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 4102 by gezelter, Fri Apr 18 18:39:38 2014 UTC

# Line 9 | Line 9
9   \usepackage{longtable}
10   \pagestyle{plain}
11   \pagenumbering{arabic}
12 + \usepackage{floatrow}
13   \oddsidemargin 0.0cm
14   \evensidemargin 0.0cm
15   \topmargin -21pt
# Line 17 | Line 18
18   \textwidth 6.5in
19   \brokenpenalty=10000
20   \renewcommand{\baselinestretch}{1.2}
21 + \usepackage[square, comma, sort&compress]{natbib}
22 + \bibpunct{[}{]}{,}{n}{}{;}
23  
24 + \DeclareFloatFont{tiny}{\scriptsize}% "scriptsize" is defined by floatrow, "tiny" not
25 + \floatsetup[table]{font=tiny}
26 +
27 +
28   %\renewcommand\citemid{\ } % no comma in optional reference note
29 < \lstset{language=C,frame=TB,basicstyle=\tiny,basicstyle=\ttfamily, %
29 > \lstset{language=C,frame=TB,basicstyle=\footnotesize\ttfamily, %
30          xleftmargin=0.25in, xrightmargin=0.25in,captionpos=b, %
31          abovecaptionskip=0.5cm, belowcaptionskip=0.5cm, escapeinside={~}{~}}
32   \renewcommand{\lstlistingname}{Scheme}
# Line 38 | Line 45
45   \newcolumntype{H}{p{0.75in}}
46   \newcolumntype{I}{p{5in}}
47  
48 + \newcolumntype{J}{p{1.5in}}
49 + \newcolumntype{K}{p{1.2in}}
50 + \newcolumntype{L}{p{1.5in}}
51 + \newcolumntype{M}{p{1.55in}}
52  
42 \title{{\sc OpenMD}: Molecular Dynamics in the Open}
53  
54 < \author{Kelsey M. Stocker, Shenyu Kuang, Charles F. Vardeman II, \\
55 <  Teng Lin, Christopher J. Fennell,  Xiuquan Sun, \\
54 > \title{{\sc OpenMD-2.2}: Molecular Dynamics in the Open}
55 >
56 > \author{Joseph Michalka, James Marr, Kelsey Stocker, Madan Lamichhane,
57 >  Patrick Louden, \\
58 >  Teng Lin, Charles F. Vardeman II, Christopher J. Fennell, Shenyu
59 >  Kuang, Xiuquan Sun, \\
60    Chunlei Li, Kyle Daily, Yang Zheng, Matthew A. Meineke, and \\
61    J. Daniel Gezelter \\
62    Department of Chemistry and Biochemistry\\
# Line 133 | Line 147 | leave an interaction region.
147   leave an interaction region.
148  
149   {\tt Atoms} may also be grouped in more traditional ways into {\tt
150 < bonds}, {\tt bends}, and {\tt torsions}.  These groupings allow the
151 < correct choice of interaction parameters for short-range interactions
152 < to be chosen from the definitions in the {\tt forceField}.
150 >  bonds}, {\tt bends}, {\tt torsions}, and {\tt inversions}.  These
151 > groupings allow the correct choice of interaction parameters for
152 > short-range interactions to be chosen from the definitions in the {\tt
153 >  forceField}.
154  
155   All of these groups of {\tt atoms} are brought together in the {\tt
156   molecule}, which is the fundamental structure for setting up and {\sc
# Line 491 | Line 506 | fs}^{-1}$), and body-fixed moments of inertia ($\mbox{
506   \endhead
507   \hline
508   \endfoot
509 < {\tt forceField} & string & Sets the force field. & Possible force
510 < fields are DUFF, WATER, LJ, EAM, SC, and CLAY. \\
509 > {\tt forceField} & string & Sets the base name for the force field file &
510 > OpenMD appends a {\tt .frc} to the end of this to look for a force
511 > field file.\\
512   {\tt component} & & Defines the molecular components of the system &
513   Every {\tt $<$MetaData$>$} block must have a component statement. \\
514   {\tt minimizer} & string & Chooses a minimizer & Possible minimizers
# Line 562 | Line 578 | column names are: {\sc time, total\_energy, potential\
578   default is the first eight of these columns in order.)  \\
579   & & \multicolumn{2}{p{3.5in}}{Allowed
580   column names are: {\sc time, total\_energy, potential\_energy, kinetic\_energy,
581 < temperature, pressure, volume, conserved\_quantity,
581 > temperature, pressure, volume, conserved\_quantity, hullvolume, gyrvolume,
582   translational\_kinetic, rotational\_kinetic,  long\_range\_potential,
583   short\_range\_potential, vanderwaals\_potential,
584 < electrostatic\_potential, bond\_potential, bend\_potential,
585 < dihedral\_potential, improper\_potential, vraw, vharm,
586 < pressure\_tensor\_x, pressure\_tensor\_y, pressure\_tensor\_z}} \\
584 > electrostatic\_potential, metallic\_potential,
585 > hydrogen\_bonding\_potential, bond\_potential, bend\_potential,
586 > dihedral\_potential, inversion\_potential, raw\_potential, restraint\_potential,
587 > pressure\_tensor, system\_dipole, heatflux, electronic\_temperature}} \\
588   {\tt printPressureTensor} & logical & sets whether {\sc OpenMD} will print
589   out the pressure tensor & can be useful for calculations of the bulk
590   modulus \\
# Line 762 | Line 779 | statistics file is denoted with the \texttt{.stat} fil
779   allowing the user to gauge the stability of the integrator. The
780   statistics file is denoted with the \texttt{.stat} file extension.
781  
782 < \chapter{\label{section:empiricalEnergy}The Empirical Energy
766 < Functions}
782 > \chapter{\label{section:forceFields}Force Fields}
783  
784 < Like many simulation packages, {\sc OpenMD} splits the potential energy
785 < into the short-ranged (bonded) portion and a long-range (non-bonded)
786 < potential,
784 > Like many molecular simulation packages, {\sc OpenMD} splits the
785 > potential energy into the short-ranged (bonded) portion and a
786 > long-range (non-bonded) potential,
787   \begin{equation}
788   V = V_{\mathrm{short-range}} + V_{\mathrm{long-range}}.
789   \end{equation}
790 < The short-ranged portion includes the explicit bonds, bends, and
791 < torsions which have been defined in the meta-data file for the
792 < molecules which are present in the simulation.  The functional forms and
793 < parameters for these interactions are defined by the force field which
794 < is chosen.
790 > The short-ranged portion includes the bonds, bends, torsions, and
791 > inversions which have been defined in the meta-data file for the
792 > molecules.  The functional forms and parameters for these interactions
793 > are defined by the force field which is selected in the MetaData
794 > section.
795  
796 < Calculating the long-range (non-bonded) potential involves a sum over
781 < all pairs of atoms (except for those atoms which are involved in a
782 < bond, bend, or torsion with each other).  If done poorly, calculating
783 < 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:
793 < \begin{equation}
794 < V_{\mathrm{long-range}} = \sum_{a} \sum_{b>a} s(r_{ab}) \sum_{i \in a} \sum_{j \in b} V_{ij}(r_{ij}),
795 < \end{equation}
796 < where $r_{ab}$ is the distance between the centers of mass of the two
797 < cutoff groups ($a$ and $b$).
796 > \section{\label{section:shortRange}The basic interactions}
797  
798 < The sums over $a$ and $b$ are over the cutoff groups that are present
799 < 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,
798 > The energy function for a system composed of $N$ molecules is
799 > traditionally written
800   \begin{equation}
801 < S(r) =
802 <        \begin{cases}
803 <        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}
801 > V = \sum^{N}_{I=1} V^{I}_{\text{Internal}}
802 >        + \sum^{N-1}_{I=1} \sum_{J>I} V^{IJ}_{\text{Cross}},
803 > \label{eq:totalPotential}
804   \end{equation}
805 < Here, $r_{\text{sw}}$ is the {\tt switchingRadius}, or the distance
806 < beyond which interactions are reduced, and $r_{\text{cut}}$ is the
807 < {\tt cutoffRadius}, or the distance at which interactions are
808 < truncated.
805 > where $V^{IJ}_{\text{Cross}}$ contains all intermolecular interactions
806 > between molecules $I$ and $J$, and $V^{I}_{\text{Internal}}$ is the
807 > internal potential of molecule $I$:
808 > \begin{align*}
809 > V^{I}_{\text{Internal}} =  &
810 > \sum_{r_{ij} \in I} V_{\text{bond}}(r_{ij})
811 > + \sum_{\theta_{ijk} \in I} V_{\text{bend}}(\theta_{ijk})
812 > + \sum_{\phi_{ijkl} \in I} V_{\text{torsion}}(\phi_{ijkl})
813 > + \sum_{\omega_{ijkl} \in I} V_{\text{inversion}}(\omega_{ijkl}) \\
814 > & + \sum_{i \in I} \sum_{(j>i+4) \in I}
815 > \biggl[ V_{\text{dispersion}}(r_{ij}) +  V_{\text{electrostatic}}
816 > (\mathbf{r}_{ij},\boldsymbol{\Omega}_{i},\boldsymbol{\Omega}_{j})
817 > \biggr].
