{\label{section:ffMultipole}The MultipoleAtomTypes
1141 > \subsection{\label{section:ffMultipole}The MultipoleAtomTypes
1142    block}
1143   For complex charge distributions that are centered on single sites, it
1144   is convenient to write the total electrostatic potential in terms of
# Line 1169 | Line 1154 | given by $C_{\bf a}$, $D_{{\bf a}\alpha}$, and $Q_{{\b
1154   \end{equation}
1155   Here, the point charge, dipole, and quadrupole for site $\bf a$ are
1156   given by $C_{\bf a}$, $D_{{\bf a}\alpha}$, and $Q_{{\bf
1157 <    a}\alpha\beta}$, respectively.  These are the primitive
1157 >    a}\alpha\beta}$, respectively.  These are the {\it primitive}
1158   multipoles.  If the site is representing a distribution of charges,
1159   these can be expressed as,
1160   \begin{align}
# Line 1203 | Line 1188 | the unit vector pointing along $\mathbf{r}_{ij}$
1188   ($\boldsymbol{\hat{r}}_{ij}=\mathbf{r}_{ij}/|\mathbf{r}_{ij}|$).
1189  
1190  
1191 < \begin{lstlisting}[caption={[An example of a MultipoleAtomTypes block.] A
1191 > \begin{code}[caption={[An example of a MultipoleAtomTypes block.] A
1192   simple example of a MultipoleAtomTypes block.   Dipoles are given in
1193   units of Debyes, and Quadrupole moments are given in units of Debye
1194   \AA~(or $10^{-26} \mathrm{~esu~cm}^2$)},
# Line 1223 | Line 1208 | end MultipoleAtomTypes
1208   // name dq phi theta psi dipole_moment  Qxx    Qyy     Qzz
1209   SSD     dq 0.0 0.0   0.0     2.35      -1.682  1.762   -0.08
1210   end MultipoleAtomTypes
1211 < \end{lstlisting}
1211 > \end{code}
1212  
1213   Specifying a MultipoleAtomType requires declaring how the
1214   electrostatic frame for the site is rotated relative to the body-fixed
# Line 1234 | Line 1219 | and Quadrupole moments in units of Debye \AA.
1219   in units of degrees.  Dipole moments are entered in units of Debye,
1220   and Quadrupole moments in units of Debye \AA.
1221  
1222 < \subsubsection{\label{section:ffGB}The FluctuatingChargeAtomTypes  block}
1223 < \subsubsection{\label{section:ffPol}The PolarizableAtomTypes block}
1239 < \subsubsection{\label{section:ffGB}The GayBerneAtomTypes block}
1222 > \subsection{\label{section:ffGB}The FluctuatingChargeAtomTypes  block}
1223 > %\subsubsection{\label{section:ffPol}The PolarizableAtomTypes block}
1224  
1225 + \subsection{\label{section:ffGB}The GayBerneAtomTypes block}
1226 +
1227   The Gay-Berne potential has been widely used in the liquid crystal
1228 < community to describe this anisotropic phase
1228 > community to describe anisotropic phase
1229   behavior.~\cite{Gay:1981yu,Berne:1972pb,Kushick:1976xy,Luckhurst:1990fy,Cleaver:1996rt}
1230   The form of the Gay-Berne potential implemented in OpenMD was
1231   generalized by Cleaver {\it et al.} and is appropriate for dissimilar
1232 < uniaxial ellipsoids.\cite{Cleaver:1996rt}  The potential is constructed in the
1233 < familiar form of the Lennard-Jones function using
1232 > uniaxial ellipsoids.\cite{Cleaver:1996rt} The potential is constructed
1233 > in the familiar form of the Lennard-Jones function using
1234   orientation-dependent $\sigma$ and $\epsilon$ parameters,
1235   \begin{equation*}
1236   V_{ij}({{\bf \hat u}_i}, {{\bf \hat u}_j}, {{\bf \hat
# Line 1289 | Line 1275 | in Ref. \citealp{Golubkov06}
1275   efficiently compute forces and torques for this potential can be found
1276   in Ref. \citealp{Golubkov06}
1277  
1278 < \begin{lstlisting}[caption={[An example of a GayBerneAtomTypes block.] A
1278 > \begin{code}[caption={[An example of a GayBerneAtomTypes block.] A
1279   simple example of a GayBerneAtomTypes block.  Distances ($d$ and $l$)
1280   are given in \AA\ and energies ($\epsilon_X, \epsilon_S, \epsilon_E$)
1281   are in units of kcal/mol. $dw$ is unitless.},
# Line 1300 | Line 1286 | end GayBerneAtomTypes                  
1286   GBC6H6          4.65    2.03    0.540           0.540           1.9818    0.6
1287   GBCH3OH         2.55    3.18    0.542           0.542           0.55826   1.0
1288   end GayBerneAtomTypes                  
1289 < \end{lstlisting}
1289 > \end{code}
1290  
1291 < \subsubsection{\label{section:ffSticky}The StickyAtomTypes block}
1291 > \subsection{\label{section:ffSticky}The StickyAtomTypes block}
1292  
1293   One of the solvents that can be simulated by {\sc OpenMD} is the
1294   extended Soft Sticky Dipole (SSD/E) water model.\cite{fennell04} The
# Line 1398 | Line 1384 | reference~\citealp{fennell04}.
1384   density corrected SSD models can be found in
1385   reference~\citealp{fennell04}.
1386  
1387 < \begin{lstlisting}[caption={[An example of a StickyAtomTypes block.] A
1387 > \begin{code}[caption={[An example of a StickyAtomTypes block.] A
1388   simple example of a StickyAtomTypes block.  Distances ($r_l$, $r_u$,
1389   $r_{l}'$ and $r_{u}'$) are given in \AA\ and energies ($v_0, v_{0}'$)
1390   are in units of kcal/mol. $w_0$ is unitless.},
# Line 1410 | Line 1396 | end StickyAtomTypes
1396   SSD     0.07715 3.7284          3.7284  2.75      3.35  2.75    4.0
1397   SSD1    0.07715 3.6613          3.6613  2.75      3.35  2.75    4.0
1398   end StickyAtomTypes
1399 < \end{lstlisting}
1399 > \end{code}
1400  
1401 < \subsection{\label{section::ffMetals}Metallic Atom Types}
1401 > \section{\label{section::ffMetals}Metallic Atom Types}
1402  
1403   {\sc OpenMD} implements a number of related potentials that describe
1404   bonding in transition metals. These potentials have an attractive
# Line 1441 | Line 1427 | repulsive interaction between atoms $i$ and $j$.
1427   The pairwise portion of the potential, $\phi_{ij}$, is usually a
1428   repulsive interaction between atoms $i$ and $j$.
1429  
1430 < \subsubsection{\label{section:ffEAM}The EAMAtomTypes block}
1430 > \subsection{\label{section:ffEAM}The EAMAtomTypes block}
1431   The Embedded Atom Method ({\sc eam}) is one of the most widely-used
1432   potentials for transition
1433   metals.~\cite{Finnis84,Ercolessi88,Chen90,Qi99,Ercolessi02,Daw84,FBD86,johnson89,Lu97}
# Line 1492 | Line 1478 | files.  
1478   $\mbox{kcal mol}^{-1}$ as in the rest of the {\sc OpenMD} force field
1479   files.  
1480  
1481 < \begin{lstlisting}[caption={[An example of a EAMAtomTypes block.] A
1481 > \begin{code}[caption={[An example of a EAMAtomTypes block.] A
1482   simple example of a EAMAtomTypes block. Here the only data provided is
1483   the name of a {\tt funcfl} file which contains the raw data for spline
1484   interpolations for the density, functional, and pair potential.},
# Line 1505 | Line 1491 | end EAMAtomTypes
1491   Pd      Pd.u3.funcfl
1492   Pt      Pt.u3.funcfl
1493   end EAMAtomTypes
1494 < \end{lstlisting}
1494 > \end{code}
1495  
1496 <
1511 < \subsubsection{\label{section:ffSC}The SuttonChenAtomTypes block}
1496 > \subsection{\label{section:ffSC}The SuttonChenAtomTypes block}
1497  
1498   The Sutton-Chen ({\sc sc})~\cite{Chen90} potential has been used to
1499   study a wide range of phenomena in metals.  Although it has the same
1500 < basic form as the {\sc eam} potential, the Sutton-Chen model takes on
1501 < a simpler form,
1500 > basic form as the {\sc eam} potential, the Sutton-Chen model requires
1501 > a simpler set of parameters,
1502   \begin{equation}
1503   \label{eq:SCP1}
1504   U_{tot}=\sum _{i}\left[ \frac{1}{2}\sum _{j\neq
# Line 1541 | Line 1526 | crystal.  Interested readers are encouraged to consult
1526   crystal.  Interested readers are encouraged to consult reference
1527   \citealp{Qi99} for further details.
1528  
1529 < \begin{lstlisting}[caption={[An example of a SCAtomTypes block.] A
1529 > \begin{code}[caption={[An example of a SCAtomTypes block.] A
1530   simple example of a SCAtomTypes block.  Distances ($\alpha$)
1531   are given in \AA\ and energies ($\epsilon$) are (by convention) given in
1532   units of eV.  These units must be specified in the {\tt Options} block
# Line 1561 | Line 1546 | end SCAtomTypes
1546   Au      0.0078052       53.581  8.0     11.0    4.0651
1547   Au2     0.0078052       53.581  8.0     11.0    4.0651
1548   end SCAtomTypes
1549 < \end{lstlisting}
1549 > \end{code}
1550  
1551 < \subsection{\label{section::ffShortRange}Short Range Interactions}
1552 < \subsubsection{\label{section:ffBond}The BondTypes block}
1553 < \subsubsection{\label{section:ffBend}The BendTypes block}
1554 < A harmonic bend potential is represented by the following function:
1555 < \begin{equation}
1556 < V_{\text{bend}}(\theta_{ijk}) = k_{\theta}( \theta_{ijk} - \theta_0
1557 < )^2, \label{eq:bendPot}
1558 < \end{equation}
1559 < where $\theta_{ijk}$ is the angle defined by atoms $i$, $j$, and $k$,
1560 < $\theta_0$ is the equilibrium bond angle, and $k_{\theta}$ is the
1561 < force constant which determines the strength of the harmonic bend.
1551 > \section{\label{section::ffShortRange}Short Range Interactions}
1552 > The internal structure of a molecule is usually specified in terms of
1553 > a set of ``bonded'' terms in the potential energy function for
1554 > molecule $I$,
1555 > \begin{align*}
1556 > V^{I}_{\text{Internal}} =  &
1557 > \sum_{r_{ij} \in I} V_{\text{bond}}(r_{ij})
1558 > + \sum_{\theta_{ijk} \in I} V_{\text{bend}}(\theta_{ijk})
1559 > + \sum_{\phi_{ijkl} \in I} V_{\text{torsion}}(\phi_{ijkl})
1560 > + \sum_{\omega_{ijkl} \in I} V_{\text{inversion}}(\omega_{ijkl}) \\
1561 > & + \sum_{i \in I} \sum_{(j>i+4) \in I}
1562 > \biggl[ V_{\text{dispersion}}(r_{ij}) +  V_{\text{electrostatic}}
1563 > (\mathbf{r}_{ij},\boldsymbol{\Omega}_{i},\boldsymbol{\Omega}_{j})
1564 > \biggr].
1565 > \label{eq:internalPotential}
1566 > \end{align*}
1567 > Here $V_{\text{bond}}, V_{\text{bend}},
1568 > V_{\text{torsion}},\mathrm{~and~} V_{\text{inversion}}$ represent the
1569 > bond, bend, torsion, and inversion potentials for all
1570 > topologically-connected sets of atoms within the molecule.  Bonds are
1571 > the primary way of specifying how the atoms are connected together to
1572 > form the molecule (i.e. they define the molecular topology).  The
1573 > other short-range interactions may be specified explicitly in the
1574 > molecule definition, or they may be deduced from bonding information.
1575 > For example, bends can be implicitly deduced from two bonds which
1576 > share an atom.  Torsions can be deduced from two bends that share a
1577 > bond.  Inversion potentials are utilized primarily to enforce
1578 > planarity around $sp^2$-hybridized sites, and these are specified with
1579 > central atoms and satellites (or an atom with bonds to exactly three
1580 > satellites). Non-bonded interactions are usually excluded for atom
1581 > pairs that are involved in the same bond, bend, or torsion, but all
1582 > other atom pairs are included in the standard non-bonded interactions.
1583  
1584 < \subsubsection{\label{section:ffTorsion}The TorsionTypes block}
1585 < The torsion potential is often represented as a cosine series of the
1586 < form:
1584 > Bond lengths, angles, and torsions (dihedrals) are ``natural''
1585 > coordinates to treat molecular motion, as it is usually in these
1586 > coordinates that most chemists understand the behavior of molecules.
1587 > The bond lengths and angles are often considered ``hard'' degrees of
1588 > freedom.  That is, we can't deform them very much without a
1589 > significant energetic penalty.  On the other hand, dihedral angles or
1590 > torsions are ``soft'' and typically undergo significant deformation
1591 > under normal conditions.
1592 >
1593 > \subsection{\label{section:ffBond}The BondTypes block}
1594 >
1595 > Bonds are the primary way to specify how the atoms are connected
1596 > together to form the molecule.  In general, bonds exert strong
1597 > restoring forces to keep the molecule compact.  The bond energy
1598 > functions are relatively simple functions of the distance between two
1599 > atomic sites,
1600   \begin{equation}
1601 < V_{\text{torsion}}(\phi) = c_1[1 + \cos \phi]
1602 <        + c_2[1 + \cos(2\phi)]
1603 <        + c_3[1 + \cos(3\phi)],
1601 > b = \left| \vec{r}_{ij} \right| = \sqrt{(x_j - x_i)^2 + (y_j - y_i)^2
1602 >  + (z_j - z_i)^2}.
1603 > \end{equation}
1604 > All BondTypes must specify two AtomType names ($i$ and $j$) that
1605 > describe when that bond should be applied, as well as an equilibrium
1606 > bond length, $b_{ij}^0$, in units of \AA. The most common forms for
1607 > bonding potentials are {\tt Harmonic} bonds,
1608 > \begin{equation}
1609 > V_{\text{bond}}(b) = \frac{k_{ij}}{2} \left(b -
1610 >  b_{ij}^0 \right)^2
1611 > \end{equation}
1612 > and {\tt Morse} bonds,
1613 > \begin{equation}
1614 > V_{\text{bond}}(b) = D_{ij} \left[ 1 - e^{-\beta_{ij} (b - b_{ij}^0)} \right]^2
1615 > \end{equation}
1616 >
1617 > \begin{figure}[h]
1618 > \centering
1619 > \includegraphics[width=2.5in]{bond.pdf}
1620 > \caption[Bond coordinates]{The coordinate describing a
1621 > a bond between atoms $i$ and $j$ is $|r_{ij}|$, the length of the
1622 > $\vec{r}_{ij}$ vector. }
1623 > \label{fig:bond}
1624 > \end{figure}
1625 >
1626 > OpenMD can also simulate some less common types of bond potentials,
1627 > including {\tt Fixed} bonds (which are constrained to be at a
1628 > specified bond length),
1629 > \begin{equation}
1630 > V_{\text{bond}}(b) = 0.
1631 > \end{equation}
1632 > {\tt Cubic} bonds include terms to model anharmonicity,
1633 > \begin{equation}
1634 > V_{\text{bond}}(b) =  K_3 (b -  b_{ij}^0)^3 + K_2 (b - b_{ij}^0)^2 + K_1 (b -  b_{ij}^0) + K_0,
1635 > \end{equation}
1636 > and {\tt Quartic} bonds, include another term in the Taylor
1637 > expansion around $b_{ij}^0$,
1638 > \begin{equation}
1639 > V_{\text{bond}}(b) = K_4 (b -  b_{ij}^0)^4 +  K_3 (b -  b_{ij}^0)^3 +
1640 > K_2 (b - b_{ij}^0)^2 + K_1 (b -  b_{ij}^0) + K_0,
1641 > \end{equation}
1642 > can also be simulated.  Note that the factor of $1/2$ that is included
1643 > in the {\tt Harmonic} bond type force constant is {\it not} present in
1644 > either the {\tt Cubic} or {\tt Quartic} bond types.
1645 >
1646 > {\tt Polynomial} bonds which can have any number of terms,
1647 > \begin{equation}
1648 > V_{\text{bond}}(b) = \sum_n K_n (b -  b_{ij}^0)^n.
1649 > \end{equation}
1650 > can also be specified by giving a sequence of integer ($n$) and force
1651 > constant ($K_n$) pairs.
1652 >
1653 > The order of terms in the BondTypes block is:
1654 > \begin{itemize}
1655 > \item {\tt AtomType} 1
1656 > \item {\tt AtomType} 2
1657 > \item {\tt BondType} (one of {\tt Harmonic}, {\tt Morse}, {\tt Fixed}, {\tt
1658 >        Cubic}, {\tt Quartic}, or {\tt Polynomial})
1659 > \item $b_{ij}^0$, the equilibrium bond length in \AA
1660 > \item any other parameters required by the {\tt BondType}
1661 > \end{itemize}
1662 >
1663 > \begin{code}[caption={[An example of a BondTypes block.] A
1664 > simple example of a BondTypes block.  Distances ($b_0$)
1665 > are given in \AA\ and force constants are given in
1666 > units so that when multiplied by the correct power of distance they
1667 > return energies in kcal/mol.  For example $k$ for a Harmonic bond is
1668 > in units of kcal/mol/\AA$^2$.},
1669 > label={sch:BondTypes}]
1670 > begin BondTypes
1671 > //Atom1 Atom2   Harmonic        b0        k (kcal/mol/A^2)
1672 > CH2     CH2     Harmonic        1.526     260
1673 > //Atom1 Atom2   Morse           b0        D       beta (A^-1)
1674 > CN      NC      Morse           1.157437  212.95  2.5802
1675 > //Atom1 Atom2   Fixed           b0
1676 > CT      HC      Fixed           1.09
1677 > //Atom1 Atom2   Cubic           b0        K3      K2      K1      K0
1678 > //Atom1 Atom2   Quartic         b0        K4      K3      K2      K1      K0
1679 > //Atom1 Atom2   Polynomial      b0        n       Kn      [m      Km]
1680 > end BondTypes
1681 > \end{code}
1682 >
1683 > There are advantages and disadvantages of all of the different types
1684 > of bonds, but specific simulation tasks may call for specific
1685 > behaviors.
1686 >
1687 > \subsection{\label{section:ffBend}The BendTypes block}
1688 > The equilibrium geometries and energy functions for bending motions in
1689 > a molecule are strongly dependent on the bonding environment of the
1690 > central atomic site.  For example, different types of hybridized
1691 > carbon centers require different bending angles and force constants to
1692 > describe the local geometry.
1693 >
1694 > The bending potential energy functions used in most force fields are
1695 > often simple functions of the angle between two bonds,
1696 > \begin{equation}
1697 > \theta_{ijk} = \cos^{-1} \left(\frac{\vec{r}_{ji} \cdot
1698 >    \vec{r}_{jk}}{\left| \vec{r}_{ji} \right| \left| \vec{r}_{jk}
1699 >    \right|} \right)
1700 > \end{equation}
1701 > Here atom $j$ is the central atom that is bonded to two partners $i$
1702 > and $k$.
1703 >
1704 > \begin{figure}[h]
1705 > \centering
1706 > \includegraphics[width=3.5in]{bend.pdf}
1707 > \caption[Bend angle coordinates]{The coordinate describing a bend
1708 >  between atoms $i$, $j$, and $k$ is the angle $\theta_{ijk} =
1709 >  \cos^{-1} \left(\hat{r}_{ji} \cdot \hat{r}_{jk}\right)$ where $\hat{r}_{ji}$ is
1710 >  the unit vector between atoms $j$ and $i$. }
1711 > \label{fig:bend}
1712 > \end{figure}
1713 >
1714 >
1715 > All BendTypes must specify three AtomType names ($i$, $j$ and $k$)
1716 > that describe when that bend potential should be applied, as well as
1717 > an equilibrium bending angle, $\theta_{ijk}^0$, in units of
1718 > degrees. The most common forms for bending potentials are {\tt
1719 >  Harmonic} bends,
1720 > \begin{equation}
1721 > V_{\text{bend}}(\theta_{ijk}) = \frac{k_{ijk}}{2}( \theta_{ijk} - \theta_{ijk}^0
1722 > )^2, \label{eq:bendPot}
1723 > \end{equation}
1724 > where $k_{ijk}$ is the force constant which determines the strength of
1725 > the harmonic bend. {\tt UreyBradley} bends utilize an additional 1-3
1726 > bond-type interaction in addition to the harmonic bending potential:
1727 > \begin{equation}
1728 >  V_{\text{bend}}(\vec{r}_i , \vec{r}_j, \vec{r}_k)
1729 >  =\frac{k_{ijk}}{2}( \theta_{ijk} - \theta_{ijk}^0)^2
1730 >  + \frac{k_{ub}}{2}( r_{ik} - s_0 )^2. \label{eq:ubBend}
1731 > \end{equation}
1732 >
1733 > A {\tt Cosine} bend is a variant on the harmonic bend which utilizes
1734 > the cosine of the angle instead of the angle itself,
1735 > \begin{equation}
1736 > V_{\text{bend}}(\theta_{ijk}) = \frac{k_{ijk}}{2}( \cos\theta_{ijk} -
1737 > \cos \theta_{ijk}^0 )^2. \label{eq:cosBend}
1738 > \end{equation}
1739 >
1740 > OpenMD can also simulate some less common types of bend potentials,
1741 > including {\tt Cubic} bends, which include terms to model anharmonicity,
1742 > \begin{equation}
1743 > V_{\text{bend}}(\theta_{ijk}) =  K_3 (\theta_{ijk} -  \theta_{ijk}^0)^3 + K_2 (\theta_{ijk} -  \theta_{ijk}^0)^2 + K_1 (\theta_{ijk} -  \theta_{ijk}^0) + K_0,
1744 > \end{equation}
1745 > and {\tt Quartic} bends, which include another term in the Taylor
1746 > expansion around $\theta_{ijk}^0$,
1747 > \begin{equation}
1748 >  V_{\text{bend}}(\theta_{ijk}) = K_4 (\theta_{ijk} -  \theta_{ijk}^0)^4 +  K_3 (\theta_{ijk} -  \theta_{ijk}^0)^3 +
1749 >  K_2 (\theta_{ijk} -  \theta_{ijk}^0)^2 + K_1 (\theta_{ijk} -
1750 >  \theta_{ijk}^0) + K_0,
1751 > \end{equation}
1752 > can also be simulated.  Note that the factor of $1/2$ that is included
1753 > in the {\tt Harmonic} bend type force constant is {\it not} present in
1754 > either the {\tt Cubic} or {\tt Quartic} bend types.
1755 >
1756 > {\tt Polynomial} bends which can have any number of terms,
1757 > \begin{equation}
1758 > V_{\text{bend}}(\theta_{ijk}) = \sum_n K_n (\theta_{ijk} -  \theta_{ijk}^0)^n.
1759 > \end{equation}
1760 > can also be specified by giving a sequence of integer ($n$) and force
1761 > constant ($K_n$) pairs.
1762 >
1763 > The order of terms in the BendTypes block is:
1764 > \begin{itemize}
1765 > \item {\tt AtomType} 1
1766 > \item {\tt AtomType} 2 (this is the central atom)
1767 > \item {\tt AtomType} 3
1768 > \item {\tt BendType} (one of {\tt Harmonic}, {\tt UreyBradley}, {\tt
1769 >    Cosine}, {\tt Cubic}, {\tt Quartic}, or {\tt Polynomial})
1770 > \item $\theta_{ijk}^0$, the equilibrium bending angle in degrees.
1771 > \item any other parameters required by the {\tt BendType}
1772 > \end{itemize}
1773 >
1774 > \begin{code}[caption={[An example of a BendTypes block.] A
1775 > simple example of a BendTypes block.  By convention, equilibrium angles
1776 > ($\theta_0$) are given in degrees but force constants are given in
1777 > units so that when multiplied by the correct power of angle (in
1778 > radians) they return energies in kcal/mol.  For example $k$ for a
1779 > Harmonic bend is in units of kcal/mol/radians$^2$.},
1780 > label={sch:BendTypes}]
1781 > begin BendTypes
1782 > //Atom1 Atom2   Atom3   Harmonic      theta0(deg) Ktheta(kcal/mol/radians^2)
1783 > CT      CT      CT      Harmonic      109.5        80.000000
1784 > CH2     CH      CH2     Harmonic      112.0       117.68
1785 > CH3     CH2     SH      Harmonic       96.0        67.220
1786 > //UreyBradley
1787 > //Atom1 Atom2   Atom3   UreyBradley   theta0      Ktheta  s0  Kub
1788 > //Cosine
1789 > //Atom1 Atom2   Atom3   Cosine        theta0      Ktheta(kcal/mol)
1790 > //Cubic
1791 > //Atom1 Atom2   Atom3   Cubic         theta0      K3      K2  K1   K0
1792 > //Quartic
1793 > //Atom1 Atom2   Atom3   Quartic       theta0      K4      K3  K2   K1   K0
1794 > //Polynomial
1795 > //Atom1 Atom2   Atom3   Polynomial    theta0      n       Kn  [m   Km]
1796 > end BendTypes
1797 > \end{code}
1798 >
1799 > Note that the parameters for a particular bend type are the same for
1800 > any bending triplet of the same atomic types (in the same or reversed
1801 > order).  Changing the AtomType in the Atom2 position will change the
1802 > matched bend types in the force field.
1803 >
1804 > \subsection{\label{section:ffTorsion}The TorsionTypes block}
1805 > The torsion potential for rotation of groups around a central bond can
1806 > often be represented with various cosine functions.  For two
1807 > tetrahedral ($sp^3$) carbons connected by a single bond, the torsion
1808 > potential might be
1809 > \begin{equation*}
1810 > V_{\text{torsion}} = \frac{v}{2} \left[ 1 + \cos( 3 \phi ) \right]
1811 > \end{equation*}
1812 > where $v$ is the barrier for going from {\em staggered} $\rightarrow$
1813 > {\em eclipsed} conformations, while for $sp^2$ carbons connected by a
1814 > double bond, the potential might be
1815 > \begin{equation*}
1816 > V_{\text{torsion}} = \frac{w}{2} \left[ 1 - \cos( 2 \phi ) \right]
1817 > \end{equation*}
1818 > where $w$ is the barrier for going from  {\em cis} $\rightarrow$ {\em
1819 >  trans} conformations.
1820 >
1821 > A general torsion potentials can be represented as a cosine series of
1822 > the form:
1823 > \begin{equation}
1824 > V_{\text{torsion}}(\phi_{ijkl}) = c_1[1 + \cos \phi_{ijkl}]
1825 >        + c_2[1 - \cos(2\phi_{ijkl})]
1826 >        + c_3[1 + \cos(3\phi_{ijkl})],
1827   \label{eq:origTorsionPot}
1828   \end{equation}
1829 < where:
1829 > where the angle $\phi_{ijkl}$ is defined
1830   \begin{equation}
1831 < \cos\phi = (\hat{\mathbf{r}}_{ij} \times \hat{\mathbf{r}}_{jk}) \cdot
1831 > \cos\phi_{ijkl} = (\hat{\mathbf{r}}_{ij} \times \hat{\mathbf{r}}_{jk}) \cdot
1832          (\hat{\mathbf{r}}_{jk} \times \hat{\mathbf{r}}_{kl}).
1833   \label{eq:torsPhi}
1834   \end{equation}
1835   Here, $\hat{\mathbf{r}}_{\alpha\beta}$ are the set of unit bond
1836 < vectors between atoms $i$, $j$, $k$, and $l$. For computational
1837 < efficiency, the torsion potential has been recast after the method of
1838 < {\sc charmm},\cite{Brooks83} in which the angle series is converted to
1839 < a power series of the form:
1836 > vectors between atoms $i$, $j$, $k$, and $l$.  Note that some force
1837 > fields define the zero of the $\phi_{ijkl}$ angle when atoms $i$ and
1838 > $l$ are in the {\em trans} configuration, while most define the zero
1839 > angle for when $i$ and $l$ are in the fully eclipsed orientation.  The
1840 > behavior of the torsion parser can be altered with the {\tt
1841 >  TorsionAngleConvention} keyword in the Options block.  The default
1842 > behavior is {\tt "180\_is\_trans"} while the opposite behavior can be
1843 > invoked by setting this keyword to {\tt "0\_is\_trans"}.
1844 >
1845 > \begin{figure}[h]
1846 > \centering
1847 > \includegraphics[width=4.5in]{torsion.pdf}
1848 > \caption[Torsion or dihedral angle coordinates]{The coordinate
1849 >  describing a torsion between atoms $i$, $j$, $k$, and $l$ is the
1850 >  dihedral angle $\phi_{ijkl}$ which measures the relative rotation of
1851 >  the two terminal atoms around the axis defined by the middle bond. }
1852 > \label{fig:torsion}
1853 > \end{figure}
1854 >
1855 > For computational efficiency, OpenMD recasts torsion potential in the
1856 > method of {\sc charmm},\cite{Brooks83} in which the angle series is
1857 > converted to a power series of the form:
1858   \begin{equation}
1859 < V_{\text{torsion}}(\phi) =  
1860 <        k_3 \cos^3 \phi + k_2 \cos^2 \phi + k_1 \cos \phi + k_0,
1859 >  V_{\text{torsion}}(\phi_{ijkl}) =  
1860 >  k_3 \cos^3 \phi_{ijkl} + k_2 \cos^2 \phi_{ijkl} + k_1 \cos \phi_{ijkl} + k_0,
1861   \label{eq:torsionPot}
1862   \end{equation}
1863   where:
1864   \begin{align*}
1865 < k_0 &= c_1 + c_3, \\
1865 > k_0 &= c_1 + 2 c_2 + c_3, \\
1866   k_1 &= c_1 - 3c_3, \\
1867 < k_2 &= 2 c_2, \\
1868 < k_3 &= 4c_3.
1867 > k_2 &= - 2 c_2, \\
1868 > k_3 &= 4 c_3.
1869   \end{align*}
1870   By recasting the potential as a power series, repeated trigonometric
1871   evaluations are avoided during the calculation of the potential
1872   energy.
1873  
1874 < \subsubsection{\label{section:ffInversion}The InversionTypes block}
1875 < \subsection{\label{section::ffLongRange}Long Range Interactions}
1876 < \subsubsection{\label{section:ffNBinteraction}The NonBondedInteraction block}
1874 > Using this framework, OpenMD implements a variety of different
1875 > potential energy functions for torsions:
1876 > \begin{itemize}
1877 > \item {\tt Cubic}:
1878 > \begin{equation*}
1879 >  V_{\text{torsion}}(\phi) =  
1880 >  k_3 \cos^3 \phi + k_2 \cos^2 \phi + k_1 \cos \phi + k_0,
1881 > \end{equation*}
1882 > \item {\tt Quartic}:
1883 > \begin{equation*}
1884 >  V_{\text{torsion}}(\phi) =  k_4 \cos^4 \phi +
1885 >  k_3 \cos^3 \phi + k_2 \cos^2 \phi + k_1 \cos \phi + k_0,
1886 > \end{equation*}
1887 > \item {\tt Polynomial}:
1888 > \begin{equation*}
1889 > V_{\text{torsion}}(\phi) =  \sum_n k_n \cos^n \phi ,
1890 > \end{equation*}
1891 > \item {\tt Charmm}:
1892 > \begin{equation*}
1893 > V_{\text{torsion}}(\phi) = \sum_n K_n \left( 1 + cos(n
1894 >  \phi - \delta_n) \right),
1895 > \end{equation*}
1896 > \item {\tt Opls}:
1897 > \begin{equation*}
1898 >  V_{\text{torsion}}(\phi) =  \frac{1}{2} \left(v_1 (1 + \cos \phi) \right)
1899 >    + v_2 (1 - \cos 2 \phi) +  v_3 (1 + \cos 3 \phi),
1900 > \end{equation*}
1901 > \item {\tt Trappe}:\cite{Siepmann1998}
1902 > \begin{equation*}
1903 >  V_{\text{torsion}}(\phi) =  c_0 + c_1 (1 + \cos \phi) + c_2 (1 - \cos 2 \phi)  +
1904 >  c_3 (1 + \cos 3 \phi),
1905 > \end{equation*}
1906 > \item {\tt Harmonic}:
1907 > \begin{equation*}
1908 > V_{\text{torsion}}(\phi) =  \frac{d_0}{2} \left(\phi - \phi^0\right).
1909 > \end{equation*}
1910 > \end{itemize}
1911  
1912 + Most torsion types don't require specific angle information in the
1913 + parameters since they are typically expressed in cosine polynomials.
1914 + {\tt Charmm} and {\tt Harmonic} torsions are a bit different.  {\tt
1915 +  Charmm} torsion types require a set of phase angles, $\delta_n$ that
1916 + are expressed in degrees, and associated periodicity numbers, $n$.
1917 + {\tt Harmonic} torsions have an equilibrium torsion angle, $\phi_0$
1918 + that is measured in degrees, while $d_0$ has units of
1919 + kcal/mol/degrees$^2$.  All other torsion parameters are measured in
1920 + units of kcal/mol.
1921  
1922 + \begin{code}[caption={[An example of a TorsionTypes block.] A
1923 + simple example of a TorsionTypes block.  Energy constants are given in
1924 + kcal / mol, and when required by the form, $\delta$ angles are given
1925 + in degrees.},
1926 + label={sch:TorsionTypes}]
1927 + begin TorsionTypes
1928 + //Cubic
1929 + //Atom1 Atom2   Atom3   Atom4   Cubic   k3       k2        k1      k0  
1930 + CH2     CH2     CH2     CH2     Cubic   5.9602   -0.2568   -3.802  2.1586
1931 + CH2     CH      CH      CH2     Cubic   3.3254   -0.4215   -1.686  1.1661
1932 + //Trappe
1933 + //Atom1 Atom2   Atom3   Atom4   Trappe  c0       c1        c2      c3
1934 + CH3     CH2     CH2     SH      Trappe  0.10507  -0.10342  0.03668 0.60874    
1935 + //Charmm
1936 + //Atom1 Atom2   Atom3   Atom4   Charmm  Kchi     n    delta  [Kchi n delta]
1937 + CT      CT      CT      C       Charmm  0.156    3    0.00
1938 + OH      CT      CT      OH      Charmm  0.144    3    0.00    1.175 2  0
1939 + HC      CT      CM      CM      Charmm  1.150    1    0.00    0.38  3 180
1940 + //Quartic
1941 + //Atom1 Atom2   Atom3   Atom4   Quartic          k4    k3    k2    k1    k0
1942 + //Polynomial
1943 + //Atom1 Atom2   Atom3   Atom4   Polynomial  n Kn     [m  Km]
1944 + S       CH2     CH2     C       Polynomial  0 2.218   1  2.905  2 -3.136  3 -0.7313  4 6.272  5 -7.528
1945 + end TorsionTypes
1946 + \end{code}
1947  
1948 < (see Fig.~\ref{fig:lipidModel}), The parameters for $k_{\theta}$ and
1949 < $\theta_0$ are borrowed from those in TraPPE.\cite{Siepmann1998}
1948 > Note that the parameters for a particular torsion type are the same
1949 > for any torsional quartet of the same atomic types (in the same or
1950 > reversed order).
1951  
1952 < Calculating the long-range (non-bonded) potential involves a sum over
1624 < all pairs of atoms (except for those atoms which are involved in a
1625 < bond, bend, or torsion with each other).  If done poorly, calculating
1626 < the the long-range interactions for $N$ atoms would involve $N(N-1)/2$
1627 < evaluations of atomic distances.  To reduce the number of distance
1628 < evaluations between pairs of atoms, {\sc OpenMD} allows the use of
1629 < switched cutoffs with Verlet neighbor lists.\cite{Allen87} Neutral
1630 < groups which contain charges will exhibit pathological forces unless
1631 < the cutoff is applied to the neutral groups evenly instead of to the
1632 < individual atoms.\cite{leach01:mm}  {\sc OpenMD} allows users to
1633 < specify cutoff groups which may contain an arbitrary number of atoms
1634 < in the molecule.  Atoms in a cutoff group are treated as a single unit
1635 < for the evaluation of the switching function:
1636 < \begin{equation}
1637 < V_{\mathrm{long-range}} = \sum_{a} \sum_{b>a} s(r_{ab}) \sum_{i \in a} \sum_{j \in b} V_{ij}(r_{ij}),
1638 < \end{equation}
1639 < where $r_{ab}$ is the distance between the centers of mass of the two
1640 < cutoff groups ($a$ and $b$).
1952 > \subsection{\label{section:ffInversion}The InversionTypes block}
1953  
1954 < The sums over $a$ and $b$ are over the cutoff groups that are present
1955 < in the simulation.  Atoms which are not explicitly defined as members
1956 < of a {\tt cutoffGroup} are treated as a group consisting of only one
1957 < atom.  The switching function, $s(r)$ is the standard cubic switching
1958 < function,
1954 > Inversion potentials are often added to force fields to enforce
1955 > planarity around $sp^2$-hybridized carbons or to correct vibrational
1956 > frequencies for umbrella-like vibrational modes for central atoms
1957 > bonded to exactly three satellite groups.
1958 >
1959 > In OpenMD's version of an inversion, the central atom is entered {\it
1960 >  first} in each line in the {\tt InversionTypes} block. Note that
1961 > this is quite different than how other codes treat Improper torsional
1962 > potentials to mimic inversion behavior.  In some other widely-used
1963 > simulation packages, the central atom is treated as atom 3 in a
1964 > standard torsion form:
1965 > \begin{itemize}
1966 >  \item OpenMD:  I - (J - K - L)  (e.g. I is $sp^2$ hybridized carbon)
1967 >  \item AMBER:   I - J - K - L   (e.g. K is $sp^2$ hybridized carbon)
1968 > \end{itemize}
1969 >
1970 > The inversion angle itself is defined as:
1971   \begin{equation}
1972 < S(r) =
1973 <        \begin{cases}
1974 <        1 & \text{if $r \le r_{\text{sw}}$},\\
1651 <        \frac{(r_{\text{cut}} + 2r - 3r_{\text{sw}})(r_{\text{cut}} - r)^2}
1652 <        {(r_{\text{cut}} - r_{\text{sw}})^3}
1653 <        & \text{if $r_{\text{sw}} < r \le r_{\text{cut}}$}, \\
1654 <        0 & \text{if $r > r_{\text{cut}}$.}
1655 <        \end{cases}
1656 < \label{eq:dipoleSwitching}
1972 > \cos\omega_{i-jkl} = \left(\hat{\mathbf{r}}_{il} \times
1973 >  \hat{\mathbf{r}}_{ij}\right)\cdot\left( \hat{\mathbf{r}}_{il} \times
1974 >  \hat{\mathbf{r}}_{ik}\right)
1975   \end{equation}
1976 < Here, $r_{\text{sw}}$ is the {\tt switchingRadius}, or the distance
1977 < beyond which interactions are reduced, and $r_{\text{cut}}$ is the
1978 < {\tt cutoffRadius}, or the distance at which interactions are
1979 < truncated.
