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 3708 by kstocke1, Mon Nov 22 22:34:45 2010 UTC vs.
Revision 4105 by gezelter, Mon Apr 28 21:41:01 2014 UTC

# Line 9 | Line 9
9   \usepackage{longtable}
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 17 | Line 20
20   \textwidth 6.5in
21   \brokenpenalty=10000
22   \renewcommand{\baselinestretch}{1.2}
23 + \usepackage[square, comma, sort&compress]{natbib}
24 + \bibpunct{[}{]}{,}{n}{}{;}
25  
26 + \DeclareFloatFont{tiny}{\scriptsize}% "scriptsize" is defined by floatrow, "tiny" not
27 + \floatsetup[table]{font=tiny}
28 +
29 +
30   %\renewcommand\citemid{\ } % no comma in optional reference note
31 < \lstset{language=C,frame=TB,basicstyle=\tiny,basicstyle=\ttfamily, %
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 38 | Line 57
57   \newcolumntype{H}{p{0.75in}}
58   \newcolumntype{I}{p{5in}}
59  
60 + \newcolumntype{J}{p{1.5in}}
61 + \newcolumntype{K}{p{1.2in}}
62 + \newcolumntype{L}{p{1.5in}}
63 + \newcolumntype{M}{p{1.55in}}
64  
42 \title{{\sc OpenMD}: Molecular Dynamics in the Open}
65  
66 < \author{Shenyu Kuang, Chunlei Li, Charles F. Vardeman II, \\
45 < Teng Lin, Christopher J. Fennell,  Xiuquan Sun, \\
46 < Kyle Daily, Yang Zheng, Matthew A. Meineke, and J. Daniel Gezelter\\
47 < Department of Chemistry and Biochemistry\\
48 < University of Notre Dame\\
49 < Notre Dame, Indiana 46556}
66 > \title{{\sc OpenMD-2.2}: Molecular Dynamics in the Open}
67  
68 + \author{Joseph Michalka, James Marr, Kelsey Stocker, Madan Lamichhane,
69 +  Patrick Louden, \\
70 +  Teng Lin, Charles F. Vardeman II, Christopher J. Fennell, Shenyu
71 +  Kuang, Xiuquan Sun, \\
72 +  Chunlei Li, Kyle Daily, Yang Zheng, Matthew A. Meineke, and \\
73 +  J. Daniel Gezelter \\
74 +  Department of Chemistry and Biochemistry\\
75 +  University of Notre Dame\\
76 +  Notre Dame, Indiana 46556}
77 +
78   \maketitle
79  
80   \section*{Preface}
# Line 132 | Line 159 | leave an interaction region.
159   leave an interaction region.
160  
161   {\tt Atoms} may also be grouped in more traditional ways into {\tt
162 < bonds}, {\tt bends}, and {\tt torsions}.  These groupings allow the
163 < correct choice of interaction parameters for short-range interactions
164 < to be chosen from the definitions in the {\tt forceField}.
162 >  bonds}, {\tt bends}, {\tt torsions}, and {\tt inversions}.  These
163 > groupings allow the correct choice of interaction parameters for
164 > short-range interactions to be chosen from the definitions in the {\tt
165 >  forceField}.
166  
167   All of these groups of {\tt atoms} are brought together in the {\tt
168   molecule}, which is the fundamental structure for setting up and {\sc
# Line 169 | 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.},
179 < 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 202 | 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 218 | 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 257 | 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 271 | 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 280 | 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 315 | 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 386 | 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 412 | 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 490 | Line 517 | fs}^{-1}$), and body-fixed moments of inertia ($\mbox{
517   \endhead
518   \hline
519   \endfoot
520 < {\tt forceField} & string & Sets the force field. & Possible force
521 < fields are DUFF, WATER, LJ, EAM, SC, and CLAY. \\
520 > {\tt forceField} & string & Sets the base name for the force field file &
521 > OpenMD appends a {\tt .frc} to the end of this to look for a force
522 > field file.\\
523   {\tt component} & & Defines the molecular components of the system &
524   Every {\tt $<$MetaData$>$} block must have a component statement. \\
525   {\tt minimizer} & string & Chooses a minimizer & Possible minimizers
526   are SD and CG. Either {\tt ensemble} or {\tt minimizer} must be specified. \\
527   {\tt ensemble} & string & Sets the ensemble. & Possible ensembles are
528 < NVE, NVT, NPTi, NPAT, NPTf, NPTxyz, LD and LHull.  Either {\tt ensemble}
528 > NVE, NVT, NPTi, NPAT, NPTf, NPTxyz, LD and LangevinHull.  Either {\tt ensemble}
529   or {\tt minimizer} must be specified. \\
530   {\tt dt} & fs & Sets the time step. & Selection of {\tt dt} should be
531   small enough to sample the fastest motion of the simulation. ({\tt
# Line 561 | Line 589 | column names are: {\sc time, total\_energy, potential\
589   default is the first eight of these columns in order.)  \\
590   & & \multicolumn{2}{p{3.5in}}{Allowed
591   column names are: {\sc time, total\_energy, potential\_energy, kinetic\_energy,
592 < temperature, pressure, volume, conserved\_quantity,
592 > temperature, pressure, volume, conserved\_quantity, hullvolume, gyrvolume,
593   translational\_kinetic, rotational\_kinetic,  long\_range\_potential,
594   short\_range\_potential, vanderwaals\_potential,
595 < electrostatic\_potential, bond\_potential, bend\_potential,
596 < dihedral\_potential, improper\_potential, vraw, vharm,
597 < pressure\_tensor\_x, pressure\_tensor\_y, pressure\_tensor\_z}} \\
595 > electrostatic\_potential, metallic\_potential,
596 > hydrogen\_bonding\_potential, bond\_potential, bend\_potential,
597 > dihedral\_potential, inversion\_potential, raw\_potential, restraint\_potential,
598 > pressure\_tensor, system\_dipole, heatflux, electronic\_temperature}} \\
599   {\tt printPressureTensor} & logical & sets whether {\sc OpenMD} will print
600   out the pressure tensor & can be useful for calculations of the bulk
601   modulus \\
# Line 618 | 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 627 | 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.},
635 < label=sch:dumpFormat]
663 > lastly, body fixed angular momentum for that integrable object.},label={sch:dumpFormat}]
664    <Snapshot>
665      <FrameData>
666          Time: 0
# Line 647 | 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 698 | 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>
736   molecule{
737    name = "I2";
738 <  atom[0]{
739 <    type = "I";
740 <  }
713 <  atom[1]{
714 <    type = "I";
715 <  }
716 <  bond{
717 <    members( 0, 1);
718 <  }
738 >  atom[0]{ type = "I"; }
739 >  atom[1]{ type = "I"; }
740 >  bond{ members( 0, 1); }
741   }
742   molecule{
743    name = "HCl"
744 <  atom[0]{
745 <    type = "H";
746 <  }
725 <  atom[1]{
726 <    type = "Cl";
727 <  }
728 <  bond{
729 <    members( 0, 1);
730 <  }
744 >  atom[0]{ type = "H";}
745 >  atom[1]{ type = "Cl";}
746 >  bond{ members( 0, 1); }
747   }
748   component{
749    type = "HCl";
# Line 757 | Line 773 | component{
773      </StuntDoubles>
774    </Snapshot>
775   </OpenMD>
776 < \end{lstlisting}
776 > \end{code}
777  
778   \section{The Statistics File}
779  
# Line 773 | 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:empiricalEnergy}The Empirical Energy
777 < Functions}
792 > \chapter{\label{chapter:forceFields}Force Fields}
793  
794 < Like many simulation packages, {\sc OpenMD} splits the potential energy
795 < into the short-ranged (bonded) portion and a long-range (non-bonded)
796 < potential,
794 > Like many molecular simulation packages, {\sc OpenMD} splits the
795 > potential energy into the short-ranged (bonded) portion and a
796 > long-range (non-bonded) potential,
797   \begin{equation}
798   V = V_{\mathrm{short-range}} + V_{\mathrm{long-range}}.
799   \end{equation}
800 < The short-ranged portion includes the explicit bonds, bends, and
801 < torsions which have been defined in the meta-data file for the
802 < molecules which are present in the simulation.  The functional forms and
803 < parameters for these interactions are defined by the force field which
804 < is chosen.
800 > The short-ranged portion includes the bonds, bends, torsions, and
801 > inversions which have been defined in the meta-data file for the
802 > molecules.  The functional forms and parameters for these interactions
803 > are defined by the force field which is selected in the MetaData
804 > section.
805  
806 < Calculating the long-range (non-bonded) potential involves a sum over
807 < all pairs of atoms (except for those atoms which are involved in a
808 < bond, bend, or torsion with each other).  If done poorly, calculating
809 < the the long-range interactions for $N$ atoms would involve $N(N-1)/2$
810 < evaluations of atomic distances.  To reduce the number of distance
796 < evaluations between pairs of atoms, {\sc OpenMD} uses a switched cutoff
797 < with Verlet neighbor lists.\cite{Allen87} It is well known that
798 < neutral groups which contain charges will exhibit pathological forces
799 < unless the cutoff is applied to the neutral groups evenly instead of
800 < to the individual atoms.\cite{leach01:mm} {\sc OpenMD} allows users to
801 < specify cutoff groups which may contain an arbitrary number of atoms
802 < in the molecule.  Atoms in a cutoff group are treated as a single unit
803 < for the evaluation of the switching function:
806 > \section{\label{section:divisionOfLabor}Separation into Internal and
807 >  Cross interactions}
808 >
809 > The classical potential energy function for a system composed of $N$
810 > molecules is traditionally written
811   \begin{equation}
812 < V_{\mathrm{long-range}} = \sum_{a} \sum_{b>a} s(r_{ab}) \sum_{i \in a} \sum_{j \in b} V_{ij}(r_{ij}),
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 $r_{ab}$ is the distance between the centers of mass of the two
817 < cutoff groups ($a$ and $b$).
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 sums over $a$ and $b$ are over the cutoff groups that are present
824 < in the simulation.  Atoms which are not explicitly defined as members
825 < of a {\tt cutoffGroup} are treated as a group consisting of only one
826 < atom.  The switching function, $s(r)$ is the standard cubic switching
827 < function,
815 < \begin{equation}
816 < S(r) =
817 <        \begin{cases}
818 <        1 & \text{if $r \le r_{\text{sw}}$},\\
819 <        \frac{(r_{\text{cut}} + 2r - 3r_{\text{sw}})(r_{\text{cut}} - r)^2}
820 <        {(r_{\text{cut}} - r_{\text{sw}})^3}
821 <        & \text{if $r_{\text{sw}} < r \le r_{\text{cut}}$}, \\
822 <        0 & \text{if $r > r_{\text{cut}}$.}
823 <        \end{cases}
824 < \label{eq:dipoleSwitching}
825 < \end{equation}
826 < Here, $r_{\text{sw}}$ is the {\tt switchingRadius}, or the distance
827 < beyond which interactions are reduced, and $r_{\text{cut}}$ is the
828 < {\tt cutoffRadius}, or the distance at which interactions are
829 < truncated.
823 > The types of atoms being simulated, as well as the specific functional
824 > forms and parameters of the intra-molecular functions and the
825 > long-range potentials are defined by the force field. In the following
826 > sections we discuss the stucture of an OpenMD force field file and the
827 > specification of blocks that may be present within these files.
828  
829 < Users of {\sc OpenMD} do not need to specify the {\tt cutoffRadius} or
832 < {\tt switchingRadius}.  In simulations containing only Lennard-Jones
833 < atoms, the cutoff radius has a default value of $2.5\sigma_{ii}$,
834 < where $\sigma_{ii}$ is the largest Lennard-Jones length parameter
835 < present in the simulation.  In simulations containing charged or
836 < dipolar atoms, the default cutoff radius is $15 \mbox{\AA}$.  
829 > \section{\label{section:frcFile}Force Field Files}
830  
831 < The {\tt switchingRadius} is set to a default value of 95\% of the
832 < {\tt cutoffRadius}.  In the special case of a simulation containing
833 < {\it only} Lennard-Jones atoms, the default switching radius takes the
834 < same value as the cutoff radius, and {\sc OpenMD} will use a shifted
835 < potential to remove discontinuities in the potential at the cutoff.
836 < Both radii may be specified in the meta-data file.
831 > Force field files have a number of ``Blocks'' to delineate different
832 > types of information.  The blocks contain AtomType data, which provide
833 > properties belonging to a single AtomType, as well as interaction
834 > information which provides information about bonded or non-bonded
835 > interactions that cannot be deduced from AtomType information alone.
836 > A simple example of a forceField file is shown in scheme
837 > \ref{sch:frcExample}.
838  
839 < Force fields can be added to {\sc OpenMD}, although it comes with a few
840 < simple examples (Lennard-Jones, {\sc duff}, {\sc water}, and {\sc
841 < eam}) which are explained in the following sections.
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}]
843 > begin Options
844 >  Name = "alkane"
845 > end Options
846  
847 < \section{\label{sec:LJPot}The Lennard Jones Force Field}
847 > begin BaseAtomTypes  
848 > //name          mass  
849 > C               12.0107
850 > end BaseAtomTypes
851  
852 < The most basic force field implemented in {\sc OpenMD} is the
853 < Lennard-Jones force field, which mimics the van der Waals interaction
854 < at long distances and uses an empirical repulsion at short
855 < distances. The Lennard-Jones potential is given by:
852 > begin AtomTypes
853 > //name  base    mass
854 > CH4     C       16.05          
855 > CH3     C       15.04          
856 > CH2     C       14.03          
857 > end AtomTypes
858 >
859 > begin LennardJonesAtomTypes
860 > //name          epsilon         sigma
861 > CH4             0.2941          3.73
862 > CH3             0.1947          3.75
863 > CH2             0.09140         3.95
864 > end LennardJonesAtomTypes
865 >
866 > begin BondTypes
867 > //AT1       AT2 Type                    r0              k
868 > CH3         CH3 Harmonic                1.526           260
869 > CH3         CH2 Harmonic                1.526           260
870 > CH2         CH2 Harmonic                1.526           260
871 > end BondTypes
872 >
873 > begin BendTypes
874 > //AT1   AT2     AT3     Type            theta0   k
875 > CH3     CH2     CH3     Harmonic        114.0    124.19
876 > CH3     CH2     CH2     Harmonic        114.0    124.19
877 > CH2     CH2     CH2     Harmonic        114.0    124.19
878 > end BendTypes
879 >
880 > begin TorsionTypes
881 > //AT1 AT2  AT3  AT4  Type    
882 > CH3   CH2  CH2  CH3  Trappe  0.0  0.70544  -0.13549  1.5723
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{code}
887 >
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
892 > tags {\tt begin Options} and {\tt end Options}.  Most options don't
893 > need to be set as they come with fairly sensible default values, but
894 > the various keywords and their possible values are given in Scheme
895 > \ref{sch:optionsBlock}.
896 >
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.},
901 > label={sch:optionsBlock}]
902 > 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"
907 > EnergyMixingRule          = "geometric"    // can also be "arithmetic" or "hhg"
908 > EnergyUnitScaling         = 1.0
909 > MetallicEnergyUnitScaling = 1.0
910 > DistanceUnitScaling       = 1.0
911 > AngleUnitScaling          = 1.0
912 > TorsionAngleConvention    = "180_is_trans" // can also be "0_is_trans"
913 > vdW-12-scale              = 0.0
914 > vdW-13-scale              = 0.0
915 > vdW-14-scale              = 0.0
916 > electrostatic-12-scale    = 0.0
917 > electrostatic-13-scale    = 0.0
918 > electrostatic-14-scale    = 0.0
919 > GayBerneMu                = 2.0
920 > GayBerneNu                = 1.0
921 > EAMMixingMethod           = "Johnson"      // can also be "Daw"
922 > end Options
923 > \end{code}
924 >
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
929 > universal properties (i.e. does this atom have a fixed charge?  What
930 > is its mass?)  Dynamic properties of an atom are not intended to be
931 > properties of an atom type.  A BaseAtomType can be used to build
932 > extended sets of related atom types that all fall back to one
933 > particular type.  For example, one might want a series of atomTypes
934 > that inherit from more basic types:
935 > \begin{displaymath}
936 > \mathtt{ALA-CA} \rightarrow \mathtt{CT} \rightarrow \mathtt{CSP3} \rightarrow \mathtt{C}
937 > \end{displaymath}
938 > where for each step to the right, the atomType falls back to more
939 > primitive data.  That is, the mass could be a property of the {\tt C}
940 > type, while Lennard-Jones parameters could be properties of the {\tt
941 >  CSP3} type.  {\tt CT} could have charge information and its own set
942 > of Lennard-Jones parameter that override the CSP3 parameters.  And the
943 > {\tt ALA-CA} type might have specific torsion or charge information
944 > that override the lower level types.  A BaseAtomType contains only
945 > information a primitive name and the mass of this atom type.
946 > BaseAtomTypes can also be useful in creating files that can be easily
947 > viewed in visualization programs.  The {\tt Dump2XYZ} utility has the
948 > ability to print out the names of the base atom types for displaying
949 > simulations in Jmol or VMD.
950 >
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
955 > //Name  mass (amu)
956 > H       1.0079
957 > O       15.9994
958 > Si      28.0855
959 > Al      26.981538
960 > Mg      24.3050
961 > Ca      40.078
962 > Fe      55.845
963 > Li      6.941
964 > Na      22.98977
965 > K       39.0983
966 > Cs      132.90545
967 > Ca      40.078
968 > Ba      137.327
969 > Cl      35.453
970 > end BaseAtomTypes
971 > \end{code}
972 >
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{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}]
984 > begin AtomTypes    
985 > //Name  baseatomtype
986 > h*      H
987 > ho      H
988 > o*      O
989 > oh      O
990 > ob      O
991 > obos    O
992 > obts    O
993 > obss    O
994 > ohs     O
995 > st      Si
996 > ao      Al
997 > at      Al
998 > mgo     Mg
999 > mgh     Mg
1000 > cao     Ca
1001 > cah     Ca
1002 > feo     Fe
1003 > lio     Li
1004 > end AtomTypes
1005 > \end{code}
1006 >
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
1011 > moments of inertias in the same way that translational motion requires
1012 > mass.  For DirectionalAtoms, OpenMD treats the mass distribution with
1013 > higher priority than electrostatic distributions; the moment of
1014 > inertia tensor, $\overleftrightarrow{\mathsf I}$, should be
1015 > diagonalized to obtain body-fixed axes, and the three diagonal moments
1016 > should correspond to rotational motion \textit{around} each of these
1017 > body-fixed axes.  Charge distributions may then result in dipole
1018 > vectors that are oriented along a linear combination of the body-axes,
1019 > and in quadrupole tensors that are not necessarily diagonal in the
1020 > body frame.
1021 >
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
1026 > //Name          I_xx    I_yy    I_zz    (All moments in (amu*Ang^2)
1027 > SSD             1.7696  0.6145  1.1550  
1028 > GBC6H6          88.781  88.781  177.561
1029 > GBCH3OH         4.056   20.258  20.999
1030 > GBH2O           1.777   0.581   1.196
1031 > CO2             43.06   43.06   0.0    // single-site model for CO2
1032 > end DirectionalAtomTypes                    
1033 >
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
1038 > case, OpenMD identifies that DirectionalAtom as having only 5 degrees
1039 > of freedom (three translations and two rotations), and will alter
1040 > calculation of temperatures to reflect this.
1041 >
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
1047 > short distances. The Lennard-Jones potential is given by:
1048   \begin{equation}
1049   V_{\text{LJ}}(r_{ij}) =
1050          4\epsilon_{ij} \biggl[
# Line 862 | Line 1055 | $\sigma_{ij}$ scales the length of the interaction, an
1055   \end{equation}
1056   where $r_{ij}$ is the distance between particles $i$ and $j$,
1057   $\sigma_{ij}$ scales the length of the interaction, and
1058 < $\epsilon_{ij}$ scales the well depth of the potential. Scheme
866 < \ref{sch:LJFF} gives an example meta-data file that
867 < sets up a system of 108 Ar particles to be simulated using the
868 < Lennard-Jones force field.
1058 > $\epsilon_{ij}$ scales the well depth of the potential.
1059  
870 \begin{lstlisting}[float,caption={[Invocation of the Lennard-Jones
871 force field] A sample startup file for a small Lennard-Jones
872 simulation.},label={sch:LJFF}]
873 <OpenMD>
874  <MetaData>
875 #include "argon.md"
876
877 component{
878  type = "Ar";
879  nMol = 108;
880 }
881
882 forceField = "LJ";
883  </MetaData>
884  <Snapshot>     // not shown in this scheme
885  </Snapshot>
886 </OpenMD>
887 \end{lstlisting}
888
1060   Interactions between dissimilar particles requires the generation of
1061   cross term parameters for $\sigma$ and $\epsilon$. These parameters
1062 < are determined using the Lorentz-Berthelot mixing
1062 > are usually determined using the Lorentz-Berthelot mixing
1063   rules:\cite{Allen87}
1064   \begin{equation}
1065   \sigma_{ij} = \frac{1}{2}[\sigma_{ii} + \sigma_{jj}],
# Line 900 | Line 1071 | and
1071   \label{eq:epsilonMix}
1072   \end{equation}
1073  
1074 < \section{\label{section:DUFF}Dipolar Unified-Atom Force Field}
1074 > The values of $\sigma_{ii}$ and $\epsilon_{ii}$ are properties of atom
1075 > type $i$, and must be specified in a section of the force field file
1076 > called the {\tt LennardJonesAtomTypes} block (see listing
1077 > \ref{sch:LJatomTypesBlock}).  Separate Lennard-Jones interactions
1078 > which are not determined by the mixing rules can also be specified in
1079 > the {\tt NonbondedInteractionTypes} block (see section
1080 > \ref{section:ffNBinteraction}).
1081  
1082 < The dipolar unified-atom force field ({\sc duff}) was developed to
1083 < simulate lipid bilayers. These types of simulations require a model
1084 < capable of forming bilayers, while still being sufficiently
1085 < computationally efficient to allow large systems ($\sim$100's of
1086 < phospholipids, $\sim$1000's of waters) to be simulated for long times
1087 < ($\sim$10's of nanoseconds). With this goal in mind, {\sc duff} has no
1088 < point charges. Charge-neutral distributions are replaced with dipoles,
1089 < while most atoms and groups of atoms are reduced to Lennard-Jones
1090 < interaction sites. This simplification reduces the length scale of
1091 < long range interactions from $\frac{1}{r}$ to $\frac{1}{r^3}$,
1092 < removing the need for the computationally expensive Ewald
1093 < sum. Instead, Verlet neighbor-lists and cutoff radii are used for the
1094 < dipolar interactions, and, if desired, a reaction field may be added
1095 < to mimic longer range interactions.
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}]
1086 > begin LennardJonesAtomTypes
1087 > //Name          epsilon             sigma      
1088 > O_TIP4P         0.1550          3.15365
1089 > O_TIP4P-Ew      0.16275         3.16435
1090 > O_TIP5P         0.16            3.12  
1091 > O_TIP5P-E       0.178           3.097  
1092 > O_SPCE          0.15532         3.16549
1093 > O_SPC           0.15532         3.16549
1094 > CH4             0.279           3.73
1095 > CH3             0.185           3.75
1096 > CH2             0.0866          3.95
1097 > CH              0.0189          4.68
1098 > end LennardJonesAtomTypes
1099 > \end{code}
1100  
1101 < As an example, lipid head-groups in {\sc duff} are represented as
921 < point dipole interaction sites.  Placing a dipole at the head group's
922 < center of mass mimics the charge separation found in common
923 < phospholipid head groups such as phosphatidylcholine.\cite{Cevc87}
924 < Additionally, a large Lennard-Jones site is located at the
925 < pseudoatom's center of mass. The model is illustrated by the red atom
926 < in Fig.~\ref{fig:lipidModel}. The water model we use to
927 < complement the dipoles of the lipids is a
928 < reparameterization\cite{fennell04} of the soft sticky dipole (SSD)
929 < model of Ichiye
930 < \emph{et al.}\cite{liu96:new_model}
1101 > \subsection{\label{section:ffCharge}The ChargeAtomTypes block}
1102  
1103 < \begin{figure}
1104 < \centering
1105 < \includegraphics[width=\linewidth]{lipidModel.pdf}
1106 < \caption[A representation of a lipid model in {\sc duff}]{A
1107 < representation of the lipid model. $\phi$ is the torsion angle,
1108 < $\theta$ is the bend angle, and $\mu$ is the dipole moment of the head
1109 < group.}
1110 < \label{fig:lipidModel}
1111 < \end{figure}
1103 > In molecular simulations, proper accumulation of the electrostatic
1104 > interactions is essential and is one of the most
1105 > computationally-demanding tasks.  Most common molecular mechanics
1106 > force fields represent atomic sites with full or partial charges
1107 > protected by Lennard-Jones (short range) interactions.  Partial charge
1108 > values, $q_i$ are empirical representations of the distribution of
1109 > electronic charge in a molecule.  This means that nearly every pair
1110 > interaction involves a calculation of charge-charge forces.  Coupled
1111 > with relatively long-ranged $r^{-1}$ decay, the monopole interactions
1112 > quickly become the most expensive part of molecular simulations.  The
1113 > interactions between point charges can be handled via a number of
1114 > different algorithms, but Coulomb's law is the fundamental physical
1115 > principle governing these interactions,
1116 > \begin{equation}
1117 >  V_{\text{charge}}(r_{ij}) = \sum_{ij}\frac{q_iq_je^2}{4 \pi \epsilon_0
1118 >    r_{ij}},
1119 > \end{equation}
1120 > where $q$ represents the charge on particle $i$ or $j$, and $e$ is the
1121 > charge of an electron in Coulombs.  $\epsilon_0$ is the permittivity
1122 > of free space.
1123  
1124 < A set of scalable parameters has been used to model the alkyl groups
1125 < with Lennard-Jones sites. For this, parameters from the TraPPE force
1126 < field of Siepmann \emph{et al.}\cite{Siepmann1998} have been
1127 < utilized. TraPPE is a unified-atom representation of n-alkanes which
1128 < is parametrized against phase equilibria using Gibbs ensemble Monte
1129 < Carlo simulation techniques.\cite{Siepmann1998} One of the advantages
1130 < of TraPPE is that it generalizes the types of atoms in an alkyl chain
1131 < to keep the number of pseudoatoms to a minimum; thus, the parameters
1132 < for a unified atom such as $\text{CH}_2$ do not change depending on
1133 < what species are bonded to it.
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}]
1128 > begin ChargeAtomTypes
1129 > // Name         charge
1130 > O_TIP3P        -0.834
1131 > O_SPCE         -0.8476
1132 > H_TIP3P         0.417
1133 > H_TIP4P         0.520
1134 > H_SPCE          0.4238
1135 > EP_TIP4P       -1.040
1136 > Na+             1.0
1137 > Cl-            -1.0
1138 > end ChargeAtomTypes
1139 > \end{code}
1140  
1141 < As is required by TraPPE, {\sc duff} also constrains all bonds to be
1142 < of fixed length. Typically, bond vibrations are the fastest motions in
1143 < a molecular dynamic simulation.  With these vibrations present, small
1144 < time steps between force evaluations must be used to ensure adequate
1145 < energy conservation in the bond degrees of freedom. By constraining
958 < the bond lengths, larger time steps may be used when integrating the
959 < equations of motion. A simulation using {\sc duff} is illustrated in
960 < Scheme \ref{sch:DUFF}.
961 <
962 < \begin{lstlisting}[float,caption={[Invocation of {\sc duff}]A portion
963 < of a startup file showing a simulation utilizing {\sc
964 < duff}},label={sch:DUFF}]  
965 < <OpenMD>
966 <  <MetaData>
967 < #include "water.md"
968 < #include "lipid.md"
969 <
970 < component{
971 <  type = "simpleLipid_16";
972 <  nMol = 60;
973 < }
974 <
975 < component{
976 <  type = "SSD_water";
977 <  nMol = 1936;
978 < }
979 <
980 < forceField = "DUFF";
981 <  </MetaData>
982 <  <Snapshot>     // not shown in this scheme
983 <  </Snapshot>
984 < </OpenMD>
985 < \end{lstlisting}
986 <
987 < \subsection{\label{section:energyFunctions}{\sc duff} Energy Functions}
988 <
989 < The total potential energy function in {\sc duff} is
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
1145 > multipole moments,
1146   \begin{equation}
1147 < V = \sum^{N}_{I=1} V^{I}_{\text{Internal}}
992 <        + \sum^{N-1}_{I=1} \sum_{J>I} V^{IJ}_{\text{Cross}},
993 < \label{eq:totalPotential}
1147 > U_{\bf{ab}}(r)=\hat{M}_{\bf a} \hat{M}_{\bf b} \frac{1}{r}  \label{kernel}.
1148   \end{equation}
1149 < where $V^{I}_{\text{Internal}}$ is the internal potential of molecule $I$:
1149 > where the multipole operator on site $\bf a$,
1150   \begin{equation}
1151 < V^{I}_{\text{Internal}} =
1152 <        \sum_{\theta_{ijk} \in I} V_{\text{bend}}(\theta_{ijk})
1153 <        + \sum_{\phi_{ijkl} \in I} V_{\text{torsion}}(\phi_{ijkl})
1000 <        + \sum_{i \in I} \sum_{(j>i+4) \in I}
1001 <        \biggl[ V_{\text{LJ}}(r_{ij}) +  V_{\text{dipole}}
1002 <        (\mathbf{r}_{ij},\boldsymbol{\Omega}_{i},\boldsymbol{\Omega}_{j})
1003 <        \biggr].
1004 < \label{eq:internalPotential}
1151 > \hat{M}_{\bf a} = C_{\bf a} - D_{{\bf a}\alpha} \frac{\partial}{\partial r_{\alpha}}
1152 > +  Q_{{\bf a}\alpha\beta}
1153 > \frac{\partial^2}{\partial r_{\alpha} \partial r_{\beta}} + \dots
1154   \end{equation}
1155 < Here $V_{\text{bend}}$ is the bend potential for all 1, 3 bonded pairs
1156 < within the molecule $I$, and $V_{\text{torsion}}$ is the torsion
1157 < potential for all 1, 4 bonded pairs.  The pairwise portions of the
1158 < non-bonded interactions are excluded for atom pairs that are involved
1159 < in the smae bond, bend, or torsion. All other atom pairs within a
1160 < molecule are subject to the LJ pair potential.
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 {\it primitive}
1158 > multipoles.  If the site is representing a distribution of charges,
1159 > these can be expressed as,
1160 > \begin{align}
1161 > C_{\bf a} =&\sum_{k \, \text{in \bf a}} q_k , \label{eq:charge} \\
1162 > D_{{\bf a}\alpha} =&\sum_{k \, \text{in \bf a}} q_k r_{k\alpha}, \label{eq:dipole}\\
1163 > Q_{{\bf a}\alpha\beta} =& \frac{1}{2} \sum_{k \, \text{in \bf a}} q_k
1164 > r_{k\alpha}  r_{k\beta} . \label{eq:quadrupole}
1165 > \end{align}
1166 > Note that the definition of the primitive quadrupole here differs from
1167 > the standard traceless form, and contains an additional Taylor-series
1168 > based factor of $1/2$.  
1169  
1170 < The bend potential of a molecule is represented by the following function:
1170 > The details of the multipolar interactions will be given later, but
1171 > many readers are familiar with the dipole-dipole potential:
1172   \begin{equation}
1173 < V_{\text{bend}}(\theta_{ijk}) = k_{\theta}( \theta_{ijk} - \theta_0
1173 > V_{\text{dipole}}(\mathbf{r}_{ij},\boldsymbol{\Omega}_{i},
1174 >        \boldsymbol{\Omega}_{j}) = \frac{|{\bf D}_i||{\bf D}_j|}{4\pi\epsilon_{0}r_{ij}^{3}} \biggl[
1175 >        \boldsymbol{\hat{u}}_{i} \cdot \boldsymbol{\hat{u}}_{j}
1176 >        -
1177 >        3(\boldsymbol{\hat{u}}_i \cdot \hat{\mathbf{r}}_{ij}) %
1178 >                (\boldsymbol{\hat{u}}_j \cdot \hat{\mathbf{r}}_{ij}) \biggr].
1179 > \label{eq:dipolePot}
1180 > \end{equation}
1181 > Here $\mathbf{r}_{ij}$ is the vector starting at atom $i$ pointing
1182 > towards $j$, and $\boldsymbol{\Omega}_i$ and $\boldsymbol{\Omega}_j$
1183 > are the orientational degrees of freedom for atoms $i$ and $j$
1184 > respectively. The magnitude of the dipole moment of atom $i$ is $|{\bf
1185 >  D}_i|$, $\boldsymbol{\hat{u}}_i$ is the standard unit orientation
1186 > vector of $\boldsymbol{\Omega}_i$, and $\boldsymbol{\hat{r}}_{ij}$ is
1187 > the unit vector pointing along $\mathbf{r}_{ij}$
1188 > ($\boldsymbol{\hat{r}}_{ij}=\mathbf{r}_{ij}/|\mathbf{r}_{ij}|$).
1189 >
1190 >
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 $10^{-26} \mathrm{~esu~cm}^2$)},
1195 > label={sch:MultipoleAtomTypesBlock}]
1196 > begin MultipoleAtomTypes
1197 > // Euler angles are given in zxz convention in units of degrees.
1198 > //
1199 > // point dipoles:
1200 > // name d phi theta psi dipole_moment
1201 > DIP     d 0.0 0.0   0.0     1.91   // dipole points along z-body axis
1202 > //
1203 > // point quadrupoles:
1204 > // name q phi theta psi Qxx Qyy Qzz
1205 > CO2     q 0.0 0.0   0.0 0.0 0.0 -0.430592  //quadrupole tensor has zz element
1206 > //
1207 > // Atoms with both dipole and quadrupole moments:
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{code}
1212 >
1213 > Specifying a MultipoleAtomType requires declaring how the
1214 > electrostatic frame for the site is rotated relative to the body-fixed
1215 > axes for that atom. The Euler angles $(\phi, \theta, \psi)$ for this
1216 > rotation must be given, and then the dipole, quadrupole, or all of
1217 > these moments are specified in the electrostatic frame.  In OpenMD,
1218 > the Euler angles are specified in the $zxz$ convention and are entered
1219 > in units of degrees.  Dipole moments are entered in units of Debye,
1220 > and Quadrupole moments in units of Debye \AA.
1221 >
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 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
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
1237 > r}_{ij}}) = 4\epsilon ({{\bf \hat u}_i}, {{\bf \hat u}_j},
1238 > {{\bf \hat r}_{ij}})\left[\left(\frac{\sigma_0}{r_{ij}-\sigma({{\bf \hat u
1239 > }_i},
1240 > {{\bf \hat u}_j}, {{\bf \hat r}_{ij}})+\sigma_0}\right)^{12}
1241 > -\left(\frac{\sigma_0}{r_{ij}-\sigma({{\bf \hat u}_i}, {{\bf \hat u}_j},
1242 > {{\bf \hat r}_{ij}})+\sigma_0}\right)^6\right]
1243 > \label{eq:gb}
1244 > \end{equation*}
1245 >
1246 > The range $(\sigma({\bf \hat{u}}_{i},{\bf \hat{u}}_{j},{\bf
1247 > \hat{r}}_{ij}))$, and strength $(\epsilon({\bf \hat{u}}_{i},{\bf
1248 > \hat{u}}_{j},{\bf \hat{r}}_{ij}))$ parameters
1249 > are dependent on the relative orientations of the two ellipsoids (${\bf
1250 > \hat{u}}_{i},{\bf \hat{u}}_{j}$) as well as the direction of the
1251 > inter-ellipsoid separation (${\bf \hat{r}}_{ij}$).  The shape and
1252 > attractiveness of each ellipsoid is governed by a relatively small set
1253 > of parameters:
1254 > \begin{itemize}
1255 > \item  $d$:  range parameter for the side-by-side (S) and cross (X) configurations
1256 > \item  $l$:  range parameter for the end-to-end (E) configuration
1257 > \item  $\epsilon_X$:  well-depth parameter for the cross (X) configuration
1258 > \item  $\epsilon_S$:  well-depth parameter for the side-by-side (S) configuration
1259 > \item  $\epsilon_E$:  well depth parameter for the end-to-end (E) configuration
1260 > \item  $dw$: The ``softness'' of the potential
1261 > \end{itemize}
1262 > Additionally, there are two universal paramters to govern the overall
1263 > importance of the purely orientational ($\nu$) and the mixed
1264 > orientational / translational ($\mu$) parts of strength of the
1265 > interactions.  These parameters have default or ``canonical'' values,
1266 > but may be changed as a force field option:
1267 > \begin{itemize}
1268 >  \item $\nu$: purely orientational part : defaults to 1
1269 >  \item $\mu$: mixed orientational / translational part : defaults to
1270 >    2
1271 > \end{itemize}
1272 > Further details of the potential are given
1273 > elsewhere,\cite{Luckhurst:1990fy,Golubkov06,SunX._jp0762020} and an
1274 > excellent overview of the computational methods that can be used to
1275 > efficiently compute forces and torques for this potential can be found
1276 > in Ref. \citealp{Golubkov06}
1277 >
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.},
1282 > label={sch:GayBerneAtomTypes}]
1283 > begin GayBerneAtomTypes
1284 > //Name          d       l       eps_X           eps_S           eps_E     dw
1285 > GBlinear        2.8104  9.993   0.774729        0.774729        0.116839  1.0
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{code}
1290 >
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
1295 > original SSD was developed by Ichiye \emph{et
1296 >  al.}\cite{liu96:new_model} as a modified form of the hard-sphere
1297 > water model proposed by Bratko, Blum, and
1298 > Luzar.\cite{Bratko85,Bratko95} It consists of a single point dipole
1299 > with a Lennard-Jones core and a sticky potential that directs the
1300 > particles to assume the proper hydrogen bond orientation in the first
1301 > solvation shell. Thus, the interaction between two SSD water molecules
1302 > \emph{i} and \emph{j} is given by the potential
1303 > \begin{equation}
1304 > V_{ij} =
1305 >        V_{ij}^{LJ} (r_{ij})\ + V_{ij}^{dp}
1306 >        (\mathbf{r}_{ij},\boldsymbol{\Omega}_i,\boldsymbol{\Omega}_j)\ +
1307 >        V_{ij}^{sp}
1308 >        (\mathbf{r}_{ij},\boldsymbol{\Omega}_i,\boldsymbol{\Omega}_j),
1309 > \label{eq:ssdPot}
1310 > \end{equation}
1311 > where the $\mathbf{r}_{ij}$ is the position vector between molecules
1312 > \emph{i} and \emph{j} with magnitude equal to the distance $r_{ij}$, and
1313 > $\boldsymbol{\Omega}_i$ and $\boldsymbol{\Omega}_j$ represent the
1314 > orientations of the respective molecules. The Lennard-Jones and dipole
1315 > parts of the potential are given by equations \ref{eq:lennardJonesPot}
1316 > and \ref{eq:dipolePot} respectively. The sticky part is described by
1317 > the following,
1318 > \begin{equation}
1319 > u_{ij}^{sp}(\mathbf{r}_{ij},\boldsymbol{\Omega}_i,\boldsymbol{\Omega}_j)=
1320 >        \frac{\nu_0}{2}[s(r_{ij})w(\mathbf{r}_{ij},
1321 >        \boldsymbol{\Omega}_i,\boldsymbol{\Omega}_j) +
1322 >        s^\prime(r_{ij})w^\prime(\mathbf{r}_{ij},
1323 >        \boldsymbol{\Omega}_i,\boldsymbol{\Omega}_j)]\ ,
1324 > \label{eq:stickyPot}
1325 > \end{equation}
1326 > where $\nu_0$ is a strength parameter for the sticky potential, and
1327 > $s$ and $s^\prime$ are cubic switching functions which turn off the
1328 > sticky interaction beyond the first solvation shell. The $w$ function
1329 > can be thought of as an attractive potential with tetrahedral
1330 > geometry:
1331 > \begin{equation}
1332 > w({\bf r}_{ij},{\bf \Omega}_i,{\bf \Omega}_j)=
1333 >        \sin\theta_{ij}\sin2\theta_{ij}\cos2\phi_{ij},
1334 > \label{eq:stickyW}
1335 > \end{equation}
1336 > while the $w^\prime$ function counters the normal aligned and
1337 > anti-aligned structures favored by point dipoles:
1338 > \begin{equation}
1339 > w^\prime({\bf r}_{ij},{\bf \Omega}_i,{\bf \Omega}_j)=
1340 >        (\cos\theta_{ij}-0.6)^2(\cos\theta_{ij}+0.8)^2-w^0,
1341 > \label{eq:stickyWprime}
1342 > \end{equation}
1343 > It should be noted that $w$ is proportional to the sum of the $Y_3^2$
1344 > and $Y_3^{-2}$ spherical harmonics (a linear combination which
1345 > enhances the tetrahedral geometry for hydrogen bonded structures),
1346 > while $w^\prime$ is a purely empirical function.  A more detailed
1347 > description of the functional parts and variables in this potential
1348 > can be found in the original SSD
1349 > articles.\cite{liu96:new_model,liu96:monte_carlo,chandra99:ssd_md,Ichiye03}
1350 >
1351 > \begin{figure}
1352 > \centering
1353 > \includegraphics[width=\linewidth]{waterAngle.pdf}
1354 > \caption[Coordinate definition for the SSD/E water model]{Coordinates
1355 > for the interaction between two SSD/E water molecules.  $\theta_{ij}$
1356 > is the angle that $r_{ij}$ makes with the $\hat{z}$ vector in the
1357 > body-fixed frame for molecule $i$.  The $\hat{z}$ vector bisects the
1358 > HOH angle in each water molecule. }
1359 > \label{fig:ssd}
1360 > \end{figure}
1361 >
1362 > Since SSD/E is a single-point {\it dipolar} model, the force
1363 > calculations are simplified significantly relative to the standard
1364 > {\it charged} multi-point models. In the original Monte Carlo
1365 > simulations using this model, Ichiye {\it et al.} reported that using
1366 > SSD decreased computer time by a factor of 6-7 compared to other
1367 > models.\cite{liu96:new_model} What is most impressive is that these
1368 > savings did not come at the expense of accurate depiction of the
1369 > liquid state properties.  Indeed, SSD/E maintains reasonable agreement
1370 > with the Head-Gordon diffraction data for the structural features of
1371 > liquid water.\cite{hura00,liu96:new_model} Additionally, the dynamical
1372 > properties exhibited by SSD/E agree with experiment better than those
1373 > of more computationally expensive models (like TIP3P and
1374 > SPC/E).\cite{chandra99:ssd_md} The combination of speed and accurate
1375 > depiction of solvent properties makes SSD/E a very attractive model
1376 > for the simulation of large scale biochemical simulations.
1377 >
1378 > Recent constant pressure simulations revealed issues in the original
1379 > SSD model that led to lower than expected densities at all target
1380 > pressures,\cite{Ichiye03,fennell04} so variants on the sticky
1381 > potential can be specified by using one of a number of substitute atom
1382 > types (see listing \ref{sch:StickyAtomTypes}).  A table of the
1383 > parameter values and the drawbacks and benefits of the different
1384 > density corrected SSD models can be found in
1385 > reference~\citealp{fennell04}.
1386 >
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.},
1391 > label={sch:StickyAtomTypes}]
1392 > begin StickyAtomTypes
1393 > //name  w0      v0 (kcal/mol)   v0p     rl (Ang)  ru    rlp     rup
1394 > SSD_E   0.07715 3.90            3.90    2.40      3.80  2.75    3.35
1395 > SSD_RF  0.07715 3.90            3.90    2.40      3.80  2.75    3.35
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{code}
1400 >
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
1405 > interaction which models ``Embedding'' a positively charged
1406 > pseudo-atom core in the electron density due to the free valance
1407 > ``sea'' of electrons created by the surrounding atoms in the system.
1408 > A pairwise part of the potential (which is primarily repulsive)
1409 > describes the interaction of the positively charged metal core ions
1410 > with one another.  These potentials have the form:
1411 > \begin{equation}
1412 > V  =  \sum_{i} F_{i}\left[\rho_{i}\right] + \sum_{i} \sum_{j \neq i}
1413 > \phi_{ij}({\bf r}_{ij})
1414 > \end{equation}
1415 > where $F_{i} $ is an embedding functional that approximates the energy
1416 > required to embed a positively-charged core ion $i$ into a linear
1417 > superposition of spherically averaged atomic electron densities given
1418 > by $\rho_{i}$,
1419 > \begin{equation}
1420 > \rho_{i}   =  \sum_{j \neq i} f_{j}({\bf r}_{ij}),
1421 > \end{equation}
1422 > Since the density at site $i$ ($\rho_i$) must be computed before the
1423 > embedding functional can be evaluated, {\sc eam} and the related
1424 > transition metal potentials require two loops through the atom pairs
1425 > to compute the inter-atomic forces.
1426 >
1427 > The pairwise portion of the potential, $\phi_{ij}$, is usually a
1428 > repulsive interaction between atoms $i$ and $j$.
1429 >
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}
1434 > It has been widely adopted in the materials science community and a
1435 > good review of {\sc eam} and other formulations of metallic potentials
1436 > was given by Voter.\cite{Voter:95}
1437 >
1438 > In the original formulation of {\sc eam}\cite{Daw84}, the pair
1439 > potential, $\phi_{ij}$ was an entirely repulsive term; however later
1440 > refinements to {\sc eam} allowed for more general forms for
1441 > $\phi$.\cite{Daw89} The effective cutoff distance, $r_{{\text cut}}$
1442 > is the distance at which the values of $f(r)$ and $\phi(r)$ drop to
1443 > zero for all atoms present in the simulation.  In practice, this
1444 > distance is fairly small, limiting the summations in the {\sc eam}
1445 > equation to the few dozen atoms surrounding atom $i$ for both the
1446 > density $\rho$ and pairwise $\phi$ interactions.
1447 >
1448 > In computing forces for alloys, OpenMD uses mixing rules outlined by
1449 > Johnson~\cite{johnson89} to compute the heterogenous pair potential,
1450 > \begin{equation}
1451 > \label{eq:johnson}
1452 > \phi_{ab}(r)=\frac{1}{2}\left(
1453 > \frac{f_{b}(r)}{f_{a}(r)}\phi_{aa}(r)+
1454 > \frac{f_{a}(r)}{f_{b}(r)}\phi_{bb}(r)
1455 > \right).
1456 > \end{equation}
1457 > No mixing rule is needed for the densities, since the density at site
1458 > $i$ is simply the linear sum of density contributions of all the other
1459 > atoms.
1460 >
1461 > The {\sc eam} force field illustrates an additional feature of {\sc
1462 > OpenMD}.  Foiles, Baskes and Daw fit {\sc eam} potentials for Cu, Ag,
1463 > Au, Ni, Pd, Pt and alloys of these metals.\cite{FBD86} These fits are
1464 > included in {\sc OpenMD} as the {\tt u3} variant of the {\sc eam} force
1465 > field.  Voter and Chen reparamaterized a set of {\sc eam} functions
1466 > which do a better job of predicting melting points.\cite{Voter:87}
1467 > These functions are included in {\sc OpenMD} as the {\tt VC} variant of
1468 > the {\sc eam} force field.  An additional set of functions (the
1469 > ``Universal 6'' functions) are included in {\sc OpenMD} as the {\tt u6}
1470 > variant of {\sc eam}.  For example, to specify the Voter-Chen variant
1471 > of the {\sc eam} force field, the user would add the {\tt
1472 > forceFieldVariant = "VC";} line to the meta-data file.
1473 >
1474 > The potential files used by the {\sc eam} force field are in the
1475 > standard {\tt funcfl} format, which is the format utilized by a number
1476 > of other codes (e.g. ParaDyn~\cite{Paradyn}, {\sc dynamo 86}).  It
1477 > should be noted that the energy units in these files are in eV, not
1478 > $\mbox{kcal mol}^{-1}$ as in the rest of the {\sc OpenMD} force field
1479 > files.  
1480 >
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.},
1485 > label={sch:EAMAtomTypes}]
1486 > begin EAMAtomTypes
1487 > Au      Au.u3.funcfl
1488 > Ag      Ag.u3.funcfl
1489 > Cu      Cu.u3.funcfl
1490 > Ni      Ni.u3.funcfl
1491 > Pd      Pd.u3.funcfl
1492 > Pt      Pt.u3.funcfl
1493 > end EAMAtomTypes
1494 > \end{code}
1495 >
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 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
1505 > i}\epsilon_{ij}V^{pair}_{ij}(r_{ij})-c_{i}\epsilon_{ii}\sqrt{\rho_{i}}\right] ,
1506 > \end{equation}
1507 > where $V^{pair}_{ij}$ and $\rho_{i}$ are given by
1508 > \begin{equation}
1509 > \label{eq:SCP2}
1510 > V^{pair}_{ij}(r)=\left(
1511 > \frac{\alpha_{ij}}{r_{ij}}\right)^{n_{ij}} \hspace{1in} \rho_{i}=\sum_{j\neq i}\left(
1512 > \frac{\alpha_{ij}}{r_{ij}}\right) ^{m_{ij}}
1513 > \end{equation}
1514 >
1515 > $V^{pair}_{ij}$ is a repulsive pairwise potential that accounts for
1516 > interactions of the pseudo-atom cores.  The $\sqrt{\rho_i}$ term in
1517 > Eq. (\ref{eq:SCP1}) is an attractive many-body potential that models
1518 > the interactions between the valence electrons and the cores of the
1519 > pseudo-atoms.  $\epsilon_{ij}$, $\epsilon_{ii}$, $c_i$ and
1520 > $\alpha_{ij}$ are parameters used to tune the potential for different
1521 > transition metals.
1522 >
1523 > The {\sc sc} potential form has also been parameterized by Qi {\it et
1524 > al.}\cite{Qi99} These parameters were obtained via empirical and {\it
1525 > ab initio} calculations to match structural features of the FCC
1526 > crystal.  Interested readers are encouraged to consult reference
1527 > \citealp{Qi99} for further details.
1528 >
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
1533 > using the keyword {\tt MetallicEnergyUnitScaling}.  Without this {\tt
1534 > Options} keyword, the default units for $\epsilon$ are kcal/mol.  The
1535 > other parameters, $m$, $n$, and $c$ are unitless.},
1536 > label={sch:SCAtomTypes}]
1537 > begin SCAtomTypes
1538 > // Name  epsilon(eV)      c      m       n      alpha(angstroms)
1539 > Ni      0.0073767       84.745  5.0     10.0    3.5157
1540 > Cu      0.0057921       84.843  5.0     10.0    3.6030
1541 > Rh      0.0024612       305.499 5.0     13.0    3.7984
1542 > Pd      0.0032864       148.205 6.0     12.0    3.8813
1543 > Ag      0.0039450       96.524  6.0     11.0    4.0691
1544 > Ir      0.0037674       224.815 6.0     13.0    3.8344  
1545 > Pt      0.0097894       71.336  7.0     11.0    3.9163
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{code}
1550 >
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 > 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 > 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 $\theta_{ijk}$ is the angle defined by atoms $i$, $j$, and $k$
1725 < (see Fig.~\ref{fig:lipidModel}), $\theta_0$ is the equilibrium
1726 < bond angle, and $k_{\theta}$ is the force constant which determines the
1727 < strength of the harmonic bend. The parameters for $k_{\theta}$ and
1728 < $\theta_0$ are borrowed from those in TraPPE.\cite{Siepmann1998}
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 the angle $\phi_{ijkl}$ is defined
1830 > \begin{equation}
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 $l$.  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 < The torsion potential and parameters are also borrowed from TraPPE. It is
1856 < of the form:
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) = c_1[1 + \cos \phi]
1860 <        + c_2[1 + \cos(2\phi)]
1029 <        + c_3[1 + \cos(3\phi)],
1030 < \label{eq:origTorsionPot}
1031 < \end{equation}
1032 < where:
1033 < \begin{equation}
1034 < \cos\phi = (\hat{\mathbf{r}}_{ij} \times \hat{\mathbf{r}}_{jk}) \cdot
1035 <        (\hat{\mathbf{r}}_{jk} \times \hat{\mathbf{r}}_{kl}).
1036 < \label{eq:torsPhi}
1037 < \end{equation}
1038 < Here, $\hat{\mathbf{r}}_{\alpha\beta}$ are the set of unit bond
1039 < vectors between atoms $i$, $j$, $k$, and $l$. For computational
1040 < efficiency, the torsion potential has been recast after the method of
1041 < {\sc charmm},\cite{Brooks83} in which the angle series is converted to
1042 < a power series of the form:
1043 < \begin{equation}
1044 < V_{\text{torsion}}(\phi) =  
1045 <        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 + 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 < The cross potential between molecules $I$ and $J$,
1913 < $V^{IJ}_{\text{Cross}}$, is as follows:
1914 < \begin{equation}
1915 < V^{IJ}_{\text{Cross}} =
1916 <        \sum_{i \in I} \sum_{j \in J}
1917 <        \biggl[ V_{\text{LJ}}(r_{ij}) +  V_{\text{dipole}}
1918 <        (\mathbf{r}_{ij},\boldsymbol{\Omega}_{i},\boldsymbol{\Omega}_{j})
1919 <        + V_{\text{sticky}}
1920 <        (\mathbf{r}_{ij},\boldsymbol{\Omega}_{i},\boldsymbol{\Omega}_{j})
1069 <        \biggr],
1070 < \label{eq:crossPotentail}
1071 < \end{equation}
1072 < where $V_{\text{LJ}}$ is the Lennard Jones potential,
1073 < $V_{\text{dipole}}$ is the dipole dipole potential, and
1074 < $V_{\text{sticky}}$ is the sticky potential defined by the SSD model
1075 < (Sec.~\ref{section:SSD}). Note that not all atom types include all
1076 < interactions.
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 < The dipole-dipole potential has the following form:
1923 < \begin{equation}
1924 < V_{\text{dipole}}(\mathbf{r}_{ij},\boldsymbol{\Omega}_{i},
1925 <        \boldsymbol{\Omega}_{j}) = \frac{|\mu_i||\mu_j|}{4\pi\epsilon_{0}r_{ij}^{3}} \biggl[
1926 <        \boldsymbol{\hat{u}}_{i} \cdot \boldsymbol{\hat{u}}_{j}
1927 <        -
1928 <        3(\boldsymbol{\hat{u}}_i \cdot \hat{\mathbf{r}}_{ij}) %
1929 <                (\boldsymbol{\hat{u}}_j \cdot \hat{\mathbf{r}}_{ij}) \biggr].
1930 < \label{eq:dipolePot}
1931 < \end{equation}
1932 < Here $\mathbf{r}_{ij}$ is the vector starting at atom $i$ pointing
1933 < towards $j$, and $\boldsymbol{\Omega}_i$ and $\boldsymbol{\Omega}_j$
1934 < are the orientational degrees of freedom for atoms $i$ and $j$
1935 < respectively. The magnitude of the dipole moment of atom $i$ is
1936 < $|\mu_i|$, $\boldsymbol{\hat{u}}_i$ is the standard unit orientation
1937 < vector of $\boldsymbol{\Omega}_i$, and $\boldsymbol{\hat{r}}_{ij}$ is
1938 < the unit vector pointing along $\mathbf{r}_{ij}$
1939 < ($\boldsymbol{\hat{r}}_{ij}=\mathbf{r}_{ij}/|\mathbf{r}_{ij}|$).
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 < \subsection{\label{section:SSD}The {\sc duff} Water Models: SSD/E
1949 < and SSD/RF}
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 < In the interest of computational efficiency, the default solvent used
1953 < by {\sc OpenMD} is the extended Soft Sticky Dipole (SSD/E) water
1954 < model.\cite{fennell04} The original SSD was developed by Ichiye
1955 < \emph{et al.}\cite{liu96:new_model} as a modified form of the hard-sphere
1956 < water model proposed by Bratko, Blum, and
1957 < Luzar.\cite{Bratko85,Bratko95} It consists of a single point dipole
1958 < with a Lennard-Jones core and a sticky potential that directs the
1959 < particles to assume the proper hydrogen bond orientation in the first
1960 < solvation shell. Thus, the interaction between two SSD water molecules
1961 < \emph{i} and \emph{j} is given by the potential
1952 > \subsection{\label{section:ffInversion}The InversionTypes block}
1953 >
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 < V_{ij} =
1973 <        V_{ij}^{LJ} (r_{ij})\ + V_{ij}^{dp}
1974 <        (\mathbf{r}_{ij},\boldsymbol{\Omega}_i,\boldsymbol{\Omega}_j)\ +
1114 <        V_{ij}^{sp}
1115 <        (\mathbf{r}_{ij},\boldsymbol{\Omega}_i,\boldsymbol{\Omega}_j),
1116 < \label{eq:ssdPot}
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 < where the $\mathbf{r}_{ij}$ is the position vector between molecules
1977 < \emph{i} and \emph{j} with magnitude equal to the distance $r_{ij}$, and
1978 < $\boldsymbol{\Omega}_i$ and $\boldsymbol{\Omega}_j$ represent the
1979 < orientations of the respective molecules. The Lennard-Jones and dipole
1980 < parts of the potential are given by equations \ref{eq:lennardJonesPot}
1123 < and \ref{eq:dipolePot} respectively. The sticky part is described by
1124 < the following,
1125 < \begin{equation}
1126 < u_{ij}^{sp}(\mathbf{r}_{ij},\boldsymbol{\Omega}_i,\boldsymbol{\Omega}_j)=
1127 <        \frac{\nu_0}{2}[s(r_{ij})w(\mathbf{r}_{ij},
1128 <        \boldsymbol{\Omega}_i,\boldsymbol{\Omega}_j) +
1129 <        s^\prime(r_{ij})w^\prime(\mathbf{r}_{ij},
1130 <        \boldsymbol{\Omega}_i,\boldsymbol{\Omega}_j)]\ ,
1131 < \label{eq:stickyPot}
1132 < \end{equation}
1133 < where $\nu_0$ is a strength parameter for the sticky potential, and
1134 < $s$ and $s^\prime$ are cubic switching functions which turn off the
1135 < sticky interaction beyond the first solvation shell. The $w$ function
1136 < can be thought of as an attractive potential with tetrahedral
1137 < geometry:
1138 < \begin{equation}
1139 < w({\bf r}_{ij},{\bf \Omega}_i,{\bf \Omega}_j)=
1140 <        \sin\theta_{ij}\sin2\theta_{ij}\cos2\phi_{ij},
1141 < \label{eq:stickyW}
1142 < \end{equation}
1143 < while the $w^\prime$ function counters the normal aligned and
1144 < anti-aligned structures favored by point dipoles:
1145 < \begin{equation}
1146 < w^\prime({\bf r}_{ij},{\bf \Omega}_i,{\bf \Omega}_j)=
1147 <        (\cos\theta_{ij}-0.6)^2(\cos\theta_{ij}+0.8)^2-w^0,
1148 < \label{eq:stickyWprime}
1149 < \end{equation}
1150 < It should be noted that $w$ is proportional to the sum of the $Y_3^2$
1151 < and $Y_3^{-2}$ spherical harmonics (a linear combination which
1152 < enhances the tetrahedral geometry for hydrogen bonded structures),
1153 < while $w^\prime$ is a purely empirical function.  A more detailed
1154 < description of the functional parts and variables in this potential
1155 < can be found in the original SSD
1156 < articles.\cite{liu96:new_model,liu96:monte_carlo,chandra99:ssd_md,Ichiye03}
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 < \begin{figure}
1983 < \centering
1984 < \includegraphics[width=\linewidth]{waterAngle.pdf}
1985 < \caption[Coordinate definition for the SSD/E water model]{Coordinates
1986 < for the interaction between two SSD/E water molecules.  $\theta_{ij}$
1987 < is the angle that $r_{ij}$ makes with the $\hat{z}$ vector in the
1988 < body-fixed frame for molecule $i$.  The $\hat{z}$ vector bisects the
1989 < HOH angle in each water molecule. }
1990 < \label{fig:ssd}
1991 < \end{figure}
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 + \section{\label{section::ffLongRange}Long Range Interactions}
2020  
2021 < Since SSD/E is a single-point {\it dipolar} model, the force
2022 < calculations are simplified significantly relative to the standard
2023 < {\it charged} multi-point models. In the original Monte Carlo
2024 < simulations using this model, Ichiye {\it et al.} reported that using
1174 < SSD decreased computer time by a factor of 6-7 compared to other
1175 < models.\cite{liu96:new_model} What is most impressive is that these
1176 < savings did not come at the expense of accurate depiction of the
1177 < liquid state properties.  Indeed, SSD/E maintains reasonable agreement
1178 < with the Head-Gordon diffraction data for the structural features of
1179 < liquid water.\cite{hura00,liu96:new_model} Additionally, the dynamical
1180 < properties exhibited by SSD/E agree with experiment better than those
1181 < of more computationally expensive models (like TIP3P and
1182 < SPC/E).\cite{chandra99:ssd_md} The combination of speed and accurate
1183 < depiction of solvent properties makes SSD/E a very attractive model
1184 < for the simulation of large scale biochemical simulations.
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 < Recent constant pressure simulations revealed issues in the original
2027 < SSD model that led to lower than expected densities at all target
1188 < pressures.\cite{Ichiye03,fennell04} The default model in {\sc OpenMD}
1189 < is therefore SSD/E, a density corrected derivative of SSD that
1190 < exhibits improved liquid structure and transport behavior. If the use
1191 < of a reaction field long-range interaction correction is desired, it
1192 < is recommended that the parameters be modified to those of the SSD/RF
1193 < model (an SSD variant parameterized for reaction field). These solvent
1194 < parameters are listed and can be easily modified in the {\sc duff}
1195 < force field file ({\tt DUFF.frc}).  A table of the parameter values
1196 < and the drawbacks and benefits of the different density corrected SSD
1197 < models can be found in reference~\cite{fennell04}.
2026 > \subsection{\label{section:ffNBinteraction}The NonBondedInteractions
2027 >  block}
2028  
2029 < \section{\label{section:WATER}The {\sc water} 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 < In addition to the {\sc duff} force field's solvent description, a
2036 < separate {\sc water} force field has been included for simulating most
2037 < of the common rigid-body water models. This force field includes the
2038 < simple and point-dipolar models (SSD, SSD1, SSD/E, SSD/RF, and DPD
2039 < water), as well as the common charge-based models (SPC, SPC/E, TIP3P,
2040 < TIP4P, and
2041 < TIP5P).\cite{liu96:new_model,Ichiye03,fennell04,Marrink01,Berendsen81,Berendsen87,Jorgensen83,Mahoney00}
2042 < In order to handle these models, charge-charge interactions were
2043 < included in the force-loop:
2044 < \begin{equation}
2045 < V_{\text{charge}}(r_{ij}) = \sum_{ij}\frac{q_iq_je^2}{r_{ij}},
2046 < \end{equation}
2047 < where $q$ represents the charge on particle $i$ or $j$, and $e$ is the
2048 < charge of an electron in Coulombs. The charge-charge interaction
2049 < support is rudimentary in the current version of {\sc OpenMD}.  As with
2050 < the other pair interactions, charges can be simulated with a pure
2051 < cutoff or a reaction field.  The various methods for performing the
2052 < Ewald summation have not yet been included.  The {\sc water} force
2053 < field can be easily expanded through modification of the {\sc water}
2054 < force field file ({\tt WATER.frc}). By adding atom types and inserting
2055 < the appropriate parameters, it is possible to extend the force field
2056 < to handle rigid molecules other than water.
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 < \section{\label{section:eam}Embedded Atom Method}
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 < {\sc OpenMD} implements a potential that describes bonding in
2069 < transition metal
2070 < systems.~\cite{Finnis84,Ercolessi88,Chen90,Qi99,Ercolessi02} This
2071 < potential has an attractive interaction which models ``Embedding'' a
2072 < positively charged pseudo-atom core in the electron density due to the
2073 < free valance ``sea'' of electrons created by the surrounding atoms in
1232 < the system.  A pairwise part of the potential (which is primarily
1233 < repulsive) describes the interaction of the positively charged metal
1234 < core ions with one another.  The Embedded Atom Method ({\sc
1235 < eam})~\cite{Daw84,FBD86,johnson89,Lu97} has been widely adopted in the
1236 < materials science community and has been included in {\sc OpenMD}. A
1237 < good review of {\sc eam} and other formulations of metallic potentials
1238 < was given by Voter.\cite{Voter:95}
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 < The {\sc eam} potential has the form:
2076 < \begin{equation}
2077 < V  =  \sum_{i} F_{i}\left[\rho_{i}\right] + \sum_{i} \sum_{j \neq i}
1243 < \phi_{ij}({\bf r}_{ij})
1244 < \end{equation}
1245 < where $F_{i} $ is an embedding functional that approximates the energy
1246 < required to embed a positively-charged core ion $i$ into a linear
1247 < superposition of spherically averaged atomic electron densities given
1248 < by $\rho_{i}$,
1249 < \begin{equation}
1250 < \rho_{i}   =  \sum_{j \neq i} f_{j}({\bf r}_{ij}),
1251 < \end{equation}
1252 < Since the density at site $i$ ($\rho_i$) must be computed before the
1253 < embedding functional can be evaluated, {\sc eam} and the related
1254 < transition metal potentials require two loops through the atom pairs
1255 < to compute the inter-atomic forces.
2075 > //Shifted Morse
2076 > //Atom1 Atom2   ShiftedMorse    r0     D0       beta0
2077 > Au      O_SPCE  ShiftedMorse    3.70   0.0424   0.769
2078  
2079 < The pairwise portion of the potential, $\phi_{ij}$, is a primarily
2080 < repulsive interaction between atoms $i$ and $j$. In the original
2081 < formulation of {\sc eam}\cite{Daw84}, $\phi_{ij}$ was an entirely
1260 < repulsive term; however later refinements to {\sc eam} allowed for
1261 < more general forms for $\phi$.\cite{Daw89} The effective cutoff
1262 < distance, $r_{{\text cut}}$ is the distance at which the values of
1263 < $f(r)$ and $\phi(r)$ drop to zero for all atoms present in the
1264 < simulation.  In practice, this distance is fairly small, limiting the
1265 < summations in the {\sc eam} equation to the few dozen atoms
1266 < surrounding atom $i$ for both the density $\rho$ and pairwise $\phi$
1267 < interactions.
2079 > //Repulsive Morse
2080 > //Atom1 Atom2   RepulsiveMorse  r0     D0       beta0
2081 > Au      H_SPCE  RepulsiveMorse  -1.00  0.00850  0.769
2082  
2083 < In computing forces for alloys, mixing rules as outlined by
2084 < Johnson~\cite{johnson89} are used to compute the heterogenous pair
2085 < potential,
2086 < \begin{equation}
2087 < \label{eq:johnson}
2088 < \phi_{ab}(r)=\frac{1}{2}\left(
1275 < \frac{f_{b}(r)}{f_{a}(r)}\phi_{aa}(r)+
1276 < \frac{f_{a}(r)}{f_{b}(r)}\phi_{bb}(r)
1277 < \right).
1278 < \end{equation}
1279 < No mixing rule is needed for the densities, since the density at site
1280 < $i$ is simply the linear sum of density contributions of all the other
1281 < atoms.
1282 <
1283 < The {\sc eam} force field illustrates an additional feature of {\sc
1284 < OpenMD}.  Foiles, Baskes and Daw fit {\sc eam} potentials for Cu, Ag,
1285 < Au, Ni, Pd, Pt and alloys of these metals.\cite{FBD86} These fits are
1286 < included in {\sc OpenMD} as the {\tt u3} variant of the {\sc eam} force
1287 < field.  Voter and Chen reparamaterized a set of {\sc eam} functions
1288 < which do a better job of predicting melting points.\cite{Voter:87}
1289 < These functions are included in {\sc OpenMD} as the {\tt VC} variant of
1290 < the {\sc eam} force field.  An additional set of functions (the
1291 < ``Universal 6'' functions) are included in {\sc OpenMD} as the {\tt u6}
1292 < variant of {\sc eam}.  For example, to specify the Voter-Chen variant
1293 < of the {\sc eam} force field, the user would add the {\tt
1294 < forceFieldVariant = "VC";} line to the meta-data file.
1295 <
1296 < The potential files used by the {\sc eam} force field are in the
1297 < standard {\tt funcfl} format, which is the format utilized by a number
1298 < of other codes (e.g. ParaDyn~\cite{Paradyn}, {\sc dynamo 86}).  It
1299 < should be noted that the energy units in these files are in eV, not
1300 < $\mbox{kcal mol}^{-1}$ as in the rest of the {\sc OpenMD} force field
1301 < files.  
1302 <
1303 < \section{\label{section:sc}The Sutton-Chen Force Field}
1304 <
1305 < The Sutton-Chen ({\sc sc})~\cite{Chen90} potential has been used to
1306 < study a wide range of phenomena in metals.  Although it is similar in
1307 < form to the {\sc eam} potential, the Sutton-Chen model takes on a
1308 < simpler form,
1309 < \begin{equation}
1310 < \label{eq:SCP1}
1311 < U_{tot}=\sum _{i}\left[ \frac{1}{2}\sum _{j\neq
1312 < i}D_{ij}V^{pair}_{ij}(r_{ij})-c_{i}D_{ii}\sqrt{\rho_{i}}\right] ,
1313 < \end{equation}
1314 < where $V^{pair}_{ij}$ and $\rho_{i}$ are given by
1315 < \begin{equation}
1316 < \label{eq:SCP2}
1317 < V^{pair}_{ij}(r)=\left(
1318 < \frac{\alpha_{ij}}{r_{ij}}\right)^{n_{ij}}, \rho_{i}=\sum_{j\neq i}\left(
1319 < \frac{\alpha_{ij}}{r_{ij}}\right) ^{m_{ij}}
1320 < \end{equation}
1321 <
1322 < $V^{pair}_{ij}$ is a repulsive pairwise potential that accounts for
1323 < interactions of the pseudo-atom cores.  The $\sqrt{\rho_i}$ term in
1324 < Eq. (\ref{eq:SCP1}) is an attractive many-body potential that models
1325 < the interactions between the valence electrons and the cores of the
1326 < pseudo-atoms.  $D_{ij}$, $D_{ii}$, $c_i$ and $\alpha_{ij}$ are
1327 < parameters used to tune the potential for different transition
1328 < metals.
1329 <
1330 < The {\sc sc} potential form has also been parameterized by Qi {\it et
1331 < al.}\cite{Qi99} These parameters were obtained via empirical and {\it
1332 < ab initio} calculations to match structural features of the FCC
1333 < crystal.  To specify the original Sutton-Chen variant of the {\sc sc}
1334 < force field, the user would add the {\tt forceFieldVariant = "SC";}
1335 < line to the meta-data file, while specification of the Qi {\it et al.}
1336 < quantum-adapted variant of the {\sc sc} potential, the user would add
1337 < the {\tt forceFieldVariant = "QSC";} line to the meta-data file.
1338 <
1339 < \section{\label{section:clay}The CLAY force field}
1340 <
1341 < The {\sc clay} force field is based on an ionic (nonbonded)
1342 < description of the metal-oxygen interactions associated with hydrated
1343 < phases. All atoms are represented as point charges and are allowed
1344 < complete translational freedom. Metal-oxygen interactions are based on
1345 < a simple Lennard-Jones potential combined with electrostatics. The
1346 < empirical parameters were optimized by Cygan {\it et
1347 < al.}\cite{Cygan04} on the basis of known mineral structures, and
1348 < partial atomic charges were derived from periodic DFT quantum chemical
1349 < calculations of simple oxide, hydroxide, and oxyhydroxide model
1350 < compounds with well-defined structures.
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  
1352
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 1479 | 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}-\
1483 < 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 1524 | Line 2260 | this reason, the default electrostatic summation metho
2260   this reason, the default electrostatic summation method utilized by
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}}
# Line 1906 | Line 2713 | LD & Langevin Dynamics & {\tt ensemble = LD;} \\
2713   &  (with separate barostats on each box dimension) & \\
2714   LD & Langevin Dynamics & {\tt ensemble = LD;} \\
2715   &  (approximates the effects of an implicit solvent) & \\
2716 < LangevinHull & Non-periodic Langevin Dynamics  & {\tt ensemble = LHull;} \\
2716 > LangevinHull & Non-periodic Langevin Dynamics  & {\tt ensemble = LangevinHull;} \\
2717   & (Langevin Dynamics for molecules on convex hull;\\
2718   & Newtonian for interior molecules) & \\
2719   \end{tabular}
# Line 2630 | Line 3437 | tensor.
3437   \label{table:ldParameters}
3438   \end{longtable}
3439  
3440 < \section{Langevin Hull Dynamics (LHull)}
3440 > \section{Constant Pressure without periodic boundary conditions (The LangevinHull)}
3441  
3442 < The Langevin Hull uses an external bath at a fixed constant pressure
3442 > The Langevin Hull\cite{Vardeman2011} uses an external bath at a fixed constant pressure
3443   ($P$) and temperature ($T$) with an effective solvent viscosity
3444   ($\eta$).  This bath interacts only with the objects on the exterior
3445   hull of the system.  Defining the hull of the atoms in a simulation is
# Line 2706 | Line 3513 | fluctuations of the random force, $\mathbf{R}(t)$, by
3513   depends on the geometry and surface area of facet $f$ and the
3514   viscosity of the bath.  The resistance tensor is related to the
3515   fluctuations of the random force, $\mathbf{R}(t)$, by the
3516 < fluctuation-dissipation theorem,
2710 < \begin{eqnarray}
2711 < \left< {\mathbf R}_f(t) \right> & = & 0 \\
2712 < \left<{\mathbf R}_f(t) {\mathbf R}_f^T(t^\prime)\right> & = & 2 k_B T\
2713 < \Xi_f(t)\delta(t-t^\prime).
2714 < \label{eq:randomForce}
2715 < \end{eqnarray}
3516 > fluctuation-dissipation theorem (see Eq. \ref{eq:randomForce}).
3517  
3518   Once the resistance tensor is known for a given facet, a stochastic
3519   vector that has the properties in Eq. (\ref{eq:randomForce}) can be
3520   calculated efficiently by carrying out a Cholesky decomposition to
3521 < obtain the square root matrix of the resistance tensor,
3522 < \begin{equation}
2722 < \Xi_f = {\bf S} {\bf S}^{T},
2723 < \label{eq:Cholesky}
2724 < \end{equation}
2725 < where ${\bf S}$ is a lower triangular matrix.\cite{Schlick2002} A
2726 < vector with the statistics required for the random force can then be
2727 < obtained by multiplying ${\bf S}$ onto a random 3-vector ${\bf Z}$ which
2728 < has elements chosen from a Gaussian distribution, such that:
2729 < \begin{equation}
2730 < \langle {\bf Z}_i \rangle = 0, \hspace{1in} \langle {\bf Z}_i \cdot
2731 < {\bf Z}_j \rangle = \frac{2 k_B T}{\delta t} \delta_{ij},
2732 < \end{equation}
2733 < where $\delta t$ is the timestep in use during the simulation. The
2734 < random force, ${\bf R}_{f} = {\bf S} {\bf Z}$, can be shown to
2735 < have the correct properties required by Eq. (\ref{eq:randomForce}).
3521 > obtain the square root matrix of the resistance tensor (see
3522 > Eq. \ref{eq:Cholesky}).
3523  
3524 < Our treatment of the resistance tensor is approximate.  $\Xi_f$ for a
3525 < rigid triangular plate would normally be treated as a $6 \times 6$
3526 < tensor that includes translational and rotational drag as well as
3527 < translational-rotational coupling. The computation of resistance
3528 < tensors for rigid bodies has been detailed
3524 > Our treatment of the resistance tensor for the Langevin Hull facets is
3525 > approximate.  $\Xi_f$ for a rigid triangular plate would normally be
3526 > treated as a $6 \times 6$ tensor that includes translational and
3527 > rotational drag as well as translational-rotational coupling. The
3528 > computation of resistance tensors for rigid bodies has been detailed
3529   elsewhere,\cite{JoseGarciadelaTorre02012000,Garcia-de-la-Torre:2001wd,GarciadelaTorreJ2002,Sun:2008fk}
3530   but the standard approach involving bead approximations would be
3531   prohibitively expensive if it were recomputed at each step in a
# Line 2790 | Line 3577 | The Delaunay triangulation and computation of the conv
3577   \item Atomic positions and velocities are propagated.
3578   \end{enumerate}
3579   The Delaunay triangulation and computation of the convex hull are done
3580 < using calls to the qhull library.\cite{Qhull} There is a minimal
3581 < penalty for computing the convex hull and resistance tensors at each
3582 < step in the molecular dynamics simulation (roughly 0.02 $\times$ cost
3583 < of a single force evaluation), and the convex hull is remarkably easy
3584 < to parallelize on distributed memory machines.
3585 <
3580 > using calls to the qhull library,\cite{Qhull} and for this reason, if
3581 > qhull is not detected during the build, the Langevin Hull integrator
3582 > will not be available.  There is a minimal penalty for computing the
3583 > convex hull and resistance tensors at each step in the molecular
3584 > dynamics simulation (roughly 0.02 $\times$ cost of a single force
3585 > evaluation).
3586  
3587   \begin{longtable}[c]{GBF}
3588   \caption{Meta-data Keywords: Required parameters for the Langevin Hull integrator}
# Line 2810 | Line 3597 | This parameter must be specified to use Langevin Hull
3597   This parameter must be specified to use Langevin Hull dynamics. \\
3598   {\tt targetPressure} & atm & Sets the target pressure of the system.
3599   This parameter must be specified to use Langevin Hull dynamics. \\
3600 < {\tt usePeriodicBoundaryConditions = false} & logical & Turns off periodic boundary conditions.
3600 > {\tt usePeriodicBoundaryConditions} & logical & Turns off periodic boundary conditions.
3601   This parameter must be set to \tt false \\
3602   \label{table:lhullParameters}
3603   \end{longtable}
# Line 2964 | Line 3751 | Harmonic Forces are used by default
3751   \label{table:zconParams}
3752   \end{longtable}
3753  
3754 < \chapter{\label{section:restraints}Restraints}
3755 < Restraints are external potentials that are added to a system to keep
3756 < particular molecules or collections of particles close to some
3757 < reference structure.  A restraint can be a collective
3754 > % \chapter{\label{section:restraints}Restraints}
3755 > % Restraints are external potentials that are added to a system to
3756 > % keep particular molecules or collections of particles close to some
3757 > % reference structure.  A restraint can be a collective
3758  
3759   \chapter{\label{section:thermInt}Thermodynamic Integration}
3760  
# Line 3106 | Line 3893 | Einstein crystal
3893   Einstein crystal
3894   \label{table:thermIntParams}
3895   \end{longtable}
3896 +
3897 + \chapter{\label{section:rnemd}Reverse Non-Equilibrium Molecular Dynamics (RNEMD)}
3898 +
3899 + There are many ways to compute transport properties from molecular
3900 + dynamics simulations.  Equilibrium Molecular Dynamics (EMD)
3901 + simulations can be used by computing relevant time correlation
3902 + functions and assuming linear response theory holds.  For some transport properties (notably thermal conductivity), EMD approaches
3903 + are subject to noise and poor convergence of the relevant
3904 + correlation functions. Traditional Non-equilibrium Molecular Dynamics
3905 + (NEMD) methods impose a gradient (e.g. thermal or momentum) on a
3906 + simulation.  However, the resulting flux is often difficult to
3907 + measure. Furthermore, problems arise for NEMD simulations of
3908 + heterogeneous systems, such as phase-phase boundaries or interfaces,
3909 + where the type of gradient to enforce at the boundary between
3910 + materials is unclear.
3911 +
3912 + {\it Reverse} Non-Equilibrium Molecular Dynamics (RNEMD) methods adopt
3913 + a different approach in that an unphysical {\it flux} is imposed
3914 + between different regions or ``slabs'' of the simulation box.  The
3915 + response of the system is to develop a temperature or momentum {\it
3916 +  gradient} between the two regions. Since the amount of the applied
3917 + flux is known exactly, and the measurement of gradient is generally
3918 + less complicated, imposed-flux methods typically take shorter
3919 + simulation times to obtain converged results for transport properties.
3920  
3921 + \begin{figure}
3922 + \includegraphics[width=\linewidth]{rnemdDemo}
3923 + \caption{The (VSS) RNEMD approach imposes unphysical transfer of both
3924 +  linear momentum and kinetic energy between a ``hot'' slab and a
3925 +  ``cold'' slab in the simulation box.  The system responds to this
3926 +  imposed flux by generating both momentum and temperature gradients.
3927 +  The slope of the gradients can then be used to compute transport
3928 +  properties (e.g. shear viscosity and thermal conductivity).}
3929 + \label{rnemdDemo}
3930 + \end{figure}
3931  
3932 + \section{\label{section:algo}Three algorithms for carrying out RNEMD simulations}
3933 + \subsection{\label{subsection:swapping}The swapping algorithm}
3934 + The original ``swapping'' approaches by M\"{u}ller-Plathe {\it et
3935 +  al.}\cite{ISI:000080382700030,MullerPlathe:1997xw} can be understood
3936 + as a sequence of imaginary elastic collisions between particles in
3937 + opposite slabs.  In each collision, the entire momentum vectors of
3938 + both particles may be exchanged to generate a thermal
3939 + flux. Alternatively, a single component of the momentum vectors may be
3940 + exchanged to generate a shear flux.  This algorithm turns out to be
3941 + quite useful in many simulations. However, the M\"{u}ller-Plathe
3942 + swapping approach perturbs the system away from ideal
3943 + Maxwell-Boltzmann distributions, and this may leads to undesirable
3944 + side-effects when the applied flux becomes large.\cite{Maginn:2010}
3945 + This limits the applicability of the swapping algorithm, so in OpenMD,
3946 + we have implemented two additional algorithms for RNEMD in addition to the
3947 + original swapping approach.
3948 +
3949 + \subsection{\label{subsection:nivs}Non-Isotropic Velocity Scaling (NIVS)}
3950 + Instead of having momentum exchange between {\it individual particles}
3951 + in each slab, the NIVS algorithm applies velocity scaling to all of
3952 + the selected particles in both slabs.\cite{kuang:164101} A combination of linear
3953 + momentum, kinetic energy, and flux constraint equations governs the
3954 + amount of velocity scaling performed at each step. Interested readers
3955 + should consult ref. \citealp{kuang:164101} for further details on the
3956 + methodology.
3957 +
3958 + NIVS has been shown to be very effective at producing thermal
3959 + gradients and for computing thermal conductivities, particularly for
3960 + heterogeneous interfaces.  Although the NIVS algorithm can also be
3961 + applied to impose a directional momentum flux, thermal anisotropy was
3962 + observed in relatively high flux simulations, and the method is not
3963 + suitable for imposing a shear flux or for computing shear viscosities.
3964 +
3965 + \subsection{\label{subsection:vss}Velocity Shearing and Scaling (VSS)}
3966 + The third RNEMD algorithm implemented in OpenMD utilizes a series of
3967 + simultaneous velocity shearing and scaling exchanges between the two
3968 + slabs.\cite{2012MolPh.110..691K}  This method results in a set of simpler equations to satisfy
3969 + the conservation constraints while creating a desired flux between the
3970 + two slabs.
3971 +
3972 + The VSS approach is versatile in that it may be used to implement both
3973 + thermal and shear transport either separately or simultaneously.
3974 + Perturbations of velocities away from the ideal Maxwell-Boltzmann
3975 + distributions are minimal, and thermal anisotropy is kept to a
3976 + minimum.  This ability to generate simultaneous thermal and shear
3977 + fluxes has been utilized to map out the shear viscosity of SPC/E water
3978 + over a wide range of temperatures (90~K) just with a single simulation.
3979 + VSS-RNEMD also allows the directional momentum flux to have
3980 + arbitary directions, which could aid in the study of anisotropic solid
3981 + surfaces in contact with liquid environments.
3982 +
3983 + \section{\label{section:usingRNEMD}Using OpenMD to perform a RNEMD simulation}
3984 + \subsection{\label{section:rnemdParams} What the user needs to specify}
3985 + To carry out a RNEMD simulation,
3986 + a user must specify a number of parameters in the MetaData (.md) file.
3987 + Because the RNEMD methods have a large number of parameters, these
3988 + must be enclosed in a {\it separate} {\tt RNEMD\{...\}} block.  The most important
3989 + parameters to specify are the {\tt useRNEMD}, {\tt fluxType} and flux
3990 + parameters. Most other parameters (summarized in table
3991 + \ref{table:rnemd}) have reasonable default values.  {\tt fluxType}
3992 + sets up the kind of exchange that will be carried out between the two
3993 + slabs (either Kinetic Energy ({\tt KE}) or momentum ({\tt Px, Py, Pz,
3994 +  Pvector}), or some combination of these).  The flux is specified
3995 + with the use of three possible parameters: {\tt kineticFlux} for
3996 + kinetic energy exchange, as well as {\tt momentumFlux} or {\tt
3997 +  momentumFluxVector} for simulations with directional exchange.
3998 +
3999 + \subsection{\label{section:rnemdResults} Processing the results}
4000 + OpenMD will generate a {\tt .rnemd}
4001 + file with the same prefix as the original {\tt .md} file.  This file
4002 + contains a running average of properties of interest computed within a
4003 + set of bins that divide the simulation cell along the $z$-axis.  The
4004 + first column of the {\tt .rnemd} file is the $z$ coordinate of the
4005 + center of each bin, while following columns may contain the average
4006 + temperature, velocity, or density within each bin.  The output format
4007 + in the {\tt .rnemd} file can be altered with the {\tt outputFields},
4008 + {\tt outputBins}, and {\tt outputFileName} parameters.  A report at
4009 + the top of the {\tt .rnemd} file contains the current exchange totals
4010 + as well as the average flux applied during the simulation.  Using the
4011 + slope of the temperature or velocity gradient obtaine from the {\tt
4012 +  .rnemd} file along with the applied flux, the user can very simply
4013 + arrive at estimates of thermal conductivities ($\lambda$),
4014 + \begin{equation}
4015 + J_z = -\lambda \frac{\partial T}{\partial z},
4016 + \end{equation}
4017 + and shear viscosities ($\eta$),
4018 + \begin{equation}
4019 + j_z(p_x) = -\eta \frac{\partial \langle v_x \rangle}{\partial z}.
4020 + \end{equation}
4021 + Here, the quantities on the left hand side are the actual flux values
4022 + (in the header of the {\tt .rnemd} file), while the slopes are
4023 + obtained from linear fits to the gradients observed in the {\tt
4024 +  .rnemd} file.
4025 +
4026 + More complicated simulations (including interfaces) require a bit more
4027 + care.  Here the second derivative may be required to compute the
4028 + interfacial thermal conductance,
4029 + \begin{align}
4030 +  G^\prime &= \left(\nabla\lambda \cdot \mathbf{\hat{n}}\right)_{z_0} \\
4031 +  &= \frac{\partial}{\partial z}\left(-\frac{J_z}{
4032 +      \left(\frac{\partial T}{\partial z}\right)}\right)_{z_0} \\
4033 +  &= J_z\left(\frac{\partial^2 T}{\partial z^2}\right)_{z_0} \Big/
4034 +  \left(\frac{\partial T}{\partial z}\right)_{z_0}^2.
4035 +  \label{derivativeG}
4036 + \end{align}
4037 + where $z_0$ is the location of the interface between two materials and
4038 + $\mathbf{\hat{n}}$ is a unit vector normal to the interface.  We
4039 + suggest that users interested in interfacial conductance consult
4040 + reference \citealp{kuang:AuThl} for other approaches to computing $G$.
4041 + Users interested in {\it friction coefficients} at heterogeneous
4042 + interfaces may also find reference \citealp{2012MolPh.110..691K}
4043 + useful.
4044 +
4045 + \newpage
4046 +
4047 + \begin{longtable}[c]{JKLM}
4048 + \caption{Meta-data Keywords: Parameters for RNEMD simulations}\\
4049 + \multicolumn{4}{c}{The following keywords must be enclosed inside a {\tt RNEMD\{...\}} block.}
4050 + \\ \hline
4051 + {\bf keyword} & {\bf units} & {\bf use} & {\bf remarks}  \\ \hline
4052 + \endhead
4053 + \hline
4054 + \endfoot
4055 + {\tt useRNEMD} & logical & perform RNEMD? & default is ``false'' \\
4056 + {\tt objectSelection} & string & see section \ref{section:syntax}
4057 + for selection syntax & default is ``select all'' \\
4058 + {\tt method} & string & exchange method & one of the following:
4059 + {\tt Swap, NIVS,} or {\tt VSS}  (default is {\tt VSS}) \\
4060 + {\tt fluxType} & string & what is being exchanged between slabs? & one
4061 + of the following: {\tt KE, Px, Py, Pz, Pvector, KE+Px, KE+Py, KE+Pvector} \\
4062 + {\tt kineticFlux} & kcal mol$^{-1}$ \AA$^{-2}$ fs$^{-1}$ & specify the kinetic energy flux &  \\
4063 + {\tt momentumFlux} & amu \AA$^{-1}$ fs$^{-2}$ & specify the momentum flux & \\
4064 + {\tt momentumFluxVector} & amu \AA$^{-1}$ fs$^{-2}$ & specify the momentum flux when
4065 + {\tt Pvector} is part of the exchange & Vector3d input\\
4066 + {\tt exchangeTime} & fs & how often to perform the exchange & default is 100 fs\\
4067 +
4068 + {\tt slabWidth} & $\mbox{\AA}$ & width of the two exchange slabs & default is $\mathsf{H}_{zz} / 10.0$ \\
4069 + {\tt slabAcenter} & $\mbox{\AA}$ & center of the end slab & default is 0 \\
4070 + {\tt slabBcenter} & $\mbox{\AA}$ & center of the middle slab & default is $\mathsf{H}_{zz} / 2$ \\
4071 + {\tt outputFileName} & string & file name for output histograms & default is the same prefix as the
4072 + .md file, but with the {\tt .rnemd} extension \\
4073 + {\tt outputBins} & int & number of $z$-bins in the output histogram &
4074 + default is 20 \\
4075 + {\tt outputFields} & string & columns to print in the {\tt .rnemd}
4076 + file where each column is separated by a pipe ($\mid$) symbol. & Allowed column names are: {\sc z, temperature, velocity, density} \\
4077 + \label{table:rnemd}
4078 + \end{longtable}
4079 +
4080   \chapter{\label{section:minimizer}Energy Minimization}
4081  
4082 < As one of the basic procedures of molecular modeling, energy
3114 < minimization is used to identify local configurations that are stable
4082 > Energy minimization is used to identify local configurations that are stable
4083   points on the potential energy surface. There is a vast literature on
4084   energy minimization algorithms have been developed to search for the
4085   global energy minimum as well as to find local structures which are
# Line 3238 | Line 4206 | diagram of the class heirarchy:
4206   \begin{figure}
4207   \centering
4208   \includegraphics[width=3in]{heirarchy.pdf}
4209 < \caption[Class heirarchy for StuntDoubles in {\sc OpenMD}-4]{ \\ The
4210 < class heirarchy of StuntDoubles in {\sc OpenMD}-4. The selection
4209 > \caption[Class heirarchy for StuntDoubles in {\sc OpenMD}]{ \\ The
4210 > class heirarchy of StuntDoubles in {\sc OpenMD}. The selection
4211   syntax allows the user to select any of the objects that are descended
4212   from a StuntDouble.}
4213   \label{fig:heirarchy}
# Line 3419 | Line 4387 | VMD. The options available for Dump2XYZ are as follows
4387    -z & {\tt -{}-zconstraint}  &                replace the atom types of zconstraint molecules  (default=off) \\
4388    -r & {\tt -{}-rigidbody}  &                  add a pseudo COM atom to rigidbody  (default=off) \\
4389    -t & {\tt -{}-watertype} &                   replace the atom type of water model (default=on) \\
4390 <  -b & {\tt -{}-basetype}  &                   using base atom type  (default=off) \\
4390 >  -b & {\tt -{}-basetype}  &                   using base atom type
4391 >  (default=off) \\
4392 >  -v& {\tt -{}-velocities}             & Print velocities in xyz file  (default=off)\\
4393 >  -f& {\tt -{}-forces}                 & Print forces xyz file  (default=off)\\
4394 >  -u& {\tt -{}-vectors}                & Print vectors (dipoles, etc) in xyz file  
4395 >                                  (default=off)\\
4396 >  -c& {\tt -{}-charges}                & Print charges in xyz file  (default=off)\\
4397 >  -e& {\tt -{}-efield}                 & Print electric field vector in xyz file  
4398 >                                  (default=off)\\
4399       & {\tt -{}-repeatX=INT}  &                 The number of images to repeat in the x direction  (default=`0') \\
4400       & {\tt -{}-repeatY=INT} &                 The number of images to repeat in the y direction  (default=`0') \\
4401       &  {\tt -{}-repeatZ=INT}  &                The number of images to repeat in the z direction  (default=`0') \\
# Line 3511 | Line 4487 | The options available for {\tt StaticProps} are as fol
4487      & {\tt -{}-sele1=selection script}   & select the first StuntDouble set \\
4488      & {\tt -{}-sele2=selection script}   & select the second StuntDouble set \\
4489      & {\tt -{}-sele3=selection script}   & select the third StuntDouble set \\
4490 <    & {\tt -{}-refsele=selection script} & select reference (can only be used with {\tt -{}-gxyz}) \\
4490 >    & {\tt -{}-refsele=selection script} & select reference (can only
4491 >    be used with {\tt -{}-gxyz}) \\
4492 >    & {\tt -{}-comsele=selection script}
4493 >                               & select stunt doubles for center-of-mass
4494 >                                  reference point\\
4495 >    & {\tt -{}-seleoffset=INT}        & global index offset for a second object (used
4496 >                                  to define a vector between sites in molecule)\\
4497 >
4498      & {\tt -{}-molname=STRING}           & molecule name \\
4499      & {\tt -{}-begin=INT}                & begin internal index \\
4500      & {\tt -{}-end=INT}                  & end internal index \\
4501 +    & {\tt -{}-radius=DOUBLE}            & nanoparticle radius\\
4502   \hline
4503   \multicolumn{3}{|l|}{One option from the following group of options is required:} \\
4504   \hline
4505 <    &  {\tt -{}-gofr}                    &  $g(r)$ \\
4506 <    &  {\tt -{}-r\_theta}                 &  $g(r, \cos(\theta))$ \\
4507 <    &  {\tt -{}-r\_omega}                 &  $g(r, \cos(\omega))$ \\
4508 <    &  {\tt -{}-theta\_omega}             &  $g(\cos(\theta), \cos(\omega))$ \\
4505 >    & {\tt -{}-bo}          & bond order parameter ({\tt -{}-rcut} must be specified) \\
4506 >    & {\tt -{}-bor}         & bond order parameter as a function of
4507 >    radius  ({\tt -{}-rcut} must be specified) \\
4508 >    & {\tt -{}-bad}         & $N(\theta)$ bond angle density within ({\tt -{}-rcut} must be specified) \\
4509 >    & {\tt -{}-count}       & count of molecules matching selection
4510 >    criteria (and associated statistics) \\
4511 >  -g&  {\tt -{}-gofr}                    &  $g(r)$ \\
4512 >    &  {\tt -{}-gofz}                    &  $g(z)$ \\
4513 >    &  {\tt -{}-r\_theta}                &  $g(r, \cos(\theta))$ \\
4514 >    &  {\tt -{}-r\_omega}                &  $g(r, \cos(\omega))$ \\
4515 >    &  {\tt -{}-r\_z}                    &  $g(r, z)$ \\
4516 >    &  {\tt -{}-theta\_omega}            &  $g(\cos(\theta), \cos(\omega))$ \\
4517      &  {\tt -{}-gxyz}                    &  $g(x, y, z)$ \\
4518 <    &  {\tt -{}-p2}                      &  $P_2$ order parameter ({\tt -{}-sele1} and {\tt -{}-sele2} must be specified) \\
4518 >    &  {\tt -{}-twodgofr}                & 2D $g(r)$ (Slab width {\tt -{}-dz} must be specified)\\
4519 >  -p&  {\tt -{}-p2}                      &  $P_2$ order parameter  ({\tt -{}-sele1} must be specified, {\tt -{}-sele2} is optional) \\
4520 >    &  {\tt -{}-rp2}                     &  Ripple order parameter ({\tt -{}-sele1} and {\tt -{}-sele2} must be specified) \\
4521      &  {\tt -{}-scd}                     &  $S_{CD}$ order parameter(either {\tt -{}-sele1}, {\tt -{}-sele2}, {\tt -{}-sele3} are specified or {\tt -{}-molname}, {\tt -{}-begin}, {\tt -{}-end} are specified) \\
4522 <    &  {\tt -{}-density}                 &  density plot ({\tt -{}-sele1} must be specified) \\
4523 <    &  {\tt -{}-slab\_density}           &  slab density ({\tt -{}-sele1} must be specified)
4522 >  -d&  {\tt -{}-density}                 &  density plot \\
4523 >    &  {\tt -{}-slab\_density}           &  slab density \\
4524 >    &  {\tt -{}-p\_angle}                & $p(\cos(\theta))$ ($\theta$
4525 >    is the angle between molecular axis and radial vector from origin\\
4526 >    &  {\tt -{}-hxy}                     & Calculates the undulation  spectrum, $h(x,y)$, of an interface \\
4527 >    &  {\tt -{}-rho\_r}                  & $\rho(r)$\\
4528 >    &  {\tt -{}-angle\_r}                &  $\theta(r)$ (spatially resolves the
4529 >    angle between the molecular axis and the radial vector from the
4530 >    origin)\\
4531 >    &  {\tt -{}-hullvol}                 & hull volume of nanoparticle\\
4532 >    &  {\tt -{}-rodlength}               & length of nanorod\\
4533 >  -Q&  {\tt -{}-tet\_param}              & tetrahedrality order parameter ($Q$)\\
4534 >    &  {\tt -{}-tet\_param\_z}           & spatially-resolved tetrahedrality order
4535 >                                   parameter $Q(z)$\\
4536 >    &  {\tt -{}-rnemdz}                  & slab-resolved RNEMD statistics (temperature,
4537 >                                  density, velocity)\\
4538 >    &  {\tt -{}-rnemdr}                  & shell-resolved RNEMD statistics (temperature,
4539 >                                  density, angular velocity)
4540   \end{longtable}
4541  
4542   \subsection{\label{section:DynamicProps}DynamicProps}
# Line 3567 | Line 4577 | The options available for DynamicProps are as follows:
4577    -o& {\tt -{}-output=filename}        & output file name \\
4578      & {\tt -{}-sele1=selection script} & select first StuntDouble set \\
4579      & {\tt -{}-sele2=selection script} & select second StuntDouble set (if sele2 is not set, use script from sele1) \\
4580 +    & {\tt -{}-order=INT}              & Lengendre Polynomial Order\\
4581 +  -z& {\tt -{}-nzbins=INT}             & Number of $z$ bins (default=`100`)\\
4582 +  -m& {\tt -{}-memory=memory specification}
4583 +                                &Available memory  
4584 +                                  (default=`2G`)\\
4585   \hline
4586   \multicolumn{3}{|l|}{One option from the following group of options is required:} \\
4587   \hline
4588 <  -r& {\tt -{}-rcorr}                  & compute mean square displacement \\
4589 <  -v& {\tt -{}-vcorr}                  & compute velocity correlation function \\
4590 <  -d& {\tt -{}-dcorr}                  & compute dipole correlation function
4588 >  -s& {\tt -{}-selecorr}               & selection correlation function \\
4589 >  -r& {\tt -{}-rcorr}                  & compute mean squared displacement \\
4590 >  -v& {\tt -{}-vcorr}                  & velocity autocorrelation function \\
4591 >  -d& {\tt -{}-dcorr}                  & dipole correlation function \\
4592 >  -l& {\tt -{}-lcorr}                  & Lengendre correlation function \\
4593 >    & {\tt -{}-lcorrZ}                 & Lengendre correlation function binned by $z$ \\
4594 >    & {\tt -{}-cohZ}                   & Lengendre correlation function for OH bond vectors binned by $z$\\
4595 >  -M& {\tt -{}-sdcorr}                 & System dipole correlation function\\
4596 >    & {\tt -{}-r\_rcorr}               & Radial mean squared displacement\\
4597 >    & {\tt -{}-thetacorr}              & Angular mean squared displacement\\
4598 >    & {\tt -{}-drcorr}                 & Directional mean squared displacement for particles with unit vectors\\
4599 >    & {\tt -{}-helfandEcorr}           & Helfand moment for thermal conductvity\\
4600 >  -p& {\tt -{}-momentum}               & Helfand momentum for viscosity\\
4601 >    & {\tt -{}-stresscorr}             & Stress tensor correlation function
4602   \end{longtable}
4603  
4604   \chapter{\label{section:PreparingInput} Preparing Input Configurations}
# Line 3639 | Line 4665 | expect the the input specifier on the command line.
4665   to {\tt atom2md}, but they use a specific input format and do not
4666   expect the the input specifier on the command line.
4667  
4668 +
4669   \section{\label{section:SimpleBuilder}SimpleBuilder}
4670  
4671   {\tt SimpleBuilder} creates simple lattice structures.  It requires an
# Line 3665 | Line 4692 | The options available for SimpleBuilder are as follows
4692      &  {\tt -{}-nz=INT}            &  number of unit cells in z
4693   \end{longtable}
4694  
4695 + \section{\label{section:icosahedralBuilder}icosahedralBuilder}
4696 +
4697 + {\tt icosahedralBuilder} creates single-component geometric solids
4698 + that can be useful in simulating nanostructures.  Like the other
4699 + builders, it requires an initial, but skeletal {\sc OpenMD} file to
4700 + specify the component that is to be placed on the lattice.  The total
4701 + number of placed molecules will be shown at the top of the
4702 + configuration file that is generated, and that number may not match
4703 + the original meta-data file, so a new meta-data file is also generated
4704 + which matches the lattice structure.
4705 +
4706 + The options available for icosahedralBuilder are as follows:
4707 + \begin{longtable}[c]{|EFG|}
4708 + \caption{icosahedralBuilder Command-line Options}
4709 + \\ \hline
4710 + {\bf option} & {\bf verbose option} & {\bf behavior} \\ \hline
4711 + \endhead
4712 + \hline
4713 + \endfoot
4714 +  -h& {\tt -{}-help}               & Print help and exit\\
4715 +  -V& {\tt -{}-version}            & Print version and exit\\
4716 +  -o& {\tt -{}-output=STRING}      & Output file name\\
4717 +  -n& {\tt -{}-shells=INT}         & Nanoparticle shells\\
4718 +  -d& {\tt -{}-latticeConstant=DOUBLE} & Lattice spacing in Angstroms for cubic lattice.\\
4719 +  -c& {\tt -{}-columnAtoms=INT}        & Number of atoms along central
4720 +  column (Decahedron only)\\
4721 +  -t& {\tt -{}-twinAtoms=INT}          & Number of atoms along twin
4722 +  boundary (Decahedron only) \\
4723 +  -p& {\tt -{}-truncatedPlanes=INT}   & Number of truncated planes (Curling-stone Decahedron only)\\
4724 + \hline
4725 + \multicolumn{3}{|l|}{One option from the following group of options is required:} \\
4726 + \hline
4727 +   & {\tt -{}-ico}    & Create an Icosahedral cluster \\
4728 +   & {\tt -{}-deca}   & Create a regualar Decahedral cluster\\
4729 +   & {\tt -{}-ino}    & Create an Ino Decahedral cluster\\
4730 +   & {\tt -{}-marks}  & Create a Marks Decahedral cluster\\
4731 +   & {\tt -{}-stone}  & Create a Curling-stone Decahedral cluster
4732 + \end{longtable}
4733 +
4734 +
4735   \section{\label{section:Hydro}Hydro}
4736   {\tt Hydro} generates resistance tensor ({\tt .diff}) files which are
4737   required when using the Langevin integrator using complex rigid
# Line 3698 | Line 4765 | hydrodynamic calculations will not be performed (defau
4765   \end{longtable}
4766  
4767  
4768 +
4769 +
4770 +
4771   \chapter{\label{section:parallelization} Parallel Simulation Implementation}
4772  
4773   Although processor power is continually improving, it is still
# Line 3781 | Line 4851 | DMR-0079647.
4851   DMR-0079647.
4852  
4853  
4854 < \bibliographystyle{jcc}
4854 > \bibliographystyle{aip}
4855   \bibliography{openmdDoc}
4856  
4857   \end{document}

Diff Legend

Removed lines
+ Added lines
< Changed lines
> Changed lines