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 3709 by gezelter, Wed Nov 24 20:44:51 2010 UTC vs.
Revision 4101 by gezelter, Wed Apr 16 12:53:07 2014 UTC

# Line 17 | Line 17
17   \textwidth 6.5in
18   \brokenpenalty=10000
19   \renewcommand{\baselinestretch}{1.2}
20 + \usepackage[square, comma, sort&compress]{natbib}
21 + \bibpunct{[}{]}{,}{n}{}{;}
22  
23 +
24   %\renewcommand\citemid{\ } % no comma in optional reference note
25 < \lstset{language=C,frame=TB,basicstyle=\tiny,basicstyle=\ttfamily, %
25 > \lstset{language=C,frame=TB,basicstyle=\footnotesize,basicstyle=\ttfamily, %
26          xleftmargin=0.25in, xrightmargin=0.25in,captionpos=b, %
27          abovecaptionskip=0.5cm, belowcaptionskip=0.5cm, escapeinside={~}{~}}
28   \renewcommand{\lstlistingname}{Scheme}
# Line 38 | Line 41
41   \newcolumntype{H}{p{0.75in}}
42   \newcolumntype{I}{p{5in}}
43  
44 + \newcolumntype{J}{p{1.5in}}
45 + \newcolumntype{K}{p{1.2in}}
46 + \newcolumntype{L}{p{1.5in}}
47 + \newcolumntype{M}{p{1.55in}}
48  
42 \title{{\sc OpenMD}: Molecular Dynamics in the Open}
49  
50 < \author{Kelsey M. Stocker, Shenyu Kuang, Charles F. Vardeman II, \\
51 <  Teng Lin, Christopher J. Fennell,  Xiuquan Sun, \\
50 > \title{{\sc OpenMD-2.1}: Molecular Dynamics in the Open}
51 >
52 > \author{Joseph Michalka, James Marr, Kelsey Stocker, Madan Lamichhane,
53 >  Patrick Louden, \\
54 >  Teng Lin, Charles F. Vardeman II, Christopher J. Fennell, Shenyu
55 >  Kuang, Xiuquan Sun, \\
56    Chunlei Li, Kyle Daily, Yang Zheng, Matthew A. Meineke, and \\
57    J. Daniel Gezelter \\
58    Department of Chemistry and Biochemistry\\
# Line 133 | Line 143 | leave an interaction region.
143   leave an interaction region.
144  
145   {\tt Atoms} may also be grouped in more traditional ways into {\tt
146 < bonds}, {\tt bends}, and {\tt torsions}.  These groupings allow the
147 < correct choice of interaction parameters for short-range interactions
148 < to be chosen from the definitions in the {\tt forceField}.
146 >  bonds}, {\tt bends}, {\tt torsions}, and {\tt inversions}.  These
147 > groupings allow the correct choice of interaction parameters for
148 > short-range interactions to be chosen from the definitions in the {\tt
149 >  forceField}.
150  
151   All of these groups of {\tt atoms} are brought together in the {\tt
152   molecule}, which is the fundamental structure for setting up and {\sc
# Line 491 | Line 502 | fs}^{-1}$), and body-fixed moments of inertia ($\mbox{
502   \endhead
503   \hline
504   \endfoot
505 < {\tt forceField} & string & Sets the force field. & Possible force
506 < fields are DUFF, WATER, LJ, EAM, SC, and CLAY. \\
505 > {\tt forceField} & string & Sets the base name for the force field file &
506 > OpenMD appends a {\tt .frc} to the end of this to look for a force
507 > field file.\\
508   {\tt component} & & Defines the molecular components of the system &
509   Every {\tt $<$MetaData$>$} block must have a component statement. \\
510   {\tt minimizer} & string & Chooses a minimizer & Possible minimizers
# Line 562 | Line 574 | column names are: {\sc time, total\_energy, potential\
574   default is the first eight of these columns in order.)  \\
575   & & \multicolumn{2}{p{3.5in}}{Allowed
576   column names are: {\sc time, total\_energy, potential\_energy, kinetic\_energy,
577 < temperature, pressure, volume, conserved\_quantity,
577 > temperature, pressure, volume, conserved\_quantity, hullvolume, gyrvolume,
578   translational\_kinetic, rotational\_kinetic,  long\_range\_potential,
579   short\_range\_potential, vanderwaals\_potential,
580 < electrostatic\_potential, bond\_potential, bend\_potential,
581 < dihedral\_potential, improper\_potential, vraw, vharm,
582 < pressure\_tensor\_x, pressure\_tensor\_y, pressure\_tensor\_z}} \\
580 > electrostatic\_potential, metallic\_potential,
581 > hydrogen\_bonding\_potential, bond\_potential, bend\_potential,
582 > dihedral\_potential, inversion\_potential, raw\_potential, restraint\_potential,
583 > pressure\_tensor, system\_dipole, heatflux, electronic\_temperature}} \\
584   {\tt printPressureTensor} & logical & sets whether {\sc OpenMD} will print
585   out the pressure tensor & can be useful for calculations of the bulk
586   modulus \\
# Line 762 | Line 775 | statistics file is denoted with the \texttt{.stat} fil
775   allowing the user to gauge the stability of the integrator. The
776   statistics file is denoted with the \texttt{.stat} file extension.
777  
778 < \chapter{\label{section:empiricalEnergy}The Empirical Energy
766 < Functions}
778 > \chapter{\label{section:forceFields}Force Fields}
779  
780 < Like many simulation packages, {\sc OpenMD} splits the potential energy
781 < into the short-ranged (bonded) portion and a long-range (non-bonded)
782 < potential,
780 > Like many molecular simulation packages, {\sc OpenMD} splits the
781 > potential energy into the short-ranged (bonded) portion and a
782 > long-range (non-bonded) potential,
783   \begin{equation}
784   V = V_{\mathrm{short-range}} + V_{\mathrm{long-range}}.
785   \end{equation}
786 < The short-ranged portion includes the explicit bonds, bends, and
787 < torsions which have been defined in the meta-data file for the
788 < molecules which are present in the simulation.  The functional forms and
789 < parameters for these interactions are defined by the force field which
790 < is chosen.
786 > The short-ranged portion includes the bonds, bends, torsions, and
787 > inversions which have been defined in the meta-data file for the
788 > molecules.  The functional forms and parameters for these interactions
789 > are defined by the force field which is selected in the MetaData
790 > section.
791  
792 < Calculating the long-range (non-bonded) potential involves a sum over
793 < all pairs of atoms (except for those atoms which are involved in a
794 < bond, bend, or torsion with each other).  If done poorly, calculating
795 < the the long-range interactions for $N$ atoms would involve $N(N-1)/2$
784 < evaluations of atomic distances.  To reduce the number of distance
785 < evaluations between pairs of atoms, {\sc OpenMD} uses a switched cutoff
786 < with Verlet neighbor lists.\cite{Allen87} It is well known that
787 < neutral groups which contain charges will exhibit pathological forces
788 < unless the cutoff is applied to the neutral groups evenly instead of
789 < to the individual atoms.\cite{leach01:mm} {\sc OpenMD} allows users to
790 < specify cutoff groups which may contain an arbitrary number of atoms
791 < in the molecule.  Atoms in a cutoff group are treated as a single unit
792 < for the evaluation of the switching function:
792 > \section{\label{section:shortRange}The basic interactions}
793 >
794 > The energy function for a system composed of $N$ molecules is
795 > traditionally written
796   \begin{equation}
797 < V_{\mathrm{long-range}} = \sum_{a} \sum_{b>a} s(r_{ab}) \sum_{i \in a} \sum_{j \in b} V_{ij}(r_{ij}),
798 < \end{equation}
799 < where $r_{ab}$ is the distance between the centers of mass of the two
797 < cutoff groups ($a$ and $b$).
798 <
799 < The sums over $a$ and $b$ are over the cutoff groups that are present
800 < in the simulation.  Atoms which are not explicitly defined as members
801 < of a {\tt cutoffGroup} are treated as a group consisting of only one
802 < atom.  The switching function, $s(r)$ is the standard cubic switching
803 < function,
804 < \begin{equation}
805 < S(r) =
806 <        \begin{cases}
807 <        1 & \text{if $r \le r_{\text{sw}}$},\\
808 <        \frac{(r_{\text{cut}} + 2r - 3r_{\text{sw}})(r_{\text{cut}} - r)^2}
809 <        {(r_{\text{cut}} - r_{\text{sw}})^3}
810 <        & \text{if $r_{\text{sw}} < r \le r_{\text{cut}}$}, \\
811 <        0 & \text{if $r > r_{\text{cut}}$.}
812 <        \end{cases}
813 < \label{eq:dipoleSwitching}
797 > V = \sum^{N}_{I=1} V^{I}_{\text{Internal}}
798 >        + \sum^{N-1}_{I=1} \sum_{J>I} V^{IJ}_{\text{Cross}},
799 > \label{eq:totalPotential}
800   \end{equation}
801 < Here, $r_{\text{sw}}$ is the {\tt switchingRadius}, or the distance
802 < beyond which interactions are reduced, and $r_{\text{cut}}$ is the
803 < {\tt cutoffRadius}, or the distance at which interactions are
804 < truncated.
801 > where $V^{IJ}_{\text{Cross}}$ contains all intermolecular interactions
802 > between molecules $I$ and $J$, and $V^{I}_{\text{Internal}}$ is the
803 > internal potential of molecule $I$:
804 > \begin{align*}
805 > V^{I}_{\text{Internal}} =  &
806 > \sum_{r_{ij} \in I} V_{\text{bond}}(r_{ij})
807 > + \sum_{\theta_{ijk} \in I} V_{\text{bend}}(\theta_{ijk})
808 > + \sum_{\phi_{ijkl} \in I} V_{\text{torsion}}(\phi_{ijkl})
809 > + \sum_{\omega_{ijkl} \in I} V_{\text{inversion}}(\omega_{ijkl}) \\
810 > & + \sum_{i \in I} \sum_{(j>i+4) \in I}
811 > \biggl[ V_{\text{dispersion}}(r_{ij}) +  V_{\text{electrostatic}}
812 > (\mathbf{r}_{ij},\boldsymbol{\Omega}_{i},\boldsymbol{\Omega}_{j})
813 > \biggr].
814 > \label{eq:internalPotential}
815 > \end{align*}
816 > Here $V_{\text{bond}}, V_{\text{bend}},
817 > V_{\text{torsion}},\mathrm{~and~} V_{\text{inversion}}$ represent the
818 > bond, bend, torsion, and inversion potentials for all
819 > topologically-connected sets of atoms within the molecule.  Bonds are
820 > the primary way of specifying how the atoms are connected together to
821 > form the molecule (i.e. they define the molecular topology).  The
822 > other short-range interactions may be specified explicitly in the
823 > molecule definition, or they may be deduced from bonding information.
824 > For example, bends can be implicitly deduced from two bonds which
825 > share an atom.  Torsions can be deduced from two bends that share a
826 > bond.  Inversion potentials are utilized primarily to enforce
827 > planarity around $sp^2$-hybridized sites, and these are specified with
828 > central atoms and satellites (or an atom with bonds to exactly three
829 > satellites).  The pairwise portions of the non-bonded interactions are
830 > usually excluded for atom pairs that are involved in the same bond,
831 > bend, or torsion. All other atom pairs within a molecule are subject
832 > to non-bonded pair potentials.
833  
834 < Users of {\sc OpenMD} do not need to specify the {\tt cutoffRadius} or
835 < {\tt switchingRadius}.  In simulations containing only Lennard-Jones
836 < atoms, the cutoff radius has a default value of $2.5\sigma_{ii}$,
837 < where $\sigma_{ii}$ is the largest Lennard-Jones length parameter
838 < present in the simulation.  In simulations containing charged or
825 < dipolar atoms, the default cutoff radius is $15 \mbox{\AA}$.  
834 > The types of atoms being simulated, as well as the specific functional
835 > forms and parameters of the intra-molecular functions and the
836 > long-range potentials are defined by the force field. In the following
837 > sections we discuss the stucture of an OpenMD force field file and the
838 > specification of blocks that may be present within these files.
839  
840 < The {\tt switchingRadius} is set to a default value of 95\% of the
828 < {\tt cutoffRadius}.  In the special case of a simulation containing
829 < {\it only} Lennard-Jones atoms, the default switching radius takes the
830 < same value as the cutoff radius, and {\sc OpenMD} will use a shifted
831 < potential to remove discontinuities in the potential at the cutoff.
832 < Both radii may be specified in the meta-data file.
840 > \section{\label{section:frcFile}Force Field Files}
841  
842 < Force fields can be added to {\sc OpenMD}, although it comes with a few
843 < simple examples (Lennard-Jones, {\sc duff}, {\sc water}, and {\sc
844 < eam}) which are explained in the following sections.
842 > Force field files have a number of ``Blocks'' to demarkate different
843 > types of information.  The blocks contain AtomType data, which provide
844 > properties belonging to a single AtomType, as well as interaction
845 > information which provides information about bonded or non-bonded
846 > interactions that cannot be deduced from AtomType information alone.
847 > A simple example of a forceField file is shown in scheme
848 > \ref{sch:frcExample}.
849  
850 < \section{\label{sec:LJPot}The Lennard Jones Force Field}
850 > \begin{lstlisting}[float,caption={[An example of a complete OpenMD
851 > force field file for straight-chain united-atom alkanes.] An example
852 > showing a complete OpenMD force field for straight-chain united-atom
853 > alkanes.}, label={sch:frcExample}]
854 > begin Options
855 >  Name = "alkane" end
856 > Options
857  
858 < The most basic force field implemented in {\sc OpenMD} is the
859 < Lennard-Jones force field, which mimics the van der Waals interaction
860 < at long distances and uses an empirical repulsion at short
858 > begin BaseAtomTypes  
859 > //name          mass  
860 > C               12.0107
861 > end BaseAtomTypes
862 >
863 > begin AtomTypes
864 > //name  base    mass
865 > CH4     C       16.05          
866 > CH3     C       15.04          
867 > CH2     C       14.03          
868 > end AtomTypes
869 >
870 > begin LennardJonesAtomTypes
871 > //name          epsilon         sigma
872 > CH4             0.2941          3.73
873 > CH3             0.1947          3.75
874 > CH2             0.09140         3.95
875 > end LennardJonesAtomTypes
876 >
877 > begin BondTypes
878 > //AT1       AT2 Type                    r0              k
879 > CH3         CH3 Harmonic                1.526           260
880 > CH3         CH2 Harmonic                1.526           260
881 > CH2         CH2 Harmonic                1.526           260
882 > end BondTypes
883 >
884 > begin BendTypes
885 > //AT1   AT2     AT3     Type            theta0   k
886 > CH3     CH2     CH3     Harmonic        114.0    124.19
887 > CH3     CH2     CH2     Harmonic        114.0    124.19
888 > CH2     CH2     CH2     Harmonic        114.0    124.19
889 > end BendTypes
890 >
891 > begin TorsionTypes
892 > //AT1 AT2  AT3  AT4  Type    
893 > CH3   CH2  CH2  CH3  Trappe  0.0  0.70544  -0.13549  1.5723
894 > CH3   CH2  CH2  CH2  Trappe  0.0  0.70544  -0.13549  1.5723  
895 > CH2   CH2  CH2  CH2  Trappe  0.0  0.70544  -0.13549  1.5723  
896 > end TorsionTypes
897 > \end{lstlisting}
898 >
899 > \subsection{\label{section:ffOptions}The Options block}
900 >
901 > The Options block defines properties governing how the force field
902 > interactions are carried out.  This block is delineated with the text
903 > tags {\tt begin Options} and {\tt end Options}.  Most options don't
904 > need to be set as they come with fairly sensible default values, but
905 > the various keywords and their possible values are given in Scheme
906 > \ref{sch:optionsBlock}.
907 >
908 > \begin{lstlisting}[caption={[A force field Options block showing default values
909 > for many force field options.] A force field Options block showing default values
910 > for many force field options.  Most of these options do not need to be
911 > specified if the default values are working.},
912 > label={sch:optionsBlock}]
913 > begin Options
914 > Name                      = "alkane"       // any string
915 > vdWtype                   = "Lennard-Jones"
916 > DistanceMixingRule        = "arithmetic"   // can also be "geometric" or "cubic"
917 > DistanceType              = "sigma"        // can also be Rmin
918 > EnergyMixingRule          = "geometric"    // can also be "arithmetic" or "hhg"
919 > EnergyUnitScaling         = 1.0
920 > MetallicEnergyUnitScaling = 1.0
921 > DistanceUnitScaling       = 1.0
922 > AngleUnitScaling          = 1.0
923 > TorsionAngleConvention    = "180_is_trans" // can also be "0_is_trans"
924 > vdW-12-scale              = 0.0
925 > vdW-13-scale              = 0.0
926 > vdW-14-scale              = 0.0
927 > electrostatic-12-scale    = 0.0
928 > electrostatic-13-scale    = 0.0
929 > electrostatic-14-scale    = 0.0
930 > GayBerneMu                = 2.0
931 > GayBerneNu                = 1.0
932 > EAMMixingMethod           = "Johnson"      // can also be "Daw"
933 > end Options
934 > \end{lstlisting}
935 >
936 > \subsection{\label{section:ffBase}The BaseAtomTypes block}
937 >
938 > An AtomType the primary data structure that OpenMD uses to store
939 > static data about an atom.  Things that belong to AtomType are
940 > universal properties (i.e. does this atom have a fixed charge?  What
941 > is its mass?)  Dynamic properties of an atom are not intended to be
942 > properties of an atom type.  A BaseAtomType can be used to build
943 > extended sets of related atom types that all fall back to one
944 > particular type.  For example, one might want a series of atomTypes
945 > that inherit from more basic types:
946 > \begin{displaymath}
947 > \mathtt{ALA-CA} \rightarrow \mathtt{CT} \rightarrow \mathtt{CSP3} \rightarrow \mathtt{C}
948 > \end{displaymath}
949 > where for each step to the right, the atomType falls back to more
950 > primitive data.  That is, the mass could be a property of the {\tt C}
951 > type, while Lennard-Jones parameters could be properties of the {\tt
952 >  CSP3} type.  {\tt CT} could have charge information and its own set
953 > of Lennard-Jones parameter that override the CSP3 parameters.  And the
954 > {\tt ALA-CA} type might have specific torsion or charge information
955 > that override the lower level types.  A BaseAtomType contains only
956 > information a primitive name and the mass of this atom type.
957 > BaseAtomTypes can also be useful in creating files that can be easily
958 > viewed in visualization programs.  The {\tt Dump2XYZ} utility has the
959 > ability to print out the names of the base atom types for displaying
960 > simulations in Jmol or VMD.
961 >
962 > \begin{lstlisting}[caption={[A simple example of a BaseAtomType
963 > block.] A simple example of a BaseAtomType block.},
964 > label={sch:baseAtomTypesBlock}]
965 > begin BaseAtomTypes
966 > //Name  mass (amu)
967 > H       1.0079
968 > O       15.9994
969 > Si      28.0855
970 > Al      26.981538
971 > Mg      24.3050
972 > Ca      40.078
973 > Fe      55.845
974 > Li      6.941
975 > Na      22.98977
976 > K       39.0983
977 > Cs      132.90545
978 > Ca      40.078
979 > Ba      137.327
980 > Cl      35.453
981 > end BaseAtomTypes
982 > \end{lstlisting}
983 >
984 > \subsection{\label{section:ffAtom}The AtomTypes block}
985 >
986 > AtomTypes inherit most properties from BaseAtomTypes, but can override
987 > their lower-level properties as well.  Scheme \ref{sch:atomTypesBlock}
988 > shows an example where multiple types of oxygen atoms can inherit mass
989 > from the oxygen base type.
990 >
991 > \begin{lstlisting}[caption={[An example of a AtomTypes block.] A
992 > simple example of an AtomType block which
993 > shows how multiple types can inherit from the same base type.},
994 > label={sch:atomTypesBlock}]
995 > begin AtomTypes    
996 > //Name  baseatomtype
997 > h*      H
998 > ho      H
999 > o*      O
1000 > oh      O
1001 > ob      O
1002 > obos    O
1003 > obts    O
1004 > obss    O
1005 > ohs     O
1006 > st      Si
1007 > ao      Al
1008 > at      Al
1009 > mgo     Mg
1010 > mgh     Mg
1011 > cao     Ca
1012 > cah     Ca
1013 > feo     Fe
1014 > lio     Li
1015 > end AtomTypes
1016 > \end{lstlisting}
1017 >
1018 > \subsection{\label{section:ffDirectionalAtom}The DirectionalAtomTypes
1019 >  block}
1020 > DirectionalAtoms have orientational degrees of freedom as well as
1021 > translation, so they have moment of inertia tensors.  
1022 >
1023 > \begin{lstlisting}[caption={[An example of a DirectionalAtomTypes block.] A
1024 > simple example of a DirectionalAtomTypes block.},
1025 > label={sch:datomTypesBlock}]
1026 > begin DirectionalAtomTypes
1027 > //Name          I_xx    I_yy    I_zz    (All moments in (amu*Ang^2)
1028 > SSD             1.7696  0.6145  1.1550  
1029 > SSD_E           1.7696  0.6145  1.1550  
1030 > GBC6H6          88.781  88.781  177.561
1031 > GBCH3OH         4.056   20.258  20.999
1032 > GBH2O           1.777   0.581   1.196
1033 > end DirectionalAtomTypes                    
1034 >
1035 > \end{lstlisting}
1036 >
1037 >
1038 > \subsection{\label{section::ffAtomProperties}AtomType properties}
1039 > \subsubsection{\label{section:ffLJ}The LennardJonesAtomTypes block}
1040 > The most basic interatomic interaction implemented in {\sc OpenMD} is
1041 > the Lennard-Jones potential, which mimics the van der Waals
1042 > interaction at long distances and uses an empirical repulsion at short
1043   distances. The Lennard-Jones potential is given by:
1044   \begin{equation}
1045   V_{\text{LJ}}(r_{ij}) =
# Line 851 | Line 1051 | $\sigma_{ij}$ scales the length of the interaction, an
1051   \end{equation}
1052   where $r_{ij}$ is the distance between particles $i$ and $j$,
1053   $\sigma_{ij}$ scales the length of the interaction, and
1054 < $\epsilon_{ij}$ scales the well depth of the potential. Scheme
855 < \ref{sch:LJFF} gives an example meta-data file that
856 < sets up a system of 108 Ar particles to be simulated using the
857 < Lennard-Jones force field.
1054 > $\epsilon_{ij}$ scales the well depth of the potential.
1055  
859 \begin{lstlisting}[float,caption={[Invocation of the Lennard-Jones
860 force field] A sample startup file for a small Lennard-Jones
861 simulation.},label={sch:LJFF}]
862 <OpenMD>
863  <MetaData>
864 #include "argon.md"
865
866 component{
867  type = "Ar";
868  nMol = 108;
869 }
870
871 forceField = "LJ";
872  </MetaData>
873  <Snapshot>     // not shown in this scheme
874  </Snapshot>
875 </OpenMD>
876 \end{lstlisting}
877
1056   Interactions between dissimilar particles requires the generation of
1057   cross term parameters for $\sigma$ and $\epsilon$. These parameters
1058   are determined using the Lorentz-Berthelot mixing
# Line 889 | Line 1067 | and
1067   \label{eq:epsilonMix}
1068   \end{equation}
1069  
1070 < \section{\label{section:DUFF}Dipolar Unified-Atom Force Field}
1071 <
894 < The dipolar unified-atom force field ({\sc duff}) was developed to
895 < simulate lipid bilayers. These types of simulations require a model
896 < capable of forming bilayers, while still being sufficiently
897 < computationally efficient to allow large systems ($\sim$100's of
898 < phospholipids, $\sim$1000's of waters) to be simulated for long times
899 < ($\sim$10's of nanoseconds). With this goal in mind, {\sc duff} has no
900 < point charges. Charge-neutral distributions are replaced with dipoles,
901 < while most atoms and groups of atoms are reduced to Lennard-Jones
902 < interaction sites. This simplification reduces the length scale of
903 < long range interactions from $\frac{1}{r}$ to $\frac{1}{r^3}$,
904 < removing the need for the computationally expensive Ewald
905 < sum. Instead, Verlet neighbor-lists and cutoff radii are used for the
906 < dipolar interactions, and, if desired, a reaction field may be added
907 < to mimic longer range interactions.
908 <
909 < As an example, lipid head-groups in {\sc duff} are represented as
910 < point dipole interaction sites.  Placing a dipole at the head group's
911 < center of mass mimics the charge separation found in common
912 < phospholipid head groups such as phosphatidylcholine.\cite{Cevc87}
913 < Additionally, a large Lennard-Jones site is located at the
914 < pseudoatom's center of mass. The model is illustrated by the red atom
915 < in Fig.~\ref{fig:lipidModel}. The water model we use to
916 < complement the dipoles of the lipids is a
917 < reparameterization\cite{fennell04} of the soft sticky dipole (SSD)
918 < model of Ichiye
919 < \emph{et al.}\cite{liu96:new_model}
920 <
921 < \begin{figure}
922 < \centering
923 < \includegraphics[width=\linewidth]{lipidModel.pdf}
924 < \caption[A representation of a lipid model in {\sc duff}]{A
925 < representation of the lipid model. $\phi$ is the torsion angle,
926 < $\theta$ is the bend angle, and $\mu$ is the dipole moment of the head
927 < group.}
928 < \label{fig:lipidModel}
929 < \end{figure}
930 <
931 < A set of scalable parameters has been used to model the alkyl groups
932 < with Lennard-Jones sites. For this, parameters from the TraPPE force
933 < field of Siepmann \emph{et al.}\cite{Siepmann1998} have been
934 < utilized. TraPPE is a unified-atom representation of n-alkanes which
935 < is parametrized against phase equilibria using Gibbs ensemble Monte
936 < Carlo simulation techniques.\cite{Siepmann1998} One of the advantages
937 < of TraPPE is that it generalizes the types of atoms in an alkyl chain
938 < to keep the number of pseudoatoms to a minimum; thus, the parameters
939 < for a unified atom such as $\text{CH}_2$ do not change depending on
940 < what species are bonded to it.
941 <
942 < As is required by TraPPE, {\sc duff} also constrains all bonds to be
943 < of fixed length. Typically, bond vibrations are the fastest motions in
944 < a molecular dynamic simulation.  With these vibrations present, small
945 < time steps between force evaluations must be used to ensure adequate
946 < energy conservation in the bond degrees of freedom. By constraining
947 < the bond lengths, larger time steps may be used when integrating the
948 < equations of motion. A simulation using {\sc duff} is illustrated in
949 < Scheme \ref{sch:DUFF}.
950 <
951 < \begin{lstlisting}[float,caption={[Invocation of {\sc duff}]A portion
952 < of a startup file showing a simulation utilizing {\sc
953 < duff}},label={sch:DUFF}]  
954 < <OpenMD>
955 <  <MetaData>
956 < #include "water.md"
957 < #include "lipid.md"
958 <
959 < component{
960 <  type = "simpleLipid_16";
961 <  nMol = 60;
962 < }
963 <
964 < component{
965 <  type = "SSD_water";
966 <  nMol = 1936;
967 < }
968 <
969 < forceField = "DUFF";
970 <  </MetaData>
971 <  <Snapshot>     // not shown in this scheme
972 <  </Snapshot>
973 < </OpenMD>
974 < \end{lstlisting}
975 <
976 < \subsection{\label{section:energyFunctions}{\sc duff} Energy Functions}
977 <
978 < The total potential energy function in {\sc duff} is
979 < \begin{equation}
980 < V = \sum^{N}_{I=1} V^{I}_{\text{Internal}}
981 <        + \sum^{N-1}_{I=1} \sum_{J>I} V^{IJ}_{\text{Cross}},
982 < \label{eq:totalPotential}
983 < \end{equation}
984 < where $V^{I}_{\text{Internal}}$ is the internal potential of molecule $I$:
985 < \begin{equation}
986 < V^{I}_{\text{Internal}} =
987 <        \sum_{\theta_{ijk} \in I} V_{\text{bend}}(\theta_{ijk})
988 <        + \sum_{\phi_{ijkl} \in I} V_{\text{torsion}}(\phi_{ijkl})
989 <        + \sum_{i \in I} \sum_{(j>i+4) \in I}
990 <        \biggl[ V_{\text{LJ}}(r_{ij}) +  V_{\text{dipole}}
991 <        (\mathbf{r}_{ij},\boldsymbol{\Omega}_{i},\boldsymbol{\Omega}_{j})
992 <        \biggr].
993 < \label{eq:internalPotential}
994 < \end{equation}
995 < Here $V_{\text{bend}}$ is the bend potential for all 1, 3 bonded pairs
996 < within the molecule $I$, and $V_{\text{torsion}}$ is the torsion
997 < potential for all 1, 4 bonded pairs.  The pairwise portions of the
998 < non-bonded interactions are excluded for atom pairs that are involved
999 < in the smae bond, bend, or torsion. All other atom pairs within a
1000 < molecule are subject to the LJ pair potential.
1001 <
1002 < The bend potential of a molecule is represented by the following function:
1003 < \begin{equation}
1004 < V_{\text{bend}}(\theta_{ijk}) = k_{\theta}( \theta_{ijk} - \theta_0
1005 < )^2, \label{eq:bendPot}
1006 < \end{equation}
1007 < where $\theta_{ijk}$ is the angle defined by atoms $i$, $j$, and $k$
1008 < (see Fig.~\ref{fig:lipidModel}), $\theta_0$ is the equilibrium
1009 < bond angle, and $k_{\theta}$ is the force constant which determines the
1010 < strength of the harmonic bend. The parameters for $k_{\theta}$ and
1011 < $\theta_0$ are borrowed from those in TraPPE.\cite{Siepmann1998}
1012 <
1013 < The torsion potential and parameters are also borrowed from TraPPE. It is
1014 < of the form:
1015 < \begin{equation}
1016 < V_{\text{torsion}}(\phi) = c_1[1 + \cos \phi]
1017 <        + c_2[1 + \cos(2\phi)]
1018 <        + c_3[1 + \cos(3\phi)],
1019 < \label{eq:origTorsionPot}
1020 < \end{equation}
1021 < where:
1022 < \begin{equation}
1023 < \cos\phi = (\hat{\mathbf{r}}_{ij} \times \hat{\mathbf{r}}_{jk}) \cdot
1024 <        (\hat{\mathbf{r}}_{jk} \times \hat{\mathbf{r}}_{kl}).
1025 < \label{eq:torsPhi}
1026 < \end{equation}
1027 < Here, $\hat{\mathbf{r}}_{\alpha\beta}$ are the set of unit bond
1028 < vectors between atoms $i$, $j$, $k$, and $l$. For computational
1029 < efficiency, the torsion potential has been recast after the method of
1030 < {\sc charmm},\cite{Brooks83} in which the angle series is converted to
1031 < a power series of the form:
1032 < \begin{equation}
1033 < V_{\text{torsion}}(\phi) =  
1034 <        k_3 \cos^3 \phi + k_2 \cos^2 \phi + k_1 \cos \phi + k_0,
1035 < \label{eq:torsionPot}
1036 < \end{equation}
1037 < where:
1038 < \begin{align*}
1039 < k_0 &= c_1 + c_3, \\
1040 < k_1 &= c_1 - 3c_3, \\
1041 < k_2 &= 2 c_2, \\
1042 < k_3 &= 4c_3.
1043 < \end{align*}
1044 < By recasting the potential as a power series, repeated trigonometric
1045 < evaluations are avoided during the calculation of the potential
1046 < energy.
1047 <
1048 <
1049 < The cross potential between molecules $I$ and $J$,
1050 < $V^{IJ}_{\text{Cross}}$, is as follows:
1051 < \begin{equation}
1052 < V^{IJ}_{\text{Cross}} =
1053 <        \sum_{i \in I} \sum_{j \in J}
1054 <        \biggl[ V_{\text{LJ}}(r_{ij}) +  V_{\text{dipole}}
1055 <        (\mathbf{r}_{ij},\boldsymbol{\Omega}_{i},\boldsymbol{\Omega}_{j})
1056 <        + V_{\text{sticky}}
1057 <        (\mathbf{r}_{ij},\boldsymbol{\Omega}_{i},\boldsymbol{\Omega}_{j})
1058 <        \biggr],
1059 < \label{eq:crossPotentail}
1060 < \end{equation}
1061 < where $V_{\text{LJ}}$ is the Lennard Jones potential,
1062 < $V_{\text{dipole}}$ is the dipole dipole potential, and
1063 < $V_{\text{sticky}}$ is the sticky potential defined by the SSD model
1064 < (Sec.~\ref{section:SSD}). Note that not all atom types include all
1065 < interactions.
1066 <
1070 > \subsubsection{\label{section:ffCharge}The ChargeAtomTypes block}
1071 > \subsubsection{\label{section:ffMultipole}The MultipoleAtomTypes  block}
1072   The dipole-dipole potential has the following form:
1073   \begin{equation}
1074   V_{\text{dipole}}(\mathbf{r}_{ij},\boldsymbol{\Omega}_{i},
# Line 1083 | Line 1088 | the unit vector pointing along $\mathbf{r}_{ij}$
1088   the unit vector pointing along $\mathbf{r}_{ij}$
1089   ($\boldsymbol{\hat{r}}_{ij}=\mathbf{r}_{ij}/|\mathbf{r}_{ij}|$).
1090  
1091 < \subsection{\label{section:SSD}The {\sc duff} Water Models: SSD/E
1092 < and SSD/RF}
1091 > \subsubsection{\label{section:ffGB}The FluctuatingChargeAtomTypes  block}
1092 > \subsubsection{\label{section:ffPol}The PolarizableAtomTypes block}
1093 > \subsubsection{\label{section:ffGB}The GayBerneAtomTypes block}
1094 > \subsubsection{\label{section:ffSticky}The StickyAtomTypes block}
1095  
1096 < In the interest of computational efficiency, the default solvent used
1097 < by {\sc OpenMD} is the extended Soft Sticky Dipole (SSD/E) water
1098 < model.\cite{fennell04} The original SSD was developed by Ichiye
1099 < \emph{et al.}\cite{liu96:new_model} as a modified form of the hard-sphere
1093 < water model proposed by Bratko, Blum, and
1096 > One of the solvents used by {\sc OpenMD} is the extended Soft Sticky
1097 > Dipole (SSD/E) water model.\cite{fennell04} The original SSD was
1098 > developed by Ichiye \emph{et al.}\cite{liu96:new_model} as a modified
1099 > form of the hard-sphere water model proposed by Bratko, Blum, and
1100   Luzar.\cite{Bratko85,Bratko95} It consists of a single point dipole
1101   with a Lennard-Jones core and a sticky potential that directs the
1102   particles to assume the proper hydrogen bond orientation in the first
# Line 1155 | Line 1161 | HOH angle in each water molecule. }
1161   \label{fig:ssd}
1162   \end{figure}
1163  
1158
1164   Since SSD/E is a single-point {\it dipolar} model, the force
1165   calculations are simplified significantly relative to the standard
1166   {\it charged} multi-point models. In the original Monte Carlo
# Line 1185 | Line 1190 | models can be found in reference~\cite{fennell04}.
1190   and the drawbacks and benefits of the different density corrected SSD
1191   models can be found in reference~\cite{fennell04}.
1192  
1193 < \section{\label{section:WATER}The {\sc water} Force Field}
1194 <
1190 < In addition to the {\sc duff} force field's solvent description, a
1191 < separate {\sc water} force field has been included for simulating most
1192 < of the common rigid-body water models. This force field includes the
1193 < simple and point-dipolar models (SSD, SSD1, SSD/E, SSD/RF, and DPD
1194 < water), as well as the common charge-based models (SPC, SPC/E, TIP3P,
1195 < TIP4P, and
1196 < TIP5P).\cite{liu96:new_model,Ichiye03,fennell04,Marrink01,Berendsen81,Berendsen87,Jorgensen83,Mahoney00}
1197 < In order to handle these models, charge-charge interactions were
1198 < included in the force-loop:
1199 < \begin{equation}
1200 < V_{\text{charge}}(r_{ij}) = \sum_{ij}\frac{q_iq_je^2}{r_{ij}},
1201 < \end{equation}
1202 < where $q$ represents the charge on particle $i$ or $j$, and $e$ is the
1203 < charge of an electron in Coulombs. The charge-charge interaction
1204 < support is rudimentary in the current version of {\sc OpenMD}.  As with
1205 < the other pair interactions, charges can be simulated with a pure
1206 < cutoff or a reaction field.  The various methods for performing the
1207 < Ewald summation have not yet been included.  The {\sc water} force
1208 < field can be easily expanded through modification of the {\sc water}
1209 < force field file ({\tt WATER.frc}). By adding atom types and inserting
1210 < the appropriate parameters, it is possible to extend the force field
1211 < to handle rigid molecules other than water.
1212 <
1213 < \section{\label{section:eam}Embedded Atom Method}
1214 <
1193 > \subsection{\label{section::ffMetals}Metallic Atom Types}
1194 > \subsubsection{\label{section:ffEAM}The EAMAtomTypes block}
1195   {\sc OpenMD} implements a potential that describes bonding in
1196   transition metal
1197   systems.~\cite{Finnis84,Ercolessi88,Chen90,Qi99,Ercolessi02} This
# Line 1289 | Line 1269 | files.  
1269   $\mbox{kcal mol}^{-1}$ as in the rest of the {\sc OpenMD} force field
1270   files.  
1271  
1272 < \section{\label{section:sc}The Sutton-Chen Force Field}
1272 > \subsubsection{\label{section:ffSC}The SuttonChenAtomTypes block}
1273  
1274   The Sutton-Chen ({\sc sc})~\cite{Chen90} potential has been used to
1275   study a wide range of phenomena in metals.  Although it is similar in
# Line 1325 | Line 1305 | the {\tt forceFieldVariant = "QSC";} line to the meta-
1305   quantum-adapted variant of the {\sc sc} potential, the user would add
1306   the {\tt forceFieldVariant = "QSC";} line to the meta-data file.
1307  
1308 + \subsection{\label{section::ffShortRange}Short Range Interactions}
1309 + \subsubsection{\label{section:ffBond}The BondTypes block}
1310 + \subsubsection{\label{section:ffBend}The BendTypes block}
1311 + A harmonic bend potential is represented by the following function:
1312 + \begin{equation}
1313 + V_{\text{bend}}(\theta_{ijk}) = k_{\theta}( \theta_{ijk} - \theta_0
1314 + )^2, \label{eq:bendPot}
1315 + \end{equation}
1316 + where $\theta_{ijk}$ is the angle defined by atoms $i$, $j$, and $k$,
1317 + $\theta_0$ is the equilibrium bond angle, and $k_{\theta}$ is the
1318 + force constant which determines the strength of the harmonic bend.
1319 +
1320 + \subsubsection{\label{section:ffTorsion}The TorsionTypes block}
1321 + The torsion potential is often represented as a cosine series of the
1322 + form:
1323 + \begin{equation}
1324 + V_{\text{torsion}}(\phi) = c_1[1 + \cos \phi]
1325 +        + c_2[1 + \cos(2\phi)]
1326 +        + c_3[1 + \cos(3\phi)],
1327 + \label{eq:origTorsionPot}
1328 + \end{equation}
1329 + where:
1330 + \begin{equation}
1331 + \cos\phi = (\hat{\mathbf{r}}_{ij} \times \hat{\mathbf{r}}_{jk}) \cdot
1332 +        (\hat{\mathbf{r}}_{jk} \times \hat{\mathbf{r}}_{kl}).
1333 + \label{eq:torsPhi}
1334 + \end{equation}
1335 + Here, $\hat{\mathbf{r}}_{\alpha\beta}$ are the set of unit bond
1336 + vectors between atoms $i$, $j$, $k$, and $l$. For computational
1337 + efficiency, the torsion potential has been recast after the method of
1338 + {\sc charmm},\cite{Brooks83} in which the angle series is converted to
1339 + a power series of the form:
1340 + \begin{equation}
1341 + V_{\text{torsion}}(\phi) =  
1342 +        k_3 \cos^3 \phi + k_2 \cos^2 \phi + k_1 \cos \phi + k_0,
1343 + \label{eq:torsionPot}
1344 + \end{equation}
1345 + where:
1346 + \begin{align*}
1347 + k_0 &= c_1 + c_3, \\
1348 + k_1 &= c_1 - 3c_3, \\
1349 + k_2 &= 2 c_2, \\
1350 + k_3 &= 4c_3.
1351 + \end{align*}
1352 + By recasting the potential as a power series, repeated trigonometric
1353 + evaluations are avoided during the calculation of the potential
1354 + energy.
1355 +
1356 + \subsubsection{\label{section:ffInversion}The InversionTypes block}
1357 + \subsection{\label{section::ffLongRange}Long Range Interactions}
1358 + \subsubsection{\label{section:ffInversion}The NonBondedInteraction block}
1359 +
1360 +
1361 +
1362 + (see Fig.~\ref{fig:lipidModel}), The parameters for $k_{\theta}$ and
1363 + $\theta_0$ are borrowed from those in TraPPE.\cite{Siepmann1998}
1364 +
1365 + Calculating the long-range (non-bonded) potential involves a sum over
1366 + all pairs of atoms (except for those atoms which are involved in a
1367 + bond, bend, or torsion with each other).  If done poorly, calculating
1368 + the the long-range interactions for $N$ atoms would involve $N(N-1)/2$
1369 + evaluations of atomic distances.  To reduce the number of distance
1370 + evaluations between pairs of atoms, {\sc OpenMD} allows the use of
1371 + switched cutoffs with Verlet neighbor lists.\cite{Allen87} Neutral
1372 + groups which contain charges will exhibit pathological forces unless
1373 + the cutoff is applied to the neutral groups evenly instead of to the
1374 + individual atoms.\cite{leach01:mm}  {\sc OpenMD} allows users to
1375 + specify cutoff groups which may contain an arbitrary number of atoms
1376 + in the molecule.  Atoms in a cutoff group are treated as a single unit
1377 + for the evaluation of the switching function:
1378 + \begin{equation}
1379 + V_{\mathrm{long-range}} = \sum_{a} \sum_{b>a} s(r_{ab}) \sum_{i \in a} \sum_{j \in b} V_{ij}(r_{ij}),
1380 + \end{equation}
1381 + where $r_{ab}$ is the distance between the centers of mass of the two
1382 + cutoff groups ($a$ and $b$).
1383 +
1384 + The sums over $a$ and $b$ are over the cutoff groups that are present
1385 + in the simulation.  Atoms which are not explicitly defined as members
1386 + of a {\tt cutoffGroup} are treated as a group consisting of only one
1387 + atom.  The switching function, $s(r)$ is the standard cubic switching
1388 + function,
1389 + \begin{equation}
1390 + S(r) =
1391 +        \begin{cases}
1392 +        1 & \text{if $r \le r_{\text{sw}}$},\\
1393 +        \frac{(r_{\text{cut}} + 2r - 3r_{\text{sw}})(r_{\text{cut}} - r)^2}
1394 +        {(r_{\text{cut}} - r_{\text{sw}})^3}
1395 +        & \text{if $r_{\text{sw}} < r \le r_{\text{cut}}$}, \\
1396 +        0 & \text{if $r > r_{\text{cut}}$.}
1397 +        \end{cases}
1398 + \label{eq:dipoleSwitching}
1399 + \end{equation}
1400 + Here, $r_{\text{sw}}$ is the {\tt switchingRadius}, or the distance
1401 + beyond which interactions are reduced, and $r_{\text{cut}}$ is the
1402 + {\tt cutoffRadius}, or the distance at which interactions are
1403 + truncated.
1404 +
1405 + Users of {\sc OpenMD} do not need to specify the {\tt cutoffRadius} or
1406 + {\tt switchingRadius}.  In simulations containing only Lennard-Jones
1407 + atoms, the cutoff radius has a default value of $2.5\sigma_{ii}$,
1408 + where $\sigma_{ii}$ is the largest Lennard-Jones length parameter
1409 + present in the simulation.  In simulations containing charged or
1410 + dipolar atoms, the default cutoff radius is $15 \mbox{\AA}$.  
1411 +
1412 + The {\tt switchingRadius} is set to a default value of 95\% of the
1413 + {\tt cutoffRadius}.  In the special case of a simulation containing
1414 + {\it only} Lennard-Jones atoms, the default switching radius takes the
1415 + same value as the cutoff radius, and {\sc OpenMD} will use a shifted
1416 + potential to remove discontinuities in the potential at the cutoff.
1417 + Both radii may be specified in the meta-data file.
1418 +
1419 + Force fields can be added to {\sc OpenMD}, although it comes with a few
1420 + simple examples (Lennard-Jones, {\sc duff}, {\sc water}, and {\sc
1421 + eam}) which are explained in the following sections.
1422 +
1423 + \section{\label{sec:LJPot}The Lennard Jones Force Field}
1424 +
1425 + Scheme
1426 + \ref{sch:LJFF} gives an example meta-data file that
1427 + sets up a system of 108 Ar particles to be simulated using the
1428 + Lennard-Jones force field.
1429 +
1430 + \begin{lstlisting}[float,caption={[Invocation of the Lennard-Jones
1431 + force field] A sample startup file for a small Lennard-Jones
1432 + simulation.},label={sch:LJFF}]
1433 + <OpenMD>
1434 +  <MetaData>
1435 + #include "argon.md"
1436 +
1437 + component{
1438 +  type = "Ar";
1439 +  nMol = 108;
1440 + }
1441 +
1442 + forceField = "LJ";
1443 +  </MetaData>
1444 +  <Snapshot>     // not shown in this scheme
1445 +  </Snapshot>
1446 + </OpenMD>
1447 + \end{lstlisting}
1448 +
1449 +
1450 + \section{\label{section:DUFF}Dipolar Unified-Atom Force Field}
1451 +
1452 + The dipolar unified-atom force field ({\sc duff}) was developed to
1453 + simulate lipid bilayers. These types of simulations require a model
1454 + capable of forming bilayers, while still being sufficiently
1455 + computationally efficient to allow large systems ($\sim$100's of
1456 + phospholipids, $\sim$1000's of waters) to be simulated for long times
1457 + ($\sim$10's of nanoseconds). With this goal in mind, {\sc duff} has no
1458 + point charges. Charge-neutral distributions are replaced with dipoles,
1459 + while most atoms and groups of atoms are reduced to Lennard-Jones
1460 + interaction sites. This simplification reduces the length scale of
1461 + long range interactions from $\frac{1}{r}$ to $\frac{1}{r^3}$,
1462 + removing the need for the computationally expensive Ewald
1463 + sum. Instead, Verlet neighbor-lists and cutoff radii are used for the
1464 + dipolar interactions, and, if desired, a reaction field may be added
1465 + to mimic longer range interactions.
1466 +
1467 + As an example, lipid head-groups in {\sc duff} are represented as
1468 + point dipole interaction sites.  Placing a dipole at the head group's
1469 + center of mass mimics the charge separation found in common
1470 + phospholipid head groups such as phosphatidylcholine.\cite{Cevc87}
1471 + Additionally, a large Lennard-Jones site is located at the
1472 + pseudoatom's center of mass. The model is illustrated by the red atom
1473 + in Fig.~\ref{fig:lipidModel}. The water model we use to
1474 + complement the dipoles of the lipids is a
1475 + reparameterization\cite{fennell04} of the soft sticky dipole (SSD)
1476 + model of Ichiye
1477 + \emph{et al.}\cite{liu96:new_model}
1478 +
1479 + \begin{figure}
1480 + \centering
1481 + \includegraphics[width=\linewidth]{lipidModel.pdf}
1482 + \caption[A representation of a lipid model in {\sc duff}]{A
1483 + representation of the lipid model. $\phi$ is the torsion angle,
1484 + $\theta$ is the bend angle, and $\mu$ is the dipole moment of the head
1485 + group.}
1486 + \label{fig:lipidModel}
1487 + \end{figure}
1488 +
1489 + A set of scalable parameters has been used to model the alkyl groups
1490 + with Lennard-Jones sites. For this, parameters from the TraPPE force
1491 + field of Siepmann \emph{et al.}\cite{Siepmann1998} have been
1492 + utilized. TraPPE is a unified-atom representation of n-alkanes which
1493 + is parametrized against phase equilibria using Gibbs ensemble Monte
1494 + Carlo simulation techniques.\cite{Siepmann1998} One of the advantages
1495 + of TraPPE is that it generalizes the types of atoms in an alkyl chain
1496 + to keep the number of pseudoatoms to a minimum; thus, the parameters
1497 + for a unified atom such as $\text{CH}_2$ do not change depending on
1498 + what species are bonded to it.
1499 +
1500 + As is required by TraPPE, {\sc duff} also constrains all bonds to be
1501 + of fixed length. Typically, bond vibrations are the fastest motions in
1502 + a molecular dynamic simulation.  With these vibrations present, small
1503 + time steps between force evaluations must be used to ensure adequate
1504 + energy conservation in the bond degrees of freedom. By constraining
1505 + the bond lengths, larger time steps may be used when integrating the
1506 + equations of motion. A simulation using {\sc duff} is illustrated in
1507 + Scheme \ref{sch:DUFF}.
1508 +
1509 + \begin{lstlisting}[float,caption={[Invocation of {\sc duff}]A portion
1510 + of a startup file showing a simulation utilizing {\sc
1511 + duff}},label={sch:DUFF}]  
1512 + <OpenMD>
1513 +  <MetaData>
1514 + #include "water.md"
1515 + #include "lipid.md"
1516 +
1517 + component{
1518 +  type = "simpleLipid_16";
1519 +  nMol = 60;
1520 + }
1521 +
1522 + component{
1523 +  type = "SSD_water";
1524 +  nMol = 1936;
1525 + }
1526 +
1527 + forceField = "DUFF";
1528 +  </MetaData>
1529 +  <Snapshot>     // not shown in this scheme
1530 +  </Snapshot>
1531 + </OpenMD>
1532 + \end{lstlisting}
1533 +
1534 +
1535 +
1536 + The cross potential between molecules $I$ and $J$,
1537 + $V^{IJ}_{\text{Cross}}$, is as follows:
1538 + \begin{equation}
1539 + V^{IJ}_{\text{Cross}} =
1540 +        \sum_{i \in I} \sum_{j \in J}
1541 +        \biggl[ V_{\text{LJ}}(r_{ij}) +  V_{\text{dipole}}
1542 +        (\mathbf{r}_{ij},\boldsymbol{\Omega}_{i},\boldsymbol{\Omega}_{j})
1543 +        + V_{\text{sticky}}
1544 +        (\mathbf{r}_{ij},\boldsymbol{\Omega}_{i},\boldsymbol{\Omega}_{j})
1545 +        \biggr],
1546 + \label{eq:crossPotentail}
1547 + \end{equation}
1548 + where $V_{\text{LJ}}$ is the Lennard Jones potential,
1549 + $V_{\text{dipole}}$ is the dipole dipole potential, and
1550 + $V_{\text{sticky}}$ is the sticky potential defined by the SSD model
1551 + (Sec.~\ref{section:SSD}). Note that not all atom types include all
1552 + interactions.
1553 +
1554 +
1555 + \section{\label{section:WATER}The {\sc water} Force Field}
1556 +
1557 + In addition to the {\sc duff} force field's solvent description, a
1558 + separate {\sc water} force field has been included for simulating most
1559 + of the common rigid-body water models. This force field includes the
1560 + simple and point-dipolar models (SSD, SSD1, SSD/E, SSD/RF, and DPD
1561 + water), as well as the common charge-based models (SPC, SPC/E, TIP3P,
1562 + TIP4P, and
1563 + TIP5P).\cite{liu96:new_model,Ichiye03,fennell04,Marrink01,Berendsen81,Berendsen87,Jorgensen83,Mahoney00}
1564 + In order to handle these models, charge-charge interactions were
1565 + included in the force-loop:
1566 + \begin{equation}
1567 + V_{\text{charge}}(r_{ij}) = \sum_{ij}\frac{q_iq_je^2}{r_{ij}},
1568 + \end{equation}
1569 + where $q$ represents the charge on particle $i$ or $j$, and $e$ is the
1570 + charge of an electron in Coulombs. The charge-charge interaction
1571 + support is rudimentary in the current version of {\sc OpenMD}.  As with
1572 + the other pair interactions, charges can be simulated with a pure
1573 + cutoff or a reaction field.  The various methods for performing the
1574 + Ewald summation have not yet been included.  The {\sc water} force
1575 + field can be easily expanded through modification of the {\sc water}
1576 + force field file ({\tt WATER.frc}). By adding atom types and inserting
1577 + the appropriate parameters, it is possible to extend the force field
1578 + to handle rigid molecules other than water.
1579 +
1580 +
1581 + \section{\label{section:sc}The Sutton-Chen Force Field}
1582 +
1583 +
1584   \section{\label{section:clay}The CLAY force field}
1585  
1586   The {\sc clay} force field is based on an ionic (nonbonded)
# Line 2621 | Line 2877 | tensor.
2877  
2878   \section{Constant Pressure without periodic boundary conditions (The LangevinHull)}
2879  
2880 < The Langevin Hull uses an external bath at a fixed constant pressure
2880 > The Langevin Hull\cite{Vardeman2011} uses an external bath at a fixed constant pressure
2881   ($P$) and temperature ($T$) with an effective solvent viscosity
2882   ($\eta$).  This bath interacts only with the objects on the exterior
2883   hull of the system.  Defining the hull of the atoms in a simulation is
# Line 2933 | Line 3189 | Harmonic Forces are used by default
3189   \label{table:zconParams}
3190   \end{longtable}
3191  
3192 < \chapter{\label{section:restraints}Restraints}
3193 < Restraints are external potentials that are added to a system to keep
3194 < particular molecules or collections of particles close to some
3195 < reference structure.  A restraint can be a collective
3192 > % \chapter{\label{section:restraints}Restraints}
3193 > % Restraints are external potentials that are added to a system to
3194 > % keep particular molecules or collections of particles close to some
3195 > % reference structure.  A restraint can be a collective
3196  
3197   \chapter{\label{section:thermInt}Thermodynamic Integration}
3198  
# Line 3076 | Line 3332 | Einstein crystal
3332   \label{table:thermIntParams}
3333   \end{longtable}
3334  
3335 + \chapter{\label{section:rnemd}Reverse Non-Equilibrium Molecular Dynamics (RNEMD)}
3336  
3337 + There are many ways to compute transport properties from molecular
3338 + dynamics simulations.  Equilibrium Molecular Dynamics (EMD)
3339 + simulations can be used by computing relevant time correlation
3340 + functions and assuming linear response theory holds.  For some transport properties (notably thermal conductivity), EMD approaches
3341 + are subject to noise and poor convergence of the relevant
3342 + correlation functions. Traditional Non-equilibrium Molecular Dynamics
3343 + (NEMD) methods impose a gradient (e.g. thermal or momentum) on a
3344 + simulation.  However, the resulting flux is often difficult to
3345 + measure. Furthermore, problems arise for NEMD simulations of
3346 + heterogeneous systems, such as phase-phase boundaries or interfaces,
3347 + where the type of gradient to enforce at the boundary between
3348 + materials is unclear.
3349 +
3350 + {\it Reverse} Non-Equilibrium Molecular Dynamics (RNEMD) methods adopt
3351 + a different approach in that an unphysical {\it flux} is imposed
3352 + between different regions or ``slabs'' of the simulation box.  The
3353 + response of the system is to develop a temperature or momentum {\it
3354 +  gradient} between the two regions. Since the amount of the applied
3355 + flux is known exactly, and the measurement of gradient is generally
3356 + less complicated, imposed-flux methods typically take shorter
3357 + simulation times to obtain converged results for transport properties.
3358 +
3359 + \begin{figure}
3360 + \includegraphics[width=\linewidth]{rnemdDemo}
3361 + \caption{The (VSS) RNEMD approach imposes unphysical transfer of both
3362 +  linear momentum and kinetic energy between a ``hot'' slab and a
3363 +  ``cold'' slab in the simulation box.  The system responds to this
3364 +  imposed flux by generating both momentum and temperature gradients.
3365 +  The slope of the gradients can then be used to compute transport
3366 +  properties (e.g. shear viscosity and thermal conductivity).}
3367 + \label{rnemdDemo}
3368 + \end{figure}
3369 +
3370 + \section{\label{section:algo}Three algorithms for carrying out RNEMD simulations}
3371 + \subsection{\label{subsection:swapping}The swapping algorithm}
3372 + The original ``swapping'' approaches by M\"{u}ller-Plathe {\it et
3373 +  al.}\cite{ISI:000080382700030,MullerPlathe:1997xw} can be understood
3374 + as a sequence of imaginary elastic collisions between particles in
3375 + opposite slabs.  In each collision, the entire momentum vectors of
3376 + both particles may be exchanged to generate a thermal
3377 + flux. Alternatively, a single component of the momentum vectors may be
3378 + exchanged to generate a shear flux.  This algorithm turns out to be
3379 + quite useful in many simulations. However, the M\"{u}ller-Plathe
3380 + swapping approach perturbs the system away from ideal
3381 + Maxwell-Boltzmann distributions, and this may leads to undesirable
3382 + side-effects when the applied flux becomes large.\cite{Maginn:2010}
3383 + This limits the applicability of the swapping algorithm, so in OpenMD,
3384 + we have implemented two additional algorithms for RNEMD in addition to the
3385 + original swapping approach.
3386 +
3387 + \subsection{\label{subsection:nivs}Non-Isotropic Velocity Scaling (NIVS)}
3388 + Instead of having momentum exchange between {\it individual particles}
3389 + in each slab, the NIVS algorithm applies velocity scaling to all of
3390 + the selected particles in both slabs.\cite{kuang:164101} A combination of linear
3391 + momentum, kinetic energy, and flux constraint equations governs the
3392 + amount of velocity scaling performed at each step. Interested readers
3393 + should consult ref. \citealp{kuang:164101} for further details on the
3394 + methodology.
3395 +
3396 + NIVS has been shown to be very effective at producing thermal
3397 + gradients and for computing thermal conductivities, particularly for
3398 + heterogeneous interfaces.  Although the NIVS algorithm can also be
3399 + applied to impose a directional momentum flux, thermal anisotropy was
3400 + observed in relatively high flux simulations, and the method is not
3401 + suitable for imposing a shear flux or for computing shear viscosities.
3402 +
3403 + \subsection{\label{subsection:vss}Velocity Shearing and Scaling (VSS)}
3404 + The third RNEMD algorithm implemented in OpenMD utilizes a series of
3405 + simultaneous velocity shearing and scaling exchanges between the two
3406 + slabs.\cite{2012MolPh.110..691K}  This method results in a set of simpler equations to satisfy
3407 + the conservation constraints while creating a desired flux between the
3408 + two slabs.
3409 +
3410 + The VSS approach is versatile in that it may be used to implement both
3411 + thermal and shear transport either separately or simultaneously.
3412 + Perturbations of velocities away from the ideal Maxwell-Boltzmann
3413 + distributions are minimal, and thermal anisotropy is kept to a
3414 + minimum.  This ability to generate simultaneous thermal and shear
3415 + fluxes has been utilized to map out the shear viscosity of SPC/E water
3416 + over a wide range of temperatures (90~K) just with a single simulation.
3417 + VSS-RNEMD also allows the directional momentum flux to have
3418 + arbitary directions, which could aid in the study of anisotropic solid
3419 + surfaces in contact with liquid environments.
3420 +
3421 + \section{\label{section:usingRNEMD}Using OpenMD to perform a RNEMD simulation}
3422 + \subsection{\label{section:rnemdParams} What the user needs to specify}
3423 + To carry out a RNEMD simulation,
3424 + a user must specify a number of parameters in the MetaData (.md) file.
3425 + Because the RNEMD methods have a large number of parameters, these
3426 + must be enclosed in a {\it separate} {\tt RNEMD\{...\}} block.  The most important
3427 + parameters to specify are the {\tt useRNEMD}, {\tt fluxType} and flux
3428 + parameters. Most other parameters (summarized in table
3429 + \ref{table:rnemd}) have reasonable default values.  {\tt fluxType}
3430 + sets up the kind of exchange that will be carried out between the two
3431 + slabs (either Kinetic Energy ({\tt KE}) or momentum ({\tt Px, Py, Pz,
3432 +  Pvector}), or some combination of these).  The flux is specified
3433 + with the use of three possible parameters: {\tt kineticFlux} for
3434 + kinetic energy exchange, as well as {\tt momentumFlux} or {\tt
3435 +  momentumFluxVector} for simulations with directional exchange.
3436 +
3437 + \subsection{\label{section:rnemdResults} Processing the results}
3438 + OpenMD will generate a {\tt .rnemd}
3439 + file with the same prefix as the original {\tt .md} file.  This file
3440 + contains a running average of properties of interest computed within a
3441 + set of bins that divide the simulation cell along the $z$-axis.  The
3442 + first column of the {\tt .rnemd} file is the $z$ coordinate of the
3443 + center of each bin, while following columns may contain the average
3444 + temperature, velocity, or density within each bin.  The output format
3445 + in the {\tt .rnemd} file can be altered with the {\tt outputFields},
3446 + {\tt outputBins}, and {\tt outputFileName} parameters.  A report at
3447 + the top of the {\tt .rnemd} file contains the current exchange totals
3448 + as well as the average flux applied during the simulation.  Using the
3449 + slope of the temperature or velocity gradient obtaine from the {\tt
3450 +  .rnemd} file along with the applied flux, the user can very simply
3451 + arrive at estimates of thermal conductivities ($\lambda$),
3452 + \begin{equation}
3453 + J_z = -\lambda \frac{\partial T}{\partial z},
3454 + \end{equation}
3455 + and shear viscosities ($\eta$),
3456 + \begin{equation}
3457 + j_z(p_x) = -\eta \frac{\partial \langle v_x \rangle}{\partial z}.
3458 + \end{equation}
3459 + Here, the quantities on the left hand side are the actual flux values
3460 + (in the header of the {\tt .rnemd} file), while the slopes are
3461 + obtained from linear fits to the gradients observed in the {\tt
3462 +  .rnemd} file.
3463 +
3464 + More complicated simulations (including interfaces) require a bit more
3465 + care.  Here the second derivative may be required to compute the
3466 + interfacial thermal conductance,
3467 + \begin{align}
3468 +  G^\prime &= \left(\nabla\lambda \cdot \mathbf{\hat{n}}\right)_{z_0} \\
3469 +  &= \frac{\partial}{\partial z}\left(-\frac{J_z}{
3470 +      \left(\frac{\partial T}{\partial z}\right)}\right)_{z_0} \\
3471 +  &= J_z\left(\frac{\partial^2 T}{\partial z^2}\right)_{z_0} \Big/
3472 +  \left(\frac{\partial T}{\partial z}\right)_{z_0}^2.
3473 +  \label{derivativeG}
3474 + \end{align}
3475 + where $z_0$ is the location of the interface between two materials and
3476 + $\mathbf{\hat{n}}$ is a unit vector normal to the interface.  We
3477 + suggest that users interested in interfacial conductance consult
3478 + reference \citealp{kuang:AuThl} for other approaches to computing $G$.
3479 + Users interested in {\it friction coefficients} at heterogeneous
3480 + interfaces may also find reference \citealp{2012MolPh.110..691K}
3481 + useful.
3482 +
3483 + \newpage
3484 +
3485 + \begin{longtable}[c]{JKLM}
3486 + \caption{Meta-data Keywords: Parameters for RNEMD simulations}\\
3487 + \multicolumn{4}{c}{The following keywords must be enclosed inside a {\tt RNEMD\{...\}} block.}
3488 + \\ \hline
3489 + {\bf keyword} & {\bf units} & {\bf use} & {\bf remarks}  \\ \hline
3490 + \endhead
3491 + \hline
3492 + \endfoot
3493 + {\tt useRNEMD} & logical & perform RNEMD? & default is ``false'' \\
3494 + {\tt objectSelection} & string & see section \ref{section:syntax}
3495 + for selection syntax & default is ``select all'' \\
3496 + {\tt method} & string & exchange method & one of the following:
3497 + {\tt Swap, NIVS,} or {\tt VSS}  (default is {\tt VSS}) \\
3498 + {\tt fluxType} & string & what is being exchanged between slabs? & one
3499 + of the following: {\tt KE, Px, Py, Pz, Pvector, KE+Px, KE+Py, KE+Pvector} \\
3500 + {\tt kineticFlux} & kcal mol$^{-1}$ \AA$^{-2}$ fs$^{-1}$ & specify the kinetic energy flux &  \\
3501 + {\tt momentumFlux} & amu \AA$^{-1}$ fs$^{-2}$ & specify the momentum flux & \\
3502 + {\tt momentumFluxVector} & amu \AA$^{-1}$ fs$^{-2}$ & specify the momentum flux when
3503 + {\tt Pvector} is part of the exchange & Vector3d input\\
3504 + {\tt exchangeTime} & fs & how often to perform the exchange & default is 100 fs\\
3505 +
3506 + {\tt slabWidth} & $\mbox{\AA}$ & width of the two exchange slabs & default is $\mathsf{H}_{zz} / 10.0$ \\
3507 + {\tt slabAcenter} & $\mbox{\AA}$ & center of the end slab & default is 0 \\
3508 + {\tt slabBcenter} & $\mbox{\AA}$ & center of the middle slab & default is $\mathsf{H}_{zz} / 2$ \\
3509 + {\tt outputFileName} & string & file name for output histograms & default is the same prefix as the
3510 + .md file, but with the {\tt .rnemd} extension \\
3511 + {\tt outputBins} & int & number of $z$-bins in the output histogram &
3512 + default is 20 \\
3513 + {\tt outputFields} & string & columns to print in the {\tt .rnemd}
3514 + file where each column is separated by a pipe ($\mid$) symbol. & Allowed column names are: {\sc z, temperature, velocity, density} \\
3515 + \label{table:rnemd}
3516 + \end{longtable}
3517 +
3518   \chapter{\label{section:minimizer}Energy Minimization}
3519  
3520 < As one of the basic procedures of molecular modeling, energy
3083 < minimization is used to identify local configurations that are stable
3520 > Energy minimization is used to identify local configurations that are stable
3521   points on the potential energy surface. There is a vast literature on
3522   energy minimization algorithms have been developed to search for the
3523   global energy minimum as well as to find local structures which are
# Line 3207 | Line 3644 | diagram of the class heirarchy:
3644   \begin{figure}
3645   \centering
3646   \includegraphics[width=3in]{heirarchy.pdf}
3647 < \caption[Class heirarchy for StuntDoubles in {\sc OpenMD}-4]{ \\ The
3648 < class heirarchy of StuntDoubles in {\sc OpenMD}-4. The selection
3647 > \caption[Class heirarchy for StuntDoubles in {\sc OpenMD}]{ \\ The
3648 > class heirarchy of StuntDoubles in {\sc OpenMD}. The selection
3649   syntax allows the user to select any of the objects that are descended
3650   from a StuntDouble.}
3651   \label{fig:heirarchy}
# Line 3388 | Line 3825 | VMD. The options available for Dump2XYZ are as follows
3825    -z & {\tt -{}-zconstraint}  &                replace the atom types of zconstraint molecules  (default=off) \\
3826    -r & {\tt -{}-rigidbody}  &                  add a pseudo COM atom to rigidbody  (default=off) \\
3827    -t & {\tt -{}-watertype} &                   replace the atom type of water model (default=on) \\
3828 <  -b & {\tt -{}-basetype}  &                   using base atom type  (default=off) \\
3828 >  -b & {\tt -{}-basetype}  &                   using base atom type
3829 >  (default=off) \\
3830 >  -v& {\tt -{}-velocities}             & Print velocities in xyz file  (default=off)\\
3831 >  -f& {\tt -{}-forces}                 & Print forces xyz file  (default=off)\\
3832 >  -u& {\tt -{}-vectors}                & Print vectors (dipoles, etc) in xyz file  
3833 >                                  (default=off)\\
3834 >  -c& {\tt -{}-charges}                & Print charges in xyz file  (default=off)\\
3835 >  -e& {\tt -{}-efield}                 & Print electric field vector in xyz file  
3836 >                                  (default=off)\\
3837       & {\tt -{}-repeatX=INT}  &                 The number of images to repeat in the x direction  (default=`0') \\
3838       & {\tt -{}-repeatY=INT} &                 The number of images to repeat in the y direction  (default=`0') \\
3839       &  {\tt -{}-repeatZ=INT}  &                The number of images to repeat in the z direction  (default=`0') \\
# Line 3480 | Line 3925 | The options available for {\tt StaticProps} are as fol
3925      & {\tt -{}-sele1=selection script}   & select the first StuntDouble set \\
3926      & {\tt -{}-sele2=selection script}   & select the second StuntDouble set \\
3927      & {\tt -{}-sele3=selection script}   & select the third StuntDouble set \\
3928 <    & {\tt -{}-refsele=selection script} & select reference (can only be used with {\tt -{}-gxyz}) \\
3928 >    & {\tt -{}-refsele=selection script} & select reference (can only
3929 >    be used with {\tt -{}-gxyz}) \\
3930 >    & {\tt -{}-comsele=selection script}
3931 >                               & select stunt doubles for center-of-mass
3932 >                                  reference point\\
3933 >    & {\tt -{}-seleoffset=INT}        & global index offset for a second object (used
3934 >                                  to define a vector between sites in molecule)\\
3935 >
3936      & {\tt -{}-molname=STRING}           & molecule name \\
3937      & {\tt -{}-begin=INT}                & begin internal index \\
3938      & {\tt -{}-end=INT}                  & end internal index \\
3939 +    & {\tt -{}-radius=DOUBLE}            & nanoparticle radius\\
3940   \hline
3941   \multicolumn{3}{|l|}{One option from the following group of options is required:} \\
3942   \hline
3943 <    &  {\tt -{}-gofr}                    &  $g(r)$ \\
3944 <    &  {\tt -{}-r\_theta}                 &  $g(r, \cos(\theta))$ \\
3945 <    &  {\tt -{}-r\_omega}                 &  $g(r, \cos(\omega))$ \\
3946 <    &  {\tt -{}-theta\_omega}             &  $g(\cos(\theta), \cos(\omega))$ \\
3943 >    & {\tt -{}-bo}          & bond order parameter ({\tt -{}-rcut} must be specified) \\
3944 >    & {\tt -{}-bor}         & bond order parameter as a function of
3945 >    radius  ({\tt -{}-rcut} must be specified) \\
3946 >    & {\tt -{}-bad}         & $N(\theta)$ bond angle density within ({\tt -{}-rcut} must be specified) \\
3947 >    & {\tt -{}-count}       & count of molecules matching selection
3948 >    criteria (and associated statistics) \\
3949 >  -g&  {\tt -{}-gofr}                    &  $g(r)$ \\
3950 >    &  {\tt -{}-gofz}                    &  $g(z)$ \\
3951 >    &  {\tt -{}-r\_theta}                &  $g(r, \cos(\theta))$ \\
3952 >    &  {\tt -{}-r\_omega}                &  $g(r, \cos(\omega))$ \\
3953 >    &  {\tt -{}-r\_z}                    &  $g(r, z)$ \\
3954 >    &  {\tt -{}-theta\_omega}            &  $g(\cos(\theta), \cos(\omega))$ \\
3955      &  {\tt -{}-gxyz}                    &  $g(x, y, z)$ \\
3956 <    &  {\tt -{}-p2}                      &  $P_2$ order parameter ({\tt -{}-sele1} and {\tt -{}-sele2} must be specified) \\
3956 >    &  {\tt -{}-twodgofr}                & 2D $g(r)$ (Slab width {\tt -{}-dz} must be specified)\\
3957 >  -p&  {\tt -{}-p2}                      &  $P_2$ order parameter  ({\tt -{}-sele1} must be specified, {\tt -{}-sele2} is optional) \\
3958 >    &  {\tt -{}-rp2}                     &  Ripple order parameter ({\tt -{}-sele1} and {\tt -{}-sele2} must be specified) \\
3959      &  {\tt -{}-scd}                     &  $S_{CD}$ order parameter(either {\tt -{}-sele1}, {\tt -{}-sele2}, {\tt -{}-sele3} are specified or {\tt -{}-molname}, {\tt -{}-begin}, {\tt -{}-end} are specified) \\
3960 <    &  {\tt -{}-density}                 &  density plot ({\tt -{}-sele1} must be specified) \\
3961 <    &  {\tt -{}-slab\_density}           &  slab density ({\tt -{}-sele1} must be specified)
3960 >  -d&  {\tt -{}-density}                 &  density plot \\
3961 >    &  {\tt -{}-slab\_density}           &  slab density \\
3962 >    &  {\tt -{}-p\_angle}                & $p(\cos(\theta))$ ($\theta$
3963 >    is the angle between molecular axis and radial vector from origin\\
3964 >    &  {\tt -{}-hxy}                     & Calculates the undulation  spectrum, $h(x,y)$, of an interface \\
3965 >    &  {\tt -{}-rho\_r}                  & $\rho(r)$\\
3966 >    &  {\tt -{}-angle\_r}                &  $\theta(r)$ (spatially resolves the
3967 >    angle between the molecular axis and the radial vector from the
3968 >    origin)\\
3969 >    &  {\tt -{}-hullvol}                 & hull volume of nanoparticle\\
3970 >    &  {\tt -{}-rodlength}               & length of nanorod\\
3971 >  -Q&  {\tt -{}-tet\_param}              & tetrahedrality order parameter ($Q$)\\
3972 >    &  {\tt -{}-tet\_param\_z}           & spatially-resolved tetrahedrality order
3973 >                                   parameter $Q(z)$\\
3974 >    &  {\tt -{}-rnemdz}                  & slab-resolved RNEMD statistics (temperature,
3975 >                                  density, velocity)\\
3976 >    &  {\tt -{}-rnemdr}                  & shell-resolved RNEMD statistics (temperature,
3977 >                                  density, angular velocity)
3978   \end{longtable}
3979  
3980   \subsection{\label{section:DynamicProps}DynamicProps}
# Line 3536 | Line 4015 | The options available for DynamicProps are as follows:
4015    -o& {\tt -{}-output=filename}        & output file name \\
4016      & {\tt -{}-sele1=selection script} & select first StuntDouble set \\
4017      & {\tt -{}-sele2=selection script} & select second StuntDouble set (if sele2 is not set, use script from sele1) \\
4018 +    & {\tt -{}-order=INT}              & Lengendre Polynomial Order\\
4019 +  -z& {\tt -{}-nzbins=INT}             & Number of $z$ bins (default=`100`)\\
4020 +  -m& {\tt -{}-memory=memory specification}
4021 +                                &Available memory  
4022 +                                  (default=`2G`)\\
4023   \hline
4024   \multicolumn{3}{|l|}{One option from the following group of options is required:} \\
4025   \hline
4026 <  -r& {\tt -{}-rcorr}                  & compute mean square displacement \\
4027 <  -v& {\tt -{}-vcorr}                  & compute velocity correlation function \\
4028 <  -d& {\tt -{}-dcorr}                  & compute dipole correlation function
4026 >  -s& {\tt -{}-selecorr}               & selection correlation function \\
4027 >  -r& {\tt -{}-rcorr}                  & compute mean squared displacement \\
4028 >  -v& {\tt -{}-vcorr}                  & velocity autocorrelation function \\
4029 >  -d& {\tt -{}-dcorr}                  & dipole correlation function \\
4030 >  -l& {\tt -{}-lcorr}                  & Lengendre correlation function \\
4031 >    & {\tt -{}-lcorrZ}                 & Lengendre correlation function binned by $z$ \\
4032 >    & {\tt -{}-cohZ}                   & Lengendre correlation function for OH bond vectors binned by $z$\\
4033 >  -M& {\tt -{}-sdcorr}                 & System dipole correlation function\\
4034 >    & {\tt -{}-r\_rcorr}               & Radial mean squared displacement\\
4035 >    & {\tt -{}-thetacorr}              & Angular mean squared displacement\\
4036 >    & {\tt -{}-drcorr}                 & Directional mean squared displacement for particles with unit vectors\\
4037 >    & {\tt -{}-helfandEcorr}           & Helfand moment for thermal conductvity\\
4038 >  -p& {\tt -{}-momentum}               & Helfand momentum for viscosity\\
4039 >    & {\tt -{}-stresscorr}             & Stress tensor correlation function
4040   \end{longtable}
4041  
4042   \chapter{\label{section:PreparingInput} Preparing Input Configurations}
# Line 3608 | Line 4103 | expect the the input specifier on the command line.
4103   to {\tt atom2md}, but they use a specific input format and do not
4104   expect the the input specifier on the command line.
4105  
4106 +
4107   \section{\label{section:SimpleBuilder}SimpleBuilder}
4108  
4109   {\tt SimpleBuilder} creates simple lattice structures.  It requires an
# Line 3632 | Line 4128 | The options available for SimpleBuilder are as follows
4128      &  {\tt -{}-nx=INT}            &  number of unit cells in x\\
4129      &  {\tt -{}-ny=INT}           &  number of unit cells in y\\
4130      &  {\tt -{}-nz=INT}            &  number of unit cells in z
4131 + \end{longtable}
4132 +
4133 + \section{\label{section:icosahedralBuilder}icosahedralBuilder}
4134 +
4135 + {\tt icosahedralBuilder} creates single-component geometric solids
4136 + that can be useful in simulating nanostructures.  Like the other
4137 + builders, it requires an initial, but skeletal {\sc OpenMD} file to
4138 + specify the component that is to be placed on the lattice.  The total
4139 + number of placed molecules will be shown at the top of the
4140 + configuration file that is generated, and that number may not match
4141 + the original meta-data file, so a new meta-data file is also generated
4142 + which matches the lattice structure.
4143 +
4144 + The options available for icosahedralBuilder are as follows:
4145 + \begin{longtable}[c]{|EFG|}
4146 + \caption{icosahedralBuilder Command-line Options}
4147 + \\ \hline
4148 + {\bf option} & {\bf verbose option} & {\bf behavior} \\ \hline
4149 + \endhead
4150 + \hline
4151 + \endfoot
4152 +  -h& {\tt -{}-help}               & Print help and exit\\
4153 +  -V& {\tt -{}-version}            & Print version and exit\\
4154 +  -o& {\tt -{}-output=STRING}      & Output file name\\
4155 +  -n& {\tt -{}-shells=INT}         & Nanoparticle shells\\
4156 +  -d& {\tt -{}-latticeConstant=DOUBLE} & Lattice spacing in Angstroms for cubic lattice.\\
4157 +  -c& {\tt -{}-columnAtoms=INT}        & Number of atoms along central
4158 +  column (Decahedron only)\\
4159 +  -t& {\tt -{}-twinAtoms=INT}          & Number of atoms along twin
4160 +  boundary (Decahedron only) \\
4161 +  -p& {\tt -{}-truncatedPlanes=INT}   & Number of truncated planes (Curling-stone Decahedron only)\\
4162 + \hline
4163 + \multicolumn{3}{|l|}{One option from the following group of options is required:} \\
4164 + \hline
4165 +   & {\tt -{}-ico}    & Create an Icosahedral cluster \\
4166 +   & {\tt -{}-deca}   & Create a regualar Decahedral cluster\\
4167 +   & {\tt -{}-ino}    & Create an Ino Decahedral cluster\\
4168 +   & {\tt -{}-marks}  & Create a Marks Decahedral cluster\\
4169 +   & {\tt -{}-stone}  & Create a Curling-stone Decahedral cluster
4170   \end{longtable}
4171  
4172 +
4173   \section{\label{section:Hydro}Hydro}
4174   {\tt Hydro} generates resistance tensor ({\tt .diff}) files which are
4175   required when using the Langevin integrator using complex rigid
# Line 3667 | Line 4203 | hydrodynamic calculations will not be performed (defau
4203   \end{longtable}
4204  
4205  
4206 +
4207 +
4208 +
4209   \chapter{\label{section:parallelization} Parallel Simulation Implementation}
4210  
4211   Although processor power is continually improving, it is still
# Line 3750 | Line 4289 | DMR-0079647.
4289   DMR-0079647.
4290  
4291  
4292 < \bibliographystyle{jcc}
4292 > \bibliographystyle{aip}
4293   \bibliography{openmdDoc}
4294  
4295   \end{document}

Diff Legend

Removed lines
+ Added lines
< Changed lines
> Changed lines