1976 > Here, $\hat{\mathbf{r}}_{\alpha\beta}$ are the set of unit bond
1977 > vectors between the central atoms $i$, and the satellite atoms $j$,
1978 > $k$, and $l$.  Note that other definitions of inversion angles are
1979 > possible, so users are encouraged to be particularly careful when
1980 > converting other force field files for use with OpenMD.
1981  
1982 < Users of {\sc OpenMD} do not need to specify the {\tt cutoffRadius} or
1983 < {\tt switchingRadius}.  In simulations containing only Lennard-Jones
1984 < atoms, the cutoff radius has a default value of $2.5\sigma_{ii}$,
1985 < where $\sigma_{ii}$ is the largest Lennard-Jones length parameter
1986 < present in the simulation.  In simulations containing charged or
1987 < dipolar atoms, the default cutoff radius is $15 \mbox{\AA}$.  
1982 > There are many common ways to create planarity or umbrella behavior in
1983 > a potential energy function, and OpenMD implements a number of the
1984 > more common functions:
1985 > \begin{itemize}
1986 > \item {\tt ImproperCosine}:
1987 > \begin{equation*}
1988 > V_{\text{torsion}}(\omega) = \sum_n \frac{K_n}{2} \left( 1 + cos(n
1989 >  \omega - \delta_n) \right),
1990 > \end{equation*}
1991 > \item {\tt AmberImproper}:
1992 > \begin{equation*}
1993 >  V_{\text{torsion}}(\omega) =  \frac{v}{2} (1 - \cos\left(2 \left(\omega - \omega_0\right)\right),
1994 > \end{equation*}
1995 > \item {\tt Harmonic}:
1996 > \begin{equation*}
1997 > V_{\text{torsion}}(\omega) =  \frac{d}{2} \left(\omega - \omega_0\right).
1998 > \end{equation*}
1999 > \end{itemize}
2000 > \begin{code}[caption={[An example of an InversionTypes block.] A
2001 > simple example of a InversionTypes block.  Angles ($\delta_n$ and
2002 > $\omega_0$) angles are given in degrees, while energy parameters ($v,
2003 > K_n$) are given in kcal / mol.   The Harmonic Inversion type has a
2004 > force constant that must be given in kcal/mol/degrees$^2$.},
2005 > label={sch:InversionTypes}]
2006 > begin InversionTypes
2007 > //Harmonic
2008 > //Atom1 Atom2   Atom3   Atom4   Harmonic  d(kcal/mol/deg^2)  omega0
2009 > RCHar3  X       X       X       Harmonic  1.21876e-2         180.0
2010 > //AmberImproper
2011 > //Atom1 Atom2   Atom3   Atom4   AmberImproper   v(kcal/mol)
2012 > C       CT      N       O       AmberImproper   10.500000
2013 > CA      CA      CA      CT      AmberImproper   1.100000
2014 > //ImproperCosine
2015 > //Atom1 Atom2   Atom3   Atom4   ImproperCosine  Kn  n  delta_n  [Kn n delta_n]
2016 > end InversionTypes
2017 > \end{code}
2018  
2019 < The {\tt switchingRadius} is set to a default value of 95\% of the
1671 < {\tt cutoffRadius}.  In the special case of a simulation containing
1672 < {\it only} Lennard-Jones atoms, the default switching radius takes the
1673 < same value as the cutoff radius, and {\sc OpenMD} will use a shifted
1674 < potential to remove discontinuities in the potential at the cutoff.
1675 < Both radii may be specified in the meta-data file.
2019 > \section{\label{section::ffLongRange}Long Range Interactions}
2020  
2021 < Force fields can be added to {\sc OpenMD}, although it comes with a few
2022 < simple examples (Lennard-Jones, {\sc duff}, {\sc water}, and {\sc
2023 < eam}) which are explained in the following sections.
2021 > Calculating the long-range (non-bonded) potential involves a sum over
2022 > all pairs of atoms (except for those atoms which are involved in a
2023 > bond, bend, or torsion with each other).  Many of these interactions
2024 > can be inferred from the AtomTypes,
2025  
2026 < \section{\label{sec:LJPot}The Lennard Jones Force Field}
2026 > \subsection{\label{section:ffNBinteraction}The NonBondedInteractions
2027 >  block}
2028  
2029 < Scheme
2030 < \ref{sch:LJFF} gives an example meta-data file that
2031 < sets up a system of 108 Ar particles to be simulated using the
2032 < Lennard-Jones force field.
2029 > The user might want like to specify explicit or special interactions
2030 > that override the default non-bodned interactions that are inferred
2031 > from the AtomTypes.  To do this, OpenMD implements a
2032 > NonBondedInteractions block to allow the user to specify the following
2033 > (pair-wise) non-bonded interactions:
2034  
2035 < \begin{lstlisting}[float,caption={[Invocation of the Lennard-Jones
2036 < force field] A sample startup file for a small Lennard-Jones
2037 < simulation.},label={sch:LJFF}]
2038 < <OpenMD>
2039 <  <MetaData>
2040 < #include "argon.md"
2035 > \begin{itemize}
2036 > \item {\tt LennardJones}:
2037 > \begin{equation*}
2038 > V_{\text{NB}}(r) = 4 \epsilon_{ij} \left(
2039 >  \left(\frac{\sigma_{ij}}{r} \right)^{12} -
2040 >  \left(\frac{\sigma_{ij}}{r} \right)^{6} \right),
2041 > \end{equation*}
2042 > \item {\tt ShiftedMorse}:
2043 > \begin{equation*}
2044 > V_{\text{NB}}(r) = D_{ij} \left( e^{-2 \beta_{ij} (r -
2045 >     r^0)} - 2 e^{- \beta_{ij} (r -
2046 >     r^0)} \right),
2047 > \end{equation*}
2048 > \item {\tt RepulsiveMorse}:
2049 > \begin{equation*}
2050 > V_{\text{NB}}(r) = D_{ij} \left( e^{-2 \beta_{ij} (r -
2051 >     r^0)} \right),
2052 > \end{equation*}
2053 > \item {\tt RepulsivePower}:
2054 > \begin{equation*}
2055 >  V_{\text{NB}}(r) = \epsilon_{ij}
2056 >  \left(\frac{\sigma_{ij}}{r} \right)^{n_{ij}}.
2057 > \end{equation*}
2058 > \end{itemize}
2059  
2060 < component{
2061 <  type = "Ar";
2062 <  nMol = 108;
2063 < }
2060 > \begin{code}[caption={[An example of a NonBondedInteractions block.] A
2061 > simple example of a NonBondedInteractions block. Distances ($\sigma,
2062 > r_0$) are given in \AA, while energies ($\epsilon, D0$) are in
2063 > kcal/mol.  The Morse potentials have an additional parameter $\beta_0$
2064 > which is in units of \AA$^{-1}$.},
2065 > label={sch:InversionTypes}]
2066 > begin NonBondedInteractions
2067  
2068 < forceField = "LJ";
2069 <  </MetaData>
2070 <  <Snapshot>     // not shown in this scheme
2071 <  </Snapshot>
2072 < </OpenMD>
2073 < \end{lstlisting}
2068 > //Lennard-Jones
2069 > //Atom1 Atom2   LennardJones    sigma  epsilon
2070 > Au      CH3     LennardJones    3.54   0.2146
2071 > Au      CH2     LennardJones    3.54   0.1749
2072 > Au      CH      LennardJones    3.54   0.1749
2073 > Au      S       LennardJones    2.40   8.465
2074  
2075 + //Shifted Morse
2076 + //Atom1 Atom2   ShiftedMorse    r0     D0       beta0
2077 + Au      O_SPCE  ShiftedMorse    3.70   0.0424   0.769
2078  
2079 < \section{\label{section:DUFF}Dipolar Unified-Atom Force Field}
2079 > //Repulsive Morse
2080 > //Atom1 Atom2   RepulsiveMorse  r0     D0       beta0
2081 > Au      H_SPCE  RepulsiveMorse  -1.00  0.00850  0.769
2082  
2083 < The dipolar unified-atom force field ({\sc duff}) was developed to
2084 < simulate lipid bilayers. These types of simulations require a model
2085 < capable of forming bilayers, while still being sufficiently
2086 < computationally efficient to allow large systems ($\sim$100's of
2087 < phospholipids, $\sim$1000's of waters) to be simulated for long times
2088 < ($\sim$10's of nanoseconds). With this goal in mind, {\sc duff} has no
1716 < point charges. Charge-neutral distributions are replaced with dipoles,
1717 < while most atoms and groups of atoms are reduced to Lennard-Jones
1718 < interaction sites. This simplification reduces the length scale of
1719 < long range interactions from $\frac{1}{r}$ to $\frac{1}{r^3}$,
1720 < removing the need for the computationally expensive Ewald
1721 < sum. Instead, Verlet neighbor-lists and cutoff radii are used for the
1722 < dipolar interactions, and, if desired, a reaction field may be added
1723 < to mimic longer range interactions.
1724 <
1725 < As an example, lipid head-groups in {\sc duff} are represented as
1726 < point dipole interaction sites.  Placing a dipole at the head group's
1727 < center of mass mimics the charge separation found in common
1728 < phospholipid head groups such as phosphatidylcholine.\cite{Cevc87}
1729 < Additionally, a large Lennard-Jones site is located at the
1730 < pseudoatom's center of mass. The model is illustrated by the red atom
1731 < in Fig.~\ref{fig:lipidModel}. The water model we use to
1732 < complement the dipoles of the lipids is a
1733 < reparameterization\cite{fennell04} of the soft sticky dipole (SSD)
1734 < model of Ichiye
1735 < \emph{et al.}\cite{liu96:new_model}
1736 <
1737 < \begin{figure}
1738 < \centering
1739 < \includegraphics[width=\linewidth]{lipidModel.pdf}
1740 < \caption[A representation of a lipid model in {\sc duff}]{A
1741 < representation of the lipid model. $\phi$ is the torsion angle,
1742 < $\theta$ is the bend angle, and $\mu$ is the dipole moment of the head
1743 < group.}
1744 < \label{fig:lipidModel}
1745 < \end{figure}
1746 <
1747 < A set of scalable parameters has been used to model the alkyl groups
1748 < with Lennard-Jones sites. For this, parameters from the TraPPE force
1749 < field of Siepmann \emph{et al.}\cite{Siepmann1998} have been
1750 < utilized. TraPPE is a unified-atom representation of n-alkanes which
1751 < is parametrized against phase equilibria using Gibbs ensemble Monte
1752 < Carlo simulation techniques.\cite{Siepmann1998} One of the advantages
1753 < of TraPPE is that it generalizes the types of atoms in an alkyl chain
1754 < to keep the number of pseudoatoms to a minimum; thus, the parameters
1755 < for a unified atom such as $\text{CH}_2$ do not change depending on
1756 < what species are bonded to it.
1757 <
1758 < As is required by TraPPE, {\sc duff} also constrains all bonds to be
1759 < of fixed length. Typically, bond vibrations are the fastest motions in
1760 < a molecular dynamic simulation.  With these vibrations present, small
1761 < time steps between force evaluations must be used to ensure adequate
1762 < energy conservation in the bond degrees of freedom. By constraining
1763 < the bond lengths, larger time steps may be used when integrating the
1764 < equations of motion. A simulation using {\sc duff} is illustrated in
1765 < Scheme \ref{sch:DUFF}.
1766 <
1767 < \begin{lstlisting}[float,caption={[Invocation of {\sc duff}]A portion
1768 < of a startup file showing a simulation utilizing {\sc
1769 < duff}},label={sch:DUFF}]  
1770 < <OpenMD>
1771 <  <MetaData>
1772 < #include "water.md"
1773 < #include "lipid.md"
1774 <
1775 < component{
1776 <  type = "simpleLipid_16";
1777 <  nMol = 60;
1778 < }
1779 <
1780 < component{
1781 <  type = "SSD_water";
1782 <  nMol = 1936;
1783 < }
1784 <
1785 < forceField = "DUFF";
1786 <  </MetaData>
1787 <  <Snapshot>     // not shown in this scheme
1788 <  </Snapshot>
1789 < </OpenMD>
1790 < \end{lstlisting}
1791 <
1792 <
1793 <
1794 < The cross potential between molecules $I$ and $J$,
1795 < $V^{IJ}_{\text{Cross}}$, is as follows:
1796 < \begin{equation}
1797 < V^{IJ}_{\text{Cross}} =
1798 <        \sum_{i \in I} \sum_{j \in J}
1799 <        \biggl[ V_{\text{LJ}}(r_{ij}) +  V_{\text{dipole}}
1800 <        (\mathbf{r}_{ij},\boldsymbol{\Omega}_{i},\boldsymbol{\Omega}_{j})
1801 <        + V_{\text{sticky}}
1802 <        (\mathbf{r}_{ij},\boldsymbol{\Omega}_{i},\boldsymbol{\Omega}_{j})
1803 <        \biggr],
1804 < \label{eq:crossPotentail}
1805 < \end{equation}
1806 < where $V_{\text{LJ}}$ is the Lennard Jones potential,
1807 < $V_{\text{dipole}}$ is the dipole dipole potential, and
1808 < $V_{\text{sticky}}$ is the sticky potential defined by the SSD model
1809 < (Sec.~\ref{section:SSD}). Note that not all atom types include all
1810 < interactions.
1811 <
1812 <
1813 < \section{\label{section:WATER}The {\sc water} Force Field}
1814 <
1815 < In addition to the {\sc duff} force field's solvent description, a
1816 < separate {\sc water} force field has been included for simulating most
1817 < of the common rigid-body water models. This force field includes the
1818 < simple and point-dipolar models (SSD, SSD1, SSD/E, SSD/RF, and DPD
1819 < water), as well as the common charge-based models (SPC, SPC/E, TIP3P,
1820 < TIP4P, and
1821 < TIP5P).\cite{liu96:new_model,Ichiye03,fennell04,Marrink01,Berendsen81,Berendsen87,Jorgensen83,Mahoney00}
1822 < In order to handle these models, charge-charge interactions were
1823 < included in the force-loop:
1824 < \begin{equation}
1825 < V_{\text{charge}}(r_{ij}) = \sum_{ij}\frac{q_iq_je^2}{r_{ij}},
1826 < \end{equation}
1827 < where $q$ represents the charge on particle $i$ or $j$, and $e$ is the
1828 < charge of an electron in Coulombs. The charge-charge interaction
1829 < support is rudimentary in the current version of {\sc OpenMD}.  As with
1830 < the other pair interactions, charges can be simulated with a pure
1831 < cutoff or a reaction field.  The various methods for performing the
1832 < Ewald summation have not yet been included.  The {\sc water} force
1833 < field can be easily expanded through modification of the {\sc water}
1834 < force field file ({\tt WATER.frc}). By adding atom types and inserting
1835 < the appropriate parameters, it is possible to extend the force field
1836 < to handle rigid molecules other than water.
2083 > //Repulsive Power
2084 > //Atom1 Atom2   RepulsivePower   sigma    epsilon    n
2085 > Au      ON      RepulsivePower   3.47005  0.186208   11
2086 > Au      NO      RepulsivePower   3.53955  0.168629   11
2087 > end NonBondedInteractions
2088 > \end{code}
2089  
1838
1839 \section{\label{section:sc}The Sutton-Chen Force Field}
1840
1841
1842 \section{\label{section:clay}The CLAY force field}
1843
1844 The {\sc clay} force field is based on an ionic (nonbonded)
1845 description of the metal-oxygen interactions associated with hydrated
1846 phases. All atoms are represented as point charges and are allowed
1847 complete translational freedom. Metal-oxygen interactions are based on
1848 a simple Lennard-Jones potential combined with electrostatics. The
1849 empirical parameters were optimized by Cygan {\it et
1850 al.}\cite{Cygan04} on the basis of known mineral structures, and
1851 partial atomic charges were derived from periodic DFT quantum chemical
1852 calculations of simple oxide, hydroxide, and oxyhydroxide model
1853 compounds with well-defined structures.
1854
1855
2090   \section{\label{section:electrostatics}Electrostatics}
2091  
2092 < To aid in performing simulations in more traditional force fields, we
2093 < have added routines to carry out electrostatic interactions using a
2094 < number of different electrostatic summation methods.  These methods
2095 < are extended from the damped and cutoff-neutralized Coulombic sum
2096 < originally proposed by Wolf, {\it et al.}\cite{Wolf99} One of these,
2097 < the damped shifted force method, shows a remarkable ability to
2098 < reproduce the energetic and dynamic characteristics exhibited by
2099 < simulations employing lattice summation techniques.  The basic idea is
2100 < to construct well-behaved real-space summation methods using two tricks:
2092 > Because nearly all force fields involve electrostatic interactions in
2093 > one form or another, OpenMD implements a number of different
2094 > electrostatic summation methods.  These methods are extended from the
2095 > damped and cutoff-neutralized Coulombic sum originally proposed by
2096 > Wolf, {\it et al.}\cite{Wolf99} One of these, the damped shifted force
2097 > method, shows a remarkable ability to reproduce the energetic and
2098 > dynamic characteristics exhibited by simulations employing lattice
2099 > summation techniques.  The basic idea is to construct well-behaved
2100 > real-space summation methods using two tricks:
2101   \begin{enumerate}
2102   \item shifting through the use of image charges, and
2103   \item damping the electrostatic interaction.
# Line 1982 | Line 2216 | the shifted potential (eq. (\ref{eq:SPPot})) becomes
2216   \end{equation}
2217   the shifted potential (eq. (\ref{eq:SPPot})) becomes
2218   \begin{equation}
2219 < V_{\textrm{DSP}}(r) = q_iq_j\left(\frac{\textrm{erfc}\left(\alpha r\right)}{r}-\
1986 < frac{\textrm{erfc}\left(\alpha R_\textrm{c}\right)}{R_\textrm{c}}\right) \quad r
2219 > V_{\textrm{DSP}}(r) = q_iq_j\left(\frac{\textrm{erfc}\left(\alpha r\right)}{r}-\frac{\textrm{erfc}\left(\alpha R_\textrm{c}\right)}{R_\textrm{c}}\right) \quad r
2220   \leqslant R_\textrm{c},
2221   \label{eq:DSPPot}
2222   \end{equation}
# Line 2028 | Line 2261 | this reason, the default electrostatic summation metho
2261   {\sc OpenMD} is the DSF (Eq. \ref{eq:DSFPot}) with a damping parameter
2262   ($\alpha$) that is set algorithmically from the cutoff radius.
2263  
2264 +
2265 + \section{\label{section:cutoffGroups}Switching Functions}
2266 +
2267 + Calculating the the long-range interactions for $N$ atoms involves
2268 + $N(N-1)/2$ evaluations of atomic distances if it is done in a brute
2269 + force manner.  To reduce the number of distance evaluations between
2270 + pairs of atoms, {\sc OpenMD} allows the use of hard or switched
2271 + cutoffs with Verlet neighbor lists.\cite{Allen87} Neutral groups which
2272 + contain charges can exhibit pathological forces unless the cutoff is
2273 + applied to the neutral groups evenly instead of to the individual
2274 + atoms.\cite{leach01:mm} {\sc OpenMD} allows users to specify cutoff
2275 + groups which may contain an arbitrary number of atoms in the molecule.
2276 + Atoms in a cutoff group are treated as a single unit for the
2277 + evaluation of the switching function:
2278 + \begin{equation}
2279 + V_{\mathrm{long-range}} = \sum_{a} \sum_{b>a} s(r_{ab}) \sum_{i \in a} \sum_{j \in b} V_{ij}(r_{ij}),
2280 + \end{equation}
2281 + where $r_{ab}$ is the distance between the centers of mass of the two
2282 + cutoff groups ($a$ and $b$).
2283 +
2284 + The sums over $a$ and $b$ are over the cutoff groups that are present
2285 + in the simulation.  Atoms which are not explicitly defined as members
2286 + of a {\tt cutoffGroup} are treated as a group consisting of only one
2287 + atom.  The switching function, $s(r)$ is the standard cubic switching
2288 + function,
2289 + \begin{equation}
2290 + S(r) =
2291 +        \begin{cases}
2292 +        1 & \text{if $r \le r_{\text{sw}}$},\\
2293 +        \frac{(r_{\text{cut}} + 2r - 3r_{\text{sw}})(r_{\text{cut}} - r)^2}
2294 +        {(r_{\text{cut}} - r_{\text{sw}})^3}
2295 +        & \text{if $r_{\text{sw}} < r \le r_{\text{cut}}$}, \\
2296 +        0 & \text{if $r > r_{\text{cut}}$.}
2297 +        \end{cases}
2298 + \label{eq:dipoleSwitching}
2299 + \end{equation}
2300 + Here, $r_{\text{sw}}$ is the {\tt switchingRadius}, or the distance
2301 + beyond which interactions are reduced, and $r_{\text{cut}}$ is the
2302 + {\tt cutoffRadius}, or the distance at which interactions are
2303 + truncated.  
2304 +
2305 + Users of {\sc OpenMD} do not need to specify the {\tt cutoffRadius} or
2306 + {\tt switchingRadius}.  
2307 + If the {\tt cutoffRadius} was not explicitly set, OpenMD will attempt
2308 + to guess an appropriate choice.  If the system contains electrostatic
2309 + atoms, the default cutoff radius is 12 \AA.  In systems without
2310 + electrostatic (charge or multipolar) atoms, the atom types present in the simulation will be
2311 + polled for suggested cutoff values (e.g. $2.5 max(\left\{ \sigma
2312 +  \right\})$ for Lennard-Jones atoms.   The largest suggested value
2313 + that was found will be used.
2314 +
2315 + By default, OpenMD uses shifted force potentials to force the
2316 + potential energy and forces to smoothly approach zero at the cutoff
2317 + radius.  If the user would like to use another cutoff method
2318 + they may do so by setting the {\tt cutoffMethod} parameter to:
2319 + \begin{itemize}
2320 + \item {\tt HARD}
2321 + \item {\tt SWITCHED}
2322 + \item {\tt SHIFTED\_FORCE} (default):
2323 + \item {\tt TAYLOR\_SHIFTED}
2324 + \item {\tt SHIFTED\_POTENTIAL}
2325 + \end{itemize}
2326 +
2327 + The {\tt switchingRadius} is set to a default value of 95\% of the
2328 + {\tt cutoffRadius}.  In the special case of a simulation containing
2329 + {\it only} Lennard-Jones atoms, the default switching radius takes the
2330 + same value as the cutoff radius, and {\sc OpenMD} will use a shifted
2331 + potential to remove discontinuities in the potential at the cutoff.
2332 + Both radii may be specified in the meta-data file.
2333 +
2334 +
2335   \section{\label{section:pbc}Periodic Boundary Conditions}
2336  
2337   \newcommand{\roundme}{\operatorname{round}}

Diff Legend

Removed lines
+ Added lines
< Changed lines
> Changed lines