818 > \label{eq:internalPotential}
819 > \end{align*}
820 > Here $V_{\text{bond}}, V_{\text{bend}},
821 > V_{\text{torsion}},\mathrm{~and~} V_{\text{inversion}}$ represent the
822 > bond, bend, torsion, and inversion potentials for all
823 > topologically-connected sets of atoms within the molecule.  Bonds are
824 > the primary way of specifying how the atoms are connected together to
825 > form the molecule (i.e. they define the molecular topology).  The
826 > other short-range interactions may be specified explicitly in the
827 > molecule definition, or they may be deduced from bonding information.
828 > For example, bends can be implicitly deduced from two bonds which
829 > share an atom.  Torsions can be deduced from two bends that share a
830 > bond.  Inversion potentials are utilized primarily to enforce
831 > planarity around $sp^2$-hybridized sites, and these are specified with
832 > central atoms and satellites (or an atom with bonds to exactly three
833 > satellites).  The pairwise portions of the non-bonded interactions are
834 > usually excluded for atom pairs that are involved in the same bond,
835 > bend, or torsion. All other atom pairs within a molecule are subject
836 > to non-bonded pair potentials.
837  
838 < Users of {\sc OpenMD} do not need to specify the {\tt cutoffRadius} or
839 < {\tt switchingRadius}.  In simulations containing only Lennard-Jones
840 < atoms, the cutoff radius has a default value of $2.5\sigma_{ii}$,
841 < where $\sigma_{ii}$ is the largest Lennard-Jones length parameter
842 < present in the simulation.  In simulations containing charged or
825 < dipolar atoms, the default cutoff radius is $15 \mbox{\AA}$.  
838 > The types of atoms being simulated, as well as the specific functional
839 > forms and parameters of the intra-molecular functions and the
840 > long-range potentials are defined by the force field. In the following
841 > sections we discuss the stucture of an OpenMD force field file and the
842 > specification of blocks that may be present within these files.
843  
844 < 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.
844 > \section{\label{section:frcFile}Force Field Files}
845  
846 < Force fields can be added to {\sc OpenMD}, although it comes with a few
847 < simple examples (Lennard-Jones, {\sc duff}, {\sc water}, and {\sc
848 < eam}) which are explained in the following sections.
846 > Force field files have a number of ``Blocks'' to delineate different
847 > types of information.  The blocks contain AtomType data, which provide
848 > properties belonging to a single AtomType, as well as interaction
849 > information which provides information about bonded or non-bonded
850 > interactions that cannot be deduced from AtomType information alone.
851 > A simple example of a forceField file is shown in scheme
852 > \ref{sch:frcExample}.
853  
854 < \section{\label{sec:LJPot}The Lennard Jones Force Field}
854 > \begin{lstlisting}[float,caption={[An example of a complete OpenMD
855 > force field file for straight-chain united-atom alkanes.] An example
856 > showing a complete OpenMD force field for straight-chain united-atom
857 > alkanes.}, label={sch:frcExample}]
858 > begin Options
859 >  Name = "alkane"
860 > end Options
861  
862 < The most basic force field implemented in {\sc OpenMD} is the
863 < Lennard-Jones force field, which mimics the van der Waals interaction
864 < at long distances and uses an empirical repulsion at short
865 < distances. The Lennard-Jones potential is given by:
862 > begin BaseAtomTypes  
863 > //name          mass  
864 > C               12.0107
865 > end BaseAtomTypes
866 >
867 > begin AtomTypes
868 > //name  base    mass
869 > CH4     C       16.05          
870 > CH3     C       15.04          
871 > CH2     C       14.03          
872 > end AtomTypes
873 >
874 > begin LennardJonesAtomTypes
875 > //name          epsilon         sigma
876 > CH4             0.2941          3.73
877 > CH3             0.1947          3.75
878 > CH2             0.09140         3.95
879 > end LennardJonesAtomTypes
880 >
881 > begin BondTypes
882 > //AT1       AT2 Type                    r0              k
883 > CH3         CH3 Harmonic                1.526           260
884 > CH3         CH2 Harmonic                1.526           260
885 > CH2         CH2 Harmonic                1.526           260
886 > end BondTypes
887 >
888 > begin BendTypes
889 > //AT1   AT2     AT3     Type            theta0   k
890 > CH3     CH2     CH3     Harmonic        114.0    124.19
891 > CH3     CH2     CH2     Harmonic        114.0    124.19
892 > CH2     CH2     CH2     Harmonic        114.0    124.19
893 > end BendTypes
894 >
895 > begin TorsionTypes
896 > //AT1 AT2  AT3  AT4  Type    
897 > CH3   CH2  CH2  CH3  Trappe  0.0  0.70544  -0.13549  1.5723
898 > CH3   CH2  CH2  CH2  Trappe  0.0  0.70544  -0.13549  1.5723  
899 > CH2   CH2  CH2  CH2  Trappe  0.0  0.70544  -0.13549  1.5723  
900 > end TorsionTypes
901 > \end{lstlisting}
902 >
903 > \subsection{\label{section:ffOptions}The Options block}
904 >
905 > The Options block defines properties governing how the force field
906 > interactions are carried out.  This block is delineated with the text
907 > tags {\tt begin Options} and {\tt end Options}.  Most options don't
908 > need to be set as they come with fairly sensible default values, but
909 > the various keywords and their possible values are given in Scheme
910 > \ref{sch:optionsBlock}.
911 >
912 > \begin{lstlisting}[caption={[A force field Options block showing default values
913 > for many force field options.] A force field Options block showing default values
914 > for many force field options.  Most of these options do not need to be
915 > specified if the default values are working.},
916 > label={sch:optionsBlock}]
917 > begin Options
918 > Name                      = "alkane"       // any string
919 > vdWtype                   = "Lennard-Jones"
920 > DistanceMixingRule        = "arithmetic"   // can also be "geometric" or "cubic"
921 > DistanceType              = "sigma"        // can also be Rmin
922 > EnergyMixingRule          = "geometric"    // can also be "arithmetic" or "hhg"
923 > EnergyUnitScaling         = 1.0
924 > MetallicEnergyUnitScaling = 1.0
925 > DistanceUnitScaling       = 1.0
926 > AngleUnitScaling          = 1.0
927 > TorsionAngleConvention    = "180_is_trans" // can also be "0_is_trans"
928 > vdW-12-scale              = 0.0
929 > vdW-13-scale              = 0.0
930 > vdW-14-scale              = 0.0
931 > electrostatic-12-scale    = 0.0
932 > electrostatic-13-scale    = 0.0
933 > electrostatic-14-scale    = 0.0
934 > GayBerneMu                = 2.0
935 > GayBerneNu                = 1.0
936 > EAMMixingMethod           = "Johnson"      // can also be "Daw"
937 > end Options
938 > \end{lstlisting}
939 >
940 > \subsection{\label{section:ffBase}The BaseAtomTypes block}
941 >
942 > An AtomType the primary data structure that OpenMD uses to store
943 > static data about an atom.  Things that belong to AtomType are
944 > universal properties (i.e. does this atom have a fixed charge?  What
945 > is its mass?)  Dynamic properties of an atom are not intended to be
946 > properties of an atom type.  A BaseAtomType can be used to build
947 > extended sets of related atom types that all fall back to one
948 > particular type.  For example, one might want a series of atomTypes
949 > that inherit from more basic types:
950 > \begin{displaymath}
951 > \mathtt{ALA-CA} \rightarrow \mathtt{CT} \rightarrow \mathtt{CSP3} \rightarrow \mathtt{C}
952 > \end{displaymath}
953 > where for each step to the right, the atomType falls back to more
954 > primitive data.  That is, the mass could be a property of the {\tt C}
955 > type, while Lennard-Jones parameters could be properties of the {\tt
956 >  CSP3} type.  {\tt CT} could have charge information and its own set
957 > of Lennard-Jones parameter that override the CSP3 parameters.  And the
958 > {\tt ALA-CA} type might have specific torsion or charge information
959 > that override the lower level types.  A BaseAtomType contains only
960 > information a primitive name and the mass of this atom type.
961 > BaseAtomTypes can also be useful in creating files that can be easily
962 > viewed in visualization programs.  The {\tt Dump2XYZ} utility has the
963 > ability to print out the names of the base atom types for displaying
964 > simulations in Jmol or VMD.
965 >
966 > \begin{lstlisting}[caption={[A simple example of a BaseAtomTypes
967 > block.] A simple example of a BaseAtomTypes block.},
968 > label={sch:baseAtomTypesBlock}]
969 > begin BaseAtomTypes
970 > //Name  mass (amu)
971 > H       1.0079
972 > O       15.9994
973 > Si      28.0855
974 > Al      26.981538
975 > Mg      24.3050
976 > Ca      40.078
977 > Fe      55.845
978 > Li      6.941
979 > Na      22.98977
980 > K       39.0983
981 > Cs      132.90545
982 > Ca      40.078
983 > Ba      137.327
984 > Cl      35.453
985 > end BaseAtomTypes
986 > \end{lstlisting}
987 >
988 > \subsection{\label{section:ffAtom}The AtomTypes block}
989 >
990 > AtomTypes inherit most properties from BaseAtomTypes, but can override
991 > their lower-level properties as well.  Scheme \ref{sch:atomTypesBlock}
992 > shows an example where multiple types of oxygen atoms can inherit mass
993 > from the oxygen base type.
994 >
995 > \begin{lstlisting}[caption={[An example of a AtomTypes block.] A
996 > simple example of an AtomTypes block which
997 > shows how multiple types can inherit from the same base type.},
998 > label={sch:atomTypesBlock}]
999 > begin AtomTypes    
1000 > //Name  baseatomtype
1001 > h*      H
1002 > ho      H
1003 > o*      O
1004 > oh      O
1005 > ob      O
1006 > obos    O
1007 > obts    O
1008 > obss    O
1009 > ohs     O
1010 > st      Si
1011 > ao      Al
1012 > at      Al
1013 > mgo     Mg
1014 > mgh     Mg
1015 > cao     Ca
1016 > cah     Ca
1017 > feo     Fe
1018 > lio     Li
1019 > end AtomTypes
1020 > \end{lstlisting}
1021 >
1022 > \subsection{\label{section:ffDirectionalAtom}The DirectionalAtomTypes
1023 >  block}
1024 > DirectionalAtoms have orientational degrees of freedom as well as
1025 > translation, so moving these atoms requires information about the
1026 > moments of inertias in the same way that translational motion requires
1027 > mass.  For DirectionalAtoms, OpenMD treats the mass distribution with
1028 > higher priority than electrostatic distributions; the moment of
1029 > inertia tensor, $\overleftrightarrow{\mathsf I}$, should be
1030 > diagonalized to obtain body-fixed axes, and the three diagonal moments
1031 > should correspond to rotational motion \textit{around} each of these
1032 > body-fixed axes.  Charge distributions may then result in dipole
1033 > vectors that are oriented along a linear combination of the body-axes,
1034 > and in quadrupole tensors that are not necessarily diagonal in the
1035 > body frame.
1036 >
1037 > \begin{lstlisting}[caption={[An example of a DirectionalAtomTypes block.] A
1038 > simple example of a DirectionalAtomTypes block.},
1039 > label={sch:datomTypesBlock}]
1040 > begin DirectionalAtomTypes
1041 > //Name          I_xx    I_yy    I_zz    (All moments in (amu*Ang^2)
1042 > SSD             1.7696  0.6145  1.1550  
1043 > GBC6H6          88.781  88.781  177.561
1044 > GBCH3OH         4.056   20.258  20.999
1045 > GBH2O           1.777   0.581   1.196
1046 > CO2             43.06   43.06   0.0    // single-site model for CO2
1047 > end DirectionalAtomTypes                    
1048 >
1049 > \end{lstlisting}
1050 >
1051 > For a DirectionalAtom that represents a linear object, it is
1052 > appropriate for one of the moments of inertia to be zero.  In this
1053 > case, OpenMD identifies that DirectionalAtom as having only 5 degrees
1054 > of freedom (three translations and two rotations), and will alter
1055 > calculation of temperatures to reflect this.
1056 >
1057 > \subsection{\label{section::ffAtomProperties}AtomType properties}
1058 > \subsubsection{\label{section:ffLJ}The LennardJonesAtomTypes block}
1059 > One of the most basic interatomic interactions implemented in {\sc
1060 >  OpenMD} is the Lennard-Jones potential, which mimics the van der
1061 > Waals interaction at long distances and uses an empirical repulsion at
1062 > short distances. The Lennard-Jones potential is given by:
1063   \begin{equation}
1064   V_{\text{LJ}}(r_{ij}) =
1065          4\epsilon_{ij} \biggl[
# Line 851 | Line 1070 | $\sigma_{ij}$ scales the length of the interaction, an
1070   \end{equation}
1071   where $r_{ij}$ is the distance between particles $i$ and $j$,
1072   $\sigma_{ij}$ scales the length of the interaction, and
1073 < $\epsilon_{ij}$ 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.
1073 > $\epsilon_{ij}$ scales the well depth of the potential.
1074  
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
1075   Interactions between dissimilar particles requires the generation of
1076   cross term parameters for $\sigma$ and $\epsilon$. These parameters
1077 < are determined using the Lorentz-Berthelot mixing
1077 > are usually determined using the Lorentz-Berthelot mixing
1078   rules:\cite{Allen87}
1079   \begin{equation}
1080   \sigma_{ij} = \frac{1}{2}[\sigma_{ii} + \sigma_{jj}],
# Line 889 | Line 1086 | and
1086   \label{eq:epsilonMix}
1087   \end{equation}
1088  
1089 < \section{\label{section:DUFF}Dipolar Unified-Atom Force Field}
1089 > The values of $\sigma_{ii}$ and $\epsilon_{ii}$ are properties of atom
1090 > type $i$, and must be specified in a section of the force field file
1091 > called the {\tt LennardJonesAtomTypes} block (see listing
1092 > \ref{sch:LJatomTypesBlock}).  Separate Lennard-Jones interactions
1093 > which are not determined by the mixing rules can also be specified in
1094 > the {\tt NonbondedInteractionTypes} block (see section
1095 > \ref{section:ffNBinteraction}).
1096  
1097 < The dipolar unified-atom force field ({\sc duff}) was developed to
1098 < simulate lipid bilayers. These types of simulations require a model
1099 < capable of forming bilayers, while still being sufficiently
1100 < computationally efficient to allow large systems ($\sim$100's of
1101 < phospholipids, $\sim$1000's of waters) to be simulated for long times
1102 < ($\sim$10's of nanoseconds). With this goal in mind, {\sc duff} has no
1103 < point charges. Charge-neutral distributions are replaced with dipoles,
1104 < while most atoms and groups of atoms are reduced to Lennard-Jones
1105 < interaction sites. This simplification reduces the length scale of
1106 < long range interactions from $\frac{1}{r}$ to $\frac{1}{r^3}$,
1107 < removing the need for the computationally expensive Ewald
1108 < sum. Instead, Verlet neighbor-lists and cutoff radii are used for the
1109 < dipolar interactions, and, if desired, a reaction field may be added
1110 < to mimic longer range interactions.
1111 <
1112 < As an example, lipid head-groups in {\sc duff} are represented as
1113 < 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>
1097 > \begin{lstlisting}[caption={[An example of a LennardJonesAtomTypes block.] A
1098 > simple example of a LennardJonesAtomTypee block.   Units for
1099 > $\epsilon$ are kcal / mol and for $\sigma$ are \AA\ .},
1100 > label={sch:LJatomTypesBlock}]
1101 > begin LennardJonesAtomTypes
1102 > //Name          epsilon             sigma      
1103 > O_TIP4P         0.1550          3.15365
1104 > O_TIP4P-Ew      0.16275         3.16435
1105 > O_TIP5P         0.16            3.12  
1106 > O_TIP5P-E       0.178           3.097  
1107 > O_SPCE          0.15532         3.16549
1108 > O_SPC           0.15532         3.16549
1109 > CH4             0.279           3.73
1110 > CH3             0.185           3.75
1111 > CH2             0.0866          3.95
1112 > CH              0.0189          4.68
1113 > end LennardJonesAtomTypes
1114   \end{lstlisting}
1115  
1116 < \subsection{\label{section:energyFunctions}{\sc duff} Energy Functions}
1116 > \subsubsection{\label{section:ffCharge}The ChargeAtomTypes block}
1117  
1118 < The total potential energy function in {\sc duff} is
1118 > In molecular simulations, proper accumulation of the electrostatic
1119 > interactions is essential and is one of the most
1120 > computationally-demanding tasks.  Most common molecular mechanics
1121 > force fields represent atomic sites with full or partial charges
1122 > protected by Lennard-Jones (short range) interactions.  Partial charge
1123 > values, $q_i$ are empirical representations of the distribution of
1124 > electronic charge in a molecule.  This means that nearly every pair
1125 > interaction involves a calculation of charge-charge forces.  Coupled
1126 > with relatively long-ranged $r^{-1}$ decay, the monopole interactions
1127 > quickly become the most expensive part of molecular simulations.  The
1128 > interactions between point charges can be handled via a number of
1129 > different algorithms, but Coulomb's law is the fundamental physical
1130 > principle governing these interactions,
1131   \begin{equation}
1132 < V = \sum^{N}_{I=1} V^{I}_{\text{Internal}}
1133 <        + \sum^{N-1}_{I=1} \sum_{J>I} V^{IJ}_{\text{Cross}},
982 < \label{eq:totalPotential}
1132 >  V_{\text{charge}}(r_{ij}) = \sum_{ij}\frac{q_iq_je^2}{4 \pi \epsilon_0
1133 >    r_{ij}},
1134   \end{equation}
1135 < where $V^{I}_{\text{Internal}}$ is the internal potential of molecule $I$:
1136 < \begin{equation}
1137 < 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.
1135 > where $q$ represents the charge on particle $i$ or $j$, and $e$ is the
1136 > charge of an electron in Coulombs.  $\epsilon_0$ is the permittivity
1137 > of free space.
1138  
1139 < The bend potential of a molecule is represented by the following function:
1140 < \begin{equation}
1141 < V_{\text{bend}}(\theta_{ijk}) = k_{\theta}( \theta_{ijk} - \theta_0
1142 < )^2, \label{eq:bendPot}
1143 < \end{equation}
1144 < where $\theta_{ijk}$ is the angle defined by atoms $i$, $j$, and $k$
1145 < (see Fig.~\ref{fig:lipidModel}), $\theta_0$ is the equilibrium
1146 < bond angle, and $k_{\theta}$ is the force constant which determines the
1147 < strength of the harmonic bend. The parameters for $k_{\theta}$ and
1148 < $\theta_0$ are borrowed from those in TraPPE.\cite{Siepmann1998}
1139 > \begin{lstlisting}[caption={[An example of a ChargeAtomTypes block.] A
1140 > simple example of a ChargeAtomTypes block.   Units for
1141 > charge are in multiples of electron charge.},
1142 > label={sch:ChargeAtomTypesBlock}]
1143 > begin ChargeAtomTypes
1144 > // Name         charge
1145 > O_TIP3P        -0.834
1146 > O_SPCE         -0.8476
1147 > H_TIP3P         0.417
1148 > H_TIP4P         0.520
1149 > H_SPCE          0.4238
1150 > EP_TIP4P       -1.040
1151 > Na+             1.0
1152 > Cl-            -1.0
1153 > end ChargeAtomTypes
1154 > \end{lstlisting}
1155  
1156 < The torsion potential and parameters are also borrowed from TraPPE. It is
1157 < of the form:
1156 > \subsubsection{\label{section:ffMultipole}The MultipoleAtomTypes
1157 >  block}
1158 > For complex charge distributions that are centered on single sites, it
1159 > is convenient to write the total electrostatic potential in terms of
1160 > multipole moments,
1161   \begin{equation}
1162 < 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}
1162 > U_{\bf{ab}}(r)=\hat{M}_{\bf a} \hat{M}_{\bf b} \frac{1}{r}  \label{kernel}.
1163   \end{equation}
1164 < where:
1164 > where the multipole operator on site $\bf a$,
1165   \begin{equation}
1166 < \cos\phi = (\hat{\mathbf{r}}_{ij} \times \hat{\mathbf{r}}_{jk}) \cdot
1167 <        (\hat{\mathbf{r}}_{jk} \times \hat{\mathbf{r}}_{kl}).
1168 < \label{eq:torsPhi}
1166 > \hat{M}_{\bf a} = C_{\bf a} - D_{{\bf a}\alpha} \frac{\partial}{\partial r_{\alpha}}
1167 > +  Q_{{\bf a}\alpha\beta}
1168 > \frac{\partial^2}{\partial r_{\alpha} \partial r_{\beta}} + \dots
1169   \end{equation}
1170 < Here, $\hat{\mathbf{r}}_{\alpha\beta}$ are the set of unit bond
1171 < vectors between atoms $i$, $j$, $k$, and $l$. For computational
1172 < efficiency, the torsion potential has been recast after the method of
1173 < {\sc charmm},\cite{Brooks83} in which the angle series is converted to
1174 < a power series of the form:
1175 < \begin{equation}
1176 < V_{\text{torsion}}(\phi) =  
1177 <        k_3 \cos^3 \phi + k_2 \cos^2 \phi + k_1 \cos \phi + k_0,
1178 < \label{eq:torsionPot}
1179 < \end{equation}
1180 < where:
1181 < \begin{align*}
1182 < k_0 &= c_1 + c_3, \\
1183 < 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.
1170 > Here, the point charge, dipole, and quadrupole for site $\bf a$ are
1171 > given by $C_{\bf a}$, $D_{{\bf a}\alpha}$, and $Q_{{\bf
1172 >    a}\alpha\beta}$, respectively.  These are the primitive
1173 > multipoles.  If the site is representing a distribution of charges,
1174 > these can be expressed as,
1175 > \begin{align}
1176 > C_{\bf a} =&\sum_{k \, \text{in \bf a}} q_k , \label{eq:charge} \\
1177 > D_{{\bf a}\alpha} =&\sum_{k \, \text{in \bf a}} q_k r_{k\alpha}, \label{eq:dipole}\\
1178 > Q_{{\bf a}\alpha\beta} =& \frac{1}{2} \sum_{k \, \text{in \bf a}} q_k
1179 > r_{k\alpha}  r_{k\beta} . \label{eq:quadrupole}
1180 > \end{align}
1181 > Note that the definition of the primitive quadrupole here differs from
1182 > the standard traceless form, and contains an additional Taylor-series
1183 > based factor of $1/2$.  
1184  
1185 <
1186 < The cross potential between molecules $I$ and $J$,
1050 < $V^{IJ}_{\text{Cross}}$, is as follows:
1185 > The details of the multipolar interactions will be given later, but
1186 > many readers are familiar with the dipole-dipole potential:
1187   \begin{equation}
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
1067 The dipole-dipole potential has the following form:
1068 \begin{equation}
1188   V_{\text{dipole}}(\mathbf{r}_{ij},\boldsymbol{\Omega}_{i},
1189 <        \boldsymbol{\Omega}_{j}) = \frac{|\mu_i||\mu_j|}{4\pi\epsilon_{0}r_{ij}^{3}} \biggl[
1189 >        \boldsymbol{\Omega}_{j}) = \frac{|{\bf D}_i||{\bf D}_j|}{4\pi\epsilon_{0}r_{ij}^{3}} \biggl[
1190          \boldsymbol{\hat{u}}_{i} \cdot \boldsymbol{\hat{u}}_{j}
1191          -
1192          3(\boldsymbol{\hat{u}}_i \cdot \hat{\mathbf{r}}_{ij}) %
# Line 1077 | Line 1196 | are the orientational degrees of freedom for atoms $i$
1196   Here $\mathbf{r}_{ij}$ is the vector starting at atom $i$ pointing
1197   towards $j$, and $\boldsymbol{\Omega}_i$ and $\boldsymbol{\Omega}_j$
1198   are the orientational degrees of freedom for atoms $i$ and $j$
1199 < respectively. The magnitude of the dipole moment of atom $i$ is
1200 < $|\mu_i|$, $\boldsymbol{\hat{u}}_i$ is the standard unit orientation
1199 > respectively. The magnitude of the dipole moment of atom $i$ is $|{\bf
1200 >  D}_i|$, $\boldsymbol{\hat{u}}_i$ is the standard unit orientation
1201   vector of $\boldsymbol{\Omega}_i$, and $\boldsymbol{\hat{r}}_{ij}$ is
1202   the unit vector pointing along $\mathbf{r}_{ij}$
1203   ($\boldsymbol{\hat{r}}_{ij}=\mathbf{r}_{ij}/|\mathbf{r}_{ij}|$).
1204  
1086 \subsection{\label{section:SSD}The {\sc duff} Water Models: SSD/E
1087 and SSD/RF}
1205  
1206 < In the interest of computational efficiency, the default solvent used
1207 < by {\sc OpenMD} is the extended Soft Sticky Dipole (SSD/E) water
1208 < model.\cite{fennell04} The original SSD was developed by Ichiye
1209 < \emph{et al.}\cite{liu96:new_model} as a modified form of the hard-sphere
1206 > \begin{lstlisting}[caption={[An example of a MultipoleAtomTypes block.] A
1207 > simple example of a MultipoleAtomTypes block.   Dipoles are given in
1208 > units of Debyes, and Quadrupole moments are given in units of Debye
1209 > \AA~(or $10^{-26} \mathrm{~esu~cm}^2$)},
1210 > label={sch:MultipoleAtomTypesBlock}]
1211 > begin MultipoleAtomTypes
1212 > // Euler angles are given in zxz convention in units of degrees.
1213 > //
1214 > // point dipoles:
1215 > // name d phi theta psi dipole_moment
1216 > DIP     d 0.0 0.0   0.0     1.91   // dipole points along z-body axis
1217 > //
1218 > // point quadrupoles:
1219 > // name q phi theta psi Qxx Qyy Qzz
1220 > CO2     q 0.0 0.0   0.0 0.0 0.0 -0.430592  //quadrupole tensor has zz element
1221 > //
1222 > // Atoms with both dipole and quadrupole moments:
1223 > // name dq phi theta psi dipole_moment  Qxx    Qyy     Qzz
1224 > SSD     dq 0.0 0.0   0.0     2.35      -1.682  1.762   -0.08
1225 > end MultipoleAtomTypes
1226 > \end{lstlisting}
1227 >
1228 > Specifying a MultipoleAtomType requires declaring how the
1229 > electrostatic frame for the site is rotated relative to the body-fixed
1230 > axes for that atom. The Euler angles $(\phi, \theta, \psi)$ for this
1231 > rotation must be given, and then the dipole, quadrupole, or all of
1232 > these moments are specified in the electrostatic frame.  In OpenMD,
1233 > the Euler angles are specified in the $zxz$ convention and are entered
1234 > in units of degrees.  Dipole moments are entered in units of Debye,
1235 > and Quadrupole moments in units of Debye \AA.
1236 >
1237 > \subsubsection{\label{section:ffGB}The FluctuatingChargeAtomTypes  block}
1238 > \subsubsection{\label{section:ffPol}The PolarizableAtomTypes block}
1239 > \subsubsection{\label{section:ffGB}The GayBerneAtomTypes block}
1240 >
1241 > The Gay-Berne potential has been widely used in the liquid crystal
1242 > community to describe this anisotropic phase
1243 > behavior.~\cite{Gay:1981yu,Berne:1972pb,Kushick:1976xy,Luckhurst:1990fy,Cleaver:1996rt}
1244 > The form of the Gay-Berne potential implemented in OpenMD was
1245 > generalized by Cleaver {\it et al.} and is appropriate for dissimilar
1246 > uniaxial ellipsoids.\cite{Cleaver:1996rt}  The potential is constructed in the
1247 > familiar form of the Lennard-Jones function using
1248 > orientation-dependent $\sigma$ and $\epsilon$ parameters,
1249 > \begin{equation*}
1250 > V_{ij}({{\bf \hat u}_i}, {{\bf \hat u}_j}, {{\bf \hat
1251 > r}_{ij}}) = 4\epsilon ({{\bf \hat u}_i}, {{\bf \hat u}_j},
1252 > {{\bf \hat r}_{ij}})\left[\left(\frac{\sigma_0}{r_{ij}-\sigma({{\bf \hat u
1253 > }_i},
1254 > {{\bf \hat u}_j}, {{\bf \hat r}_{ij}})+\sigma_0}\right)^{12}
1255 > -\left(\frac{\sigma_0}{r_{ij}-\sigma({{\bf \hat u}_i}, {{\bf \hat u}_j},
1256 > {{\bf \hat r}_{ij}})+\sigma_0}\right)^6\right]
1257 > \label{eq:gb}
1258 > \end{equation*}
1259 >
1260 > The range $(\sigma({\bf \hat{u}}_{i},{\bf \hat{u}}_{j},{\bf
1261 > \hat{r}}_{ij}))$, and strength $(\epsilon({\bf \hat{u}}_{i},{\bf
1262 > \hat{u}}_{j},{\bf \hat{r}}_{ij}))$ parameters
1263 > are dependent on the relative orientations of the two ellipsoids (${\bf
1264 > \hat{u}}_{i},{\bf \hat{u}}_{j}$) as well as the direction of the
1265 > inter-ellipsoid separation (${\bf \hat{r}}_{ij}$).  The shape and
1266 > attractiveness of each ellipsoid is governed by a relatively small set
1267 > of parameters:
1268 > \begin{itemize}
1269 > \item  $d$:  range parameter for the side-by-side (S) and cross (X) configurations
1270 > \item  $l$:  range parameter for the end-to-end (E) configuration
1271 > \item  $\epsilon_X$:  well-depth parameter for the cross (X) configuration
1272 > \item  $\epsilon_S$:  well-depth parameter for the side-by-side (S) configuration
1273 > \item  $\epsilon_E$:  well depth parameter for the end-to-end (E) configuration
1274 > \item  $dw$: The ``softness'' of the potential
1275 > \end{itemize}
1276 > Additionally, there are two universal paramters to govern the overall
1277 > importance of the purely orientational ($\nu$) and the mixed
1278 > orientational / translational ($\mu$) parts of strength of the
1279 > interactions.  These parameters have default or ``canonical'' values,
1280 > but may be changed as a force field option:
1281 > \begin{itemize}
1282 >  \item $\nu$: purely orientational part : defaults to 1
1283 >  \item $\mu$: mixed orientational / translational part : defaults to
1284 >    2
1285 > \end{itemize}
1286 > Further details of the potential are given
1287 > elsewhere,\cite{Luckhurst:1990fy,Golubkov06,SunX._jp0762020} and an
1288 > excellent overview of the computational methods that can be used to
1289 > efficiently compute forces and torques for this potential can be found
1290 > in Ref. \citealp{Golubkov06}
1291 >
1292 > \begin{lstlisting}[caption={[An example of a GayBerneAtomTypes block.] A
1293 > simple example of a GayBerneAtomTypes block.  Distances ($d$ and $l$)
1294 > are given in \AA\ and energies ($\epsilon_X, \epsilon_S, \epsilon_E$)
1295 > are in units of kcal/mol. $dw$ is unitless.},
1296 > label={sch:GayBerneAtomTypes}]
1297 > begin GayBerneAtomTypes
1298 > //Name          d       l       eps_X           eps_S           eps_E     dw
1299 > GBlinear        2.8104  9.993   0.774729        0.774729        0.116839  1.0
1300 > GBC6H6          4.65    2.03    0.540           0.540           1.9818    0.6
1301 > GBCH3OH         2.55    3.18    0.542           0.542           0.55826   1.0
1302 > end GayBerneAtomTypes                  
1303 > \end{lstlisting}
1304 >
1305 > \subsubsection{\label{section:ffSticky}The StickyAtomTypes block}
1306 >
1307 > One of the solvents that can be simulated by {\sc OpenMD} is the
1308 > extended Soft Sticky Dipole (SSD/E) water model.\cite{fennell04} The
1309 > original SSD was developed by Ichiye \emph{et
1310 >  al.}\cite{liu96:new_model} as a modified form of the hard-sphere
1311   water model proposed by Bratko, Blum, and
1312   Luzar.\cite{Bratko85,Bratko95} It consists of a single point dipole
1313   with a Lennard-Jones core and a sticky potential that directs the
# Line 1155 | Line 1373 | HOH angle in each water molecule. }
1373   \label{fig:ssd}
1374   \end{figure}
1375  
1158
1376   Since SSD/E is a single-point {\it dipolar} model, the force
1377   calculations are simplified significantly relative to the standard
1378   {\it charged} multi-point models. In the original Monte Carlo
# Line 1174 | Line 1391 | SSD model that led to lower than expected densities at
1391  
1392   Recent constant pressure simulations revealed issues in the original
1393   SSD model that led to lower than expected densities at all target
1394 < pressures.\cite{Ichiye03,fennell04} The default model in {\sc OpenMD}
1395 < is therefore SSD/E, a density corrected derivative of SSD that
1396 < exhibits improved liquid structure and transport behavior. If the use
1397 < of a reaction field long-range interaction correction is desired, it
1398 < is recommended that the parameters be modified to those of the SSD/RF
1399 < model (an SSD variant parameterized for reaction field). These solvent
1183 < parameters are listed and can be easily modified in the {\sc duff}
1184 < force field file ({\tt DUFF.frc}).  A table of the parameter values
1185 < and the drawbacks and benefits of the different density corrected SSD
1186 < models can be found in reference~\cite{fennell04}.
1394 > pressures,\cite{Ichiye03,fennell04} so variants on the sticky
1395 > potential can be specified by using one of a number of substitute atom
1396 > types (see listing \ref{sch:StickyAtomTypes}).  A table of the
1397 > parameter values and the drawbacks and benefits of the different
1398 > density corrected SSD models can be found in
1399 > reference~\citealp{fennell04}.
1400  
1401 < \section{\label{section:WATER}The {\sc water} Force Field}
1401 > \begin{lstlisting}[caption={[An example of a StickyAtomTypes block.] A
1402 > simple example of a StickyAtomTypes block.  Distances ($r_l$, $r_u$,
1403 > $r_{l}'$ and $r_{u}'$) are given in \AA\ and energies ($v_0, v_{0}'$)
1404 > are in units of kcal/mol. $w_0$ is unitless.},
1405 > label={sch:StickyAtomTypes}]
1406 > begin StickyAtomTypes
1407 > //name  w0      v0 (kcal/mol)   v0p     rl (Ang)  ru    rlp     rup
1408 > SSD_E   0.07715 3.90            3.90    2.40      3.80  2.75    3.35
1409 > SSD_RF  0.07715 3.90            3.90    2.40      3.80  2.75    3.35
1410 > SSD     0.07715 3.7284          3.7284  2.75      3.35  2.75    4.0
1411 > SSD1    0.07715 3.6613          3.6613  2.75      3.35  2.75    4.0
1412 > end StickyAtomTypes
1413 > \end{lstlisting}
1414  
1415 < 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.
1415 > \subsection{\label{section::ffMetals}Metallic Atom Types}
1416  
1417 < \section{\label{section:eam}Embedded Atom Method}
1418 <
1419 < {\sc OpenMD} implements a potential that describes bonding in
1420 < transition metal
1421 < systems.~\cite{Finnis84,Ercolessi88,Chen90,Qi99,Ercolessi02} This
1422 < potential has an attractive interaction which models ``Embedding'' a
1423 < positively charged pseudo-atom core in the electron density due to the
1424 < free valance ``sea'' of electrons created by the surrounding atoms in
1221 < the system.  A pairwise part of the potential (which is primarily
1222 < repulsive) describes the interaction of the positively charged metal
1223 < core ions with one another.  The Embedded Atom Method ({\sc
1224 < eam})~\cite{Daw84,FBD86,johnson89,Lu97} has been widely adopted in the
1225 < materials science community and has been included in {\sc OpenMD}. A
1226 < good review of {\sc eam} and other formulations of metallic potentials
1227 < was given by Voter.\cite{Voter:95}
1228 <
1229 < The {\sc eam} potential has the form:
1417 > {\sc OpenMD} implements a number of related potentials that describe
1418 > bonding in transition metals. These potentials have an attractive
1419 > interaction which models ``Embedding'' a positively charged
1420 > pseudo-atom core in the electron density due to the free valance
1421 > ``sea'' of electrons created by the surrounding atoms in the system.
1422 > A pairwise part of the potential (which is primarily repulsive)
1423 > describes the interaction of the positively charged metal core ions
1424 > with one another.  These potentials have the form:
1425   \begin{equation}
1426   V  =  \sum_{i} F_{i}\left[\rho_{i}\right] + \sum_{i} \sum_{j \neq i}
1427   \phi_{ij}({\bf r}_{ij})
# Line 1243 | Line 1438 | to compute the inter-atomic forces.
1438   transition metal potentials require two loops through the atom pairs
1439   to compute the inter-atomic forces.
1440  
1441 < The pairwise portion of the potential, $\phi_{ij}$, is a primarily
1442 < repulsive interaction between atoms $i$ and $j$. In the original
1248 < formulation of {\sc eam}\cite{Daw84}, $\phi_{ij}$ was an entirely
1249 < repulsive term; however later refinements to {\sc eam} allowed for
1250 < more general forms for $\phi$.\cite{Daw89} The effective cutoff
1251 < distance, $r_{{\text cut}}$ is the distance at which the values of
1252 < $f(r)$ and $\phi(r)$ drop to zero for all atoms present in the
1253 < simulation.  In practice, this distance is fairly small, limiting the
1254 < summations in the {\sc eam} equation to the few dozen atoms
1255 < surrounding atom $i$ for both the density $\rho$ and pairwise $\phi$
1256 < interactions.
1441 > The pairwise portion of the potential, $\phi_{ij}$, is usually a
1442 > repulsive interaction between atoms $i$ and $j$.
1443  
1444 < In computing forces for alloys, mixing rules as outlined by
1445 < Johnson~\cite{johnson89} are used to compute the heterogenous pair
1446 < potential,
1444 > \subsubsection{\label{section:ffEAM}The EAMAtomTypes block}
1445 > The Embedded Atom Method ({\sc eam}) is one of the most widely-used
1446 > potentials for transition
1447 > metals.~\cite{Finnis84,Ercolessi88,Chen90,Qi99,Ercolessi02,Daw84,FBD86,johnson89,Lu97}
1448 > It has been widely adopted in the materials science community and a
1449 > good review of {\sc eam} and other formulations of metallic potentials
1450 > was given by Voter.\cite{Voter:95}
1451 >
1452 > In the original formulation of {\sc eam}\cite{Daw84}, the pair
1453 > potential, $\phi_{ij}$ was an entirely repulsive term; however later
1454 > refinements to {\sc eam} allowed for more general forms for
1455 > $\phi$.\cite{Daw89} The effective cutoff distance, $r_{{\text cut}}$
1456 > is the distance at which the values of $f(r)$ and $\phi(r)$ drop to
1457 > zero for all atoms present in the simulation.  In practice, this
1458 > distance is fairly small, limiting the summations in the {\sc eam}
1459 > equation to the few dozen atoms surrounding atom $i$ for both the
1460 > density $\rho$ and pairwise $\phi$ interactions.
1461 >
1462 > In computing forces for alloys, OpenMD uses mixing rules outlined by
1463 > Johnson~\cite{johnson89} to compute the heterogenous pair potential,
1464   \begin{equation}
1465   \label{eq:johnson}
1466   \phi_{ab}(r)=\frac{1}{2}\left(
# Line 1289 | Line 1492 | files.  
1492   $\mbox{kcal mol}^{-1}$ as in the rest of the {\sc OpenMD} force field
1493   files.  
1494  
1495 < \section{\label{section:sc}The Sutton-Chen Force Field}
1495 > \begin{lstlisting}[caption={[An example of a EAMAtomTypes block.] A
1496 > simple example of a EAMAtomTypes block. Here the only data provided is
1497 > the name of a {\tt funcfl} file which contains the raw data for spline
1498 > interpolations for the density, functional, and pair potential.},
1499 > label={sch:EAMAtomTypes}]
1500 > begin EAMAtomTypes
1501 > Au      Au.u3.funcfl
1502 > Ag      Ag.u3.funcfl
1503 > Cu      Cu.u3.funcfl
1504 > Ni      Ni.u3.funcfl
1505 > Pd      Pd.u3.funcfl
1506 > Pt      Pt.u3.funcfl
1507 > end EAMAtomTypes
1508 > \end{lstlisting}
1509  
1510 +
1511 + \subsubsection{\label{section:ffSC}The SuttonChenAtomTypes block}
1512 +
1513   The Sutton-Chen ({\sc sc})~\cite{Chen90} potential has been used to
1514 < study a wide range of phenomena in metals.  Although it is similar in
1515 < form to the {\sc eam} potential, the Sutton-Chen model takes on a
1516 < simpler form,
1514 > study a wide range of phenomena in metals.  Although it has the same
1515 > basic form as the {\sc eam} potential, the Sutton-Chen model takes on
1516 > a simpler form,
1517   \begin{equation}
1518   \label{eq:SCP1}
1519   U_{tot}=\sum _{i}\left[ \frac{1}{2}\sum _{j\neq
1520 < i}D_{ij}V^{pair}_{ij}(r_{ij})-c_{i}D_{ii}\sqrt{\rho_{i}}\right] ,
1520 > i}\epsilon_{ij}V^{pair}_{ij}(r_{ij})-c_{i}\epsilon_{ii}\sqrt{\rho_{i}}\right] ,
1521   \end{equation}
1522   where $V^{pair}_{ij}$ and $\rho_{i}$ are given by
1523   \begin{equation}
1524   \label{eq:SCP2}
1525   V^{pair}_{ij}(r)=\left(
1526 < \frac{\alpha_{ij}}{r_{ij}}\right)^{n_{ij}}, \rho_{i}=\sum_{j\neq i}\left(
1526 > \frac{\alpha_{ij}}{r_{ij}}\right)^{n_{ij}} \hspace{1in} \rho_{i}=\sum_{j\neq i}\left(
1527   \frac{\alpha_{ij}}{r_{ij}}\right) ^{m_{ij}}
1528   \end{equation}
1529  
# Line 1312 | Line 1531 | the interactions between the valence electrons and the
1531   interactions of the pseudo-atom cores.  The $\sqrt{\rho_i}$ term in
1532   Eq. (\ref{eq:SCP1}) is an attractive many-body potential that models
1533   the interactions between the valence electrons and the cores of the
1534 < pseudo-atoms.  $D_{ij}$, $D_{ii}$, $c_i$ and $\alpha_{ij}$ are
1535 < parameters used to tune the potential for different transition
1536 < metals.
1534 > pseudo-atoms.  $\epsilon_{ij}$, $\epsilon_{ii}$, $c_i$ and
1535 > $\alpha_{ij}$ are parameters used to tune the potential for different
1536 > transition metals.
1537  
1538   The {\sc sc} potential form has also been parameterized by Qi {\it et
1539   al.}\cite{Qi99} These parameters were obtained via empirical and {\it
1540   ab initio} calculations to match structural features of the FCC
1541 < crystal.  To specify the original Sutton-Chen variant of the {\sc sc}
1542 < force field, the user would add the {\tt forceFieldVariant = "SC";}
1543 < line to the meta-data file, while specification of the Qi {\it et al.}
1544 < quantum-adapted variant of the {\sc sc} potential, the user would add
1545 < the {\tt forceFieldVariant = "QSC";} line to the meta-data file.
1541 > crystal.  Interested readers are encouraged to consult reference
1542 > \citealp{Qi99} for further details.
1543 >
1544 > \begin{lstlisting}[caption={[An example of a SCAtomTypes block.] A
1545 > simple example of a SCAtomTypes block.  Distances ($\alpha$)
1546 > are given in \AA\ and energies ($\epsilon$) are (by convention) given in
1547 > units of eV.  These units must be specified in the {\tt Options} block
1548 > using the keyword {\tt MetallicEnergyUnitScaling}.  Without this {\tt
1549 > Options} keyword, the default units for $\epsilon$ are kcal/mol.  The
1550 > other parameters, $m$, $n$, and $c$ are unitless.},
1551 > label={sch:SCAtomTypes}]
1552 > begin SCAtomTypes
1553 > // Name  epsilon(eV)      c      m       n      alpha(angstroms)
1554 > Ni      0.0073767       84.745  5.0     10.0    3.5157
1555 > Cu      0.0057921       84.843  5.0     10.0    3.6030
1556 > Rh      0.0024612       305.499 5.0     13.0    3.7984
1557 > Pd      0.0032864       148.205 6.0     12.0    3.8813
1558 > Ag      0.0039450       96.524  6.0     11.0    4.0691
1559 > Ir      0.0037674       224.815 6.0     13.0    3.8344  
1560 > Pt      0.0097894       71.336  7.0     11.0    3.9163
1561 > Au      0.0078052       53.581  8.0     11.0    4.0651
1562 > Au2     0.0078052       53.581  8.0     11.0    4.0651
1563 > end SCAtomTypes
1564 > \end{lstlisting}
1565 >
1566 > \subsection{\label{section::ffShortRange}Short Range Interactions}
1567 > \subsubsection{\label{section:ffBond}The BondTypes block}
1568 > \subsubsection{\label{section:ffBend}The BendTypes block}
1569 > A harmonic bend potential is represented by the following function:
1570 > \begin{equation}
1571 > V_{\text{bend}}(\theta_{ijk}) = k_{\theta}( \theta_{ijk} - \theta_0
1572 > )^2, \label{eq:bendPot}
1573 > \end{equation}
1574 > where $\theta_{ijk}$ is the angle defined by atoms $i$, $j$, and $k$,
1575 > $\theta_0$ is the equilibrium bond angle, and $k_{\theta}$ is the
1576 > force constant which determines the strength of the harmonic bend.
1577 >
1578 > \subsubsection{\label{section:ffTorsion}The TorsionTypes block}
1579 > The torsion potential is often represented as a cosine series of the
1580 > form:
1581 > \begin{equation}
1582 > V_{\text{torsion}}(\phi) = c_1[1 + \cos \phi]
1583 >        + c_2[1 + \cos(2\phi)]
1584 >        + c_3[1 + \cos(3\phi)],
1585 > \label{eq:origTorsionPot}
1586 > \end{equation}
1587 > where:
1588 > \begin{equation}
1589 > \cos\phi = (\hat{\mathbf{r}}_{ij} \times \hat{\mathbf{r}}_{jk}) \cdot
1590 >        (\hat{\mathbf{r}}_{jk} \times \hat{\mathbf{r}}_{kl}).
1591 > \label{eq:torsPhi}
1592 > \end{equation}
1593 > Here, $\hat{\mathbf{r}}_{\alpha\beta}$ are the set of unit bond
1594 > vectors between atoms $i$, $j$, $k$, and $l$. For computational
1595 > efficiency, the torsion potential has been recast after the method of
1596 > {\sc charmm},\cite{Brooks83} in which the angle series is converted to
1597 > a power series of the form:
1598 > \begin{equation}
1599 > V_{\text{torsion}}(\phi) =  
1600 >        k_3 \cos^3 \phi + k_2 \cos^2 \phi + k_1 \cos \phi + k_0,
1601 > \label{eq:torsionPot}
1602 > \end{equation}
1603 > where:
1604 > \begin{align*}
1605 > k_0 &= c_1 + c_3, \\
1606 > k_1 &= c_1 - 3c_3, \\
1607 > k_2 &= 2 c_2, \\
1608 > k_3 &= 4c_3.
1609 > \end{align*}
1610 > By recasting the potential as a power series, repeated trigonometric
1611 > evaluations are avoided during the calculation of the potential
1612 > energy.
1613  
1614 + \subsubsection{\label{section:ffInversion}The InversionTypes block}
1615 + \subsection{\label{section::ffLongRange}Long Range Interactions}
1616 + \subsubsection{\label{section:ffNBinteraction}The NonBondedInteraction block}
1617 +
1618 +
1619 +
1620 + (see Fig.~\ref{fig:lipidModel}), The parameters for $k_{\theta}$ and
1621 + $\theta_0$ are borrowed from those in TraPPE.\cite{Siepmann1998}
1622 +
1623 + Calculating the long-range (non-bonded) potential involves a sum over
1624 + all pairs of atoms (except for those atoms which are involved in a
1625 + bond, bend, or torsion with each other).  If done poorly, calculating
1626 + the the long-range interactions for $N$ atoms would involve $N(N-1)/2$
1627 + evaluations of atomic distances.  To reduce the number of distance
1628 + evaluations between pairs of atoms, {\sc OpenMD} allows the use of
1629 + switched cutoffs with Verlet neighbor lists.\cite{Allen87} Neutral
1630 + groups which contain charges will exhibit pathological forces unless
1631 + the cutoff is applied to the neutral groups evenly instead of to the
1632 + individual atoms.\cite{leach01:mm}  {\sc OpenMD} allows users to
1633 + specify cutoff groups which may contain an arbitrary number of atoms
1634 + in the molecule.  Atoms in a cutoff group are treated as a single unit
1635 + for the evaluation of the switching function:
1636 + \begin{equation}
1637 + V_{\mathrm{long-range}} = \sum_{a} \sum_{b>a} s(r_{ab}) \sum_{i \in a} \sum_{j \in b} V_{ij}(r_{ij}),
1638 + \end{equation}
1639 + where $r_{ab}$ is the distance between the centers of mass of the two
1640 + cutoff groups ($a$ and $b$).
1641 +
1642 + The sums over $a$ and $b$ are over the cutoff groups that are present
1643 + in the simulation.  Atoms which are not explicitly defined as members
1644 + of a {\tt cutoffGroup} are treated as a group consisting of only one
1645 + atom.  The switching function, $s(r)$ is the standard cubic switching
1646 + function,
1647 + \begin{equation}
1648 + S(r) =
1649 +        \begin{cases}
1650 +        1 & \text{if $r \le r_{\text{sw}}$},\\
1651 +        \frac{(r_{\text{cut}} + 2r - 3r_{\text{sw}})(r_{\text{cut}} - r)^2}
1652 +        {(r_{\text{cut}} - r_{\text{sw}})^3}
1653 +        & \text{if $r_{\text{sw}} < r \le r_{\text{cut}}$}, \\
1654 +        0 & \text{if $r > r_{\text{cut}}$.}
1655 +        \end{cases}
1656 + \label{eq:dipoleSwitching}
1657 + \end{equation}
1658 + Here, $r_{\text{sw}}$ is the {\tt switchingRadius}, or the distance
1659 + beyond which interactions are reduced, and $r_{\text{cut}}$ is the
1660 + {\tt cutoffRadius}, or the distance at which interactions are
1661 + truncated.
1662 +
1663 + Users of {\sc OpenMD} do not need to specify the {\tt cutoffRadius} or
1664 + {\tt switchingRadius}.  In simulations containing only Lennard-Jones
1665 + atoms, the cutoff radius has a default value of $2.5\sigma_{ii}$,
1666 + where $\sigma_{ii}$ is the largest Lennard-Jones length parameter
1667 + present in the simulation.  In simulations containing charged or
1668 + dipolar atoms, the default cutoff radius is $15 \mbox{\AA}$.  
1669 +
1670 + The {\tt switchingRadius} is set to a default value of 95\% of the
1671 + {\tt cutoffRadius}.  In the special case of a simulation containing
1672 + {\it only} Lennard-Jones atoms, the default switching radius takes the
1673 + same value as the cutoff radius, and {\sc OpenMD} will use a shifted
1674 + potential to remove discontinuities in the potential at the cutoff.
1675 + Both radii may be specified in the meta-data file.
1676 +
1677 + Force fields can be added to {\sc OpenMD}, although it comes with a few
1678 + simple examples (Lennard-Jones, {\sc duff}, {\sc water}, and {\sc
1679 + eam}) which are explained in the following sections.
1680 +
1681 + \section{\label{sec:LJPot}The Lennard Jones Force Field}
1682 +
1683 + Scheme
1684 + \ref{sch:LJFF} gives an example meta-data file that
1685 + sets up a system of 108 Ar particles to be simulated using the
1686 + Lennard-Jones force field.
1687 +
1688 + \begin{lstlisting}[float,caption={[Invocation of the Lennard-Jones
1689 + force field] A sample startup file for a small Lennard-Jones
1690 + simulation.},label={sch:LJFF}]
1691 + <OpenMD>
1692 +  <MetaData>
1693 + #include "argon.md"
1694 +
1695 + component{
1696 +  type = "Ar";
1697 +  nMol = 108;
1698 + }
1699 +
1700 + forceField = "LJ";
1701 +  </MetaData>
1702 +  <Snapshot>     // not shown in this scheme
1703 +  </Snapshot>
1704 + </OpenMD>
1705 + \end{lstlisting}
1706 +
1707 +
1708 + \section{\label{section:DUFF}Dipolar Unified-Atom Force Field}
1709 +
1710 + The dipolar unified-atom force field ({\sc duff}) was developed to
1711 + simulate lipid bilayers. These types of simulations require a model
1712 + capable of forming bilayers, while still being sufficiently
1713 + computationally efficient to allow large systems ($\sim$100's of
1714 + phospholipids, $\sim$1000's of waters) to be simulated for long times
1715 + ($\sim$10's of nanoseconds). With this goal in mind, {\sc duff} has no
1716 + point charges. Charge-neutral distributions are replaced with dipoles,
1717 + while most atoms and groups of atoms are reduced to Lennard-Jones
1718 + interaction sites. This simplification reduces the length scale of
1719 + long range interactions from $\frac{1}{r}$ to $\frac{1}{r^3}$,
1720 + removing the need for the computationally expensive Ewald
1721 + sum. Instead, Verlet neighbor-lists and cutoff radii are used for the
1722 + dipolar interactions, and, if desired, a reaction field may be added
1723 + to mimic longer range interactions.
1724 +
1725 + As an example, lipid head-groups in {\sc duff} are represented as
1726 + point dipole interaction sites.  Placing a dipole at the head group's
1727 + center of mass mimics the charge separation found in common
1728 + phospholipid head groups such as phosphatidylcholine.\cite{Cevc87}
1729 + Additionally, a large Lennard-Jones site is located at the
1730 + pseudoatom's center of mass. The model is illustrated by the red atom
1731 + in Fig.~\ref{fig:lipidModel}. The water model we use to
1732 + complement the dipoles of the lipids is a
1733 + reparameterization\cite{fennell04} of the soft sticky dipole (SSD)
1734 + model of Ichiye
1735 + \emph{et al.}\cite{liu96:new_model}
1736 +
1737 + \begin{figure}
1738 + \centering
1739 + \includegraphics[width=\linewidth]{lipidModel.pdf}
1740 + \caption[A representation of a lipid model in {\sc duff}]{A
1741 + representation of the lipid model. $\phi$ is the torsion angle,
1742 + $\theta$ is the bend angle, and $\mu$ is the dipole moment of the head
1743 + group.}
1744 + \label{fig:lipidModel}
1745 + \end{figure}
1746 +
1747 + A set of scalable parameters has been used to model the alkyl groups
1748 + with Lennard-Jones sites. For this, parameters from the TraPPE force
1749 + field of Siepmann \emph{et al.}\cite{Siepmann1998} have been
1750 + utilized. TraPPE is a unified-atom representation of n-alkanes which
1751 + is parametrized against phase equilibria using Gibbs ensemble Monte
1752 + Carlo simulation techniques.\cite{Siepmann1998} One of the advantages
1753 + of TraPPE is that it generalizes the types of atoms in an alkyl chain
1754 + to keep the number of pseudoatoms to a minimum; thus, the parameters
1755 + for a unified atom such as $\text{CH}_2$ do not change depending on
1756 + what species are bonded to it.
1757 +
1758 + As is required by TraPPE, {\sc duff} also constrains all bonds to be
1759 + of fixed length. Typically, bond vibrations are the fastest motions in
1760 + a molecular dynamic simulation.  With these vibrations present, small
1761 + time steps between force evaluations must be used to ensure adequate
1762 + energy conservation in the bond degrees of freedom. By constraining
1763 + the bond lengths, larger time steps may be used when integrating the
1764 + equations of motion. A simulation using {\sc duff} is illustrated in
1765 + Scheme \ref{sch:DUFF}.
1766 +
1767 + \begin{lstlisting}[float,caption={[Invocation of {\sc duff}]A portion
1768 + of a startup file showing a simulation utilizing {\sc
1769 + duff}},label={sch:DUFF}]  
1770 + <OpenMD>
1771 +  <MetaData>
1772 + #include "water.md"
1773 + #include "lipid.md"
1774 +
1775 + component{
1776 +  type = "simpleLipid_16";
1777 +  nMol = 60;
1778 + }
1779 +
1780 + component{
1781 +  type = "SSD_water";
1782 +  nMol = 1936;
1783 + }
1784 +
1785 + forceField = "DUFF";
1786 +  </MetaData>
1787 +  <Snapshot>     // not shown in this scheme
1788 +  </Snapshot>
1789 + </OpenMD>
1790 + \end{lstlisting}
1791 +
1792 +
1793 +
1794 + The cross potential between molecules $I$ and $J$,
1795 + $V^{IJ}_{\text{Cross}}$, is as follows:
1796 + \begin{equation}
1797 + V^{IJ}_{\text{Cross}} =
1798 +        \sum_{i \in I} \sum_{j \in J}
1799 +        \biggl[ V_{\text{LJ}}(r_{ij}) +  V_{\text{dipole}}
1800 +        (\mathbf{r}_{ij},\boldsymbol{\Omega}_{i},\boldsymbol{\Omega}_{j})
1801 +        + V_{\text{sticky}}
1802 +        (\mathbf{r}_{ij},\boldsymbol{\Omega}_{i},\boldsymbol{\Omega}_{j})
1803 +        \biggr],
1804 + \label{eq:crossPotentail}
1805 + \end{equation}
1806 + where $V_{\text{LJ}}$ is the Lennard Jones potential,
1807 + $V_{\text{dipole}}$ is the dipole dipole potential, and
1808 + $V_{\text{sticky}}$ is the sticky potential defined by the SSD model
1809 + (Sec.~\ref{section:SSD}). Note that not all atom types include all
1810 + interactions.
1811 +
1812 +
1813 + \section{\label{section:WATER}The {\sc water} Force Field}
1814 +
1815 + In addition to the {\sc duff} force field's solvent description, a
1816 + separate {\sc water} force field has been included for simulating most
1817 + of the common rigid-body water models. This force field includes the
1818 + simple and point-dipolar models (SSD, SSD1, SSD/E, SSD/RF, and DPD
1819 + water), as well as the common charge-based models (SPC, SPC/E, TIP3P,
1820 + TIP4P, and
1821 + TIP5P).\cite{liu96:new_model,Ichiye03,fennell04,Marrink01,Berendsen81,Berendsen87,Jorgensen83,Mahoney00}
1822 + In order to handle these models, charge-charge interactions were
1823 + included in the force-loop:
1824 + \begin{equation}
1825 + V_{\text{charge}}(r_{ij}) = \sum_{ij}\frac{q_iq_je^2}{r_{ij}},
1826 + \end{equation}
1827 + where $q$ represents the charge on particle $i$ or $j$, and $e$ is the
1828 + charge of an electron in Coulombs. The charge-charge interaction
1829 + support is rudimentary in the current version of {\sc OpenMD}.  As with
1830 + the other pair interactions, charges can be simulated with a pure
1831 + cutoff or a reaction field.  The various methods for performing the
1832 + Ewald summation have not yet been included.  The {\sc water} force
1833 + field can be easily expanded through modification of the {\sc water}
1834 + force field file ({\tt WATER.frc}). By adding atom types and inserting
1835 + the appropriate parameters, it is possible to extend the force field
1836 + to handle rigid molecules other than water.
1837 +
1838 +
1839 + \section{\label{section:sc}The Sutton-Chen Force Field}
1840 +
1841 +
1842   \section{\label{section:clay}The CLAY force field}
1843  
1844   The {\sc clay} force field is based on an ionic (nonbonded)
# Line 2621 | Line 3135 | tensor.
3135  
3136   \section{Constant Pressure without periodic boundary conditions (The LangevinHull)}
3137  
3138 < The Langevin Hull uses an external bath at a fixed constant pressure
3138 > The Langevin Hull\cite{Vardeman2011} uses an external bath at a fixed constant pressure
3139   ($P$) and temperature ($T$) with an effective solvent viscosity
3140   ($\eta$).  This bath interacts only with the objects on the exterior
3141   hull of the system.  Defining the hull of the atoms in a simulation is
# Line 2933 | Line 3447 | Harmonic Forces are used by default
3447   \label{table:zconParams}
3448   \end{longtable}
3449  
3450 < \chapter{\label{section:restraints}Restraints}
3451 < Restraints are external potentials that are added to a system to keep
3452 < particular molecules or collections of particles close to some
3453 < reference structure.  A restraint can be a collective
3450 > % \chapter{\label{section:restraints}Restraints}
3451 > % Restraints are external potentials that are added to a system to
3452 > % keep particular molecules or collections of particles close to some
3453 > % reference structure.  A restraint can be a collective
3454  
3455   \chapter{\label{section:thermInt}Thermodynamic Integration}
3456  
# Line 3075 | Line 3589 | Einstein crystal
3589   Einstein crystal
3590   \label{table:thermIntParams}
3591   \end{longtable}
3592 +
3593 + \chapter{\label{section:rnemd}Reverse Non-Equilibrium Molecular Dynamics (RNEMD)}
3594 +
3595 + There are many ways to compute transport properties from molecular
3596 + dynamics simulations.  Equilibrium Molecular Dynamics (EMD)
3597 + simulations can be used by computing relevant time correlation
3598 + functions and assuming linear response theory holds.  For some transport properties (notably thermal conductivity), EMD approaches
3599 + are subject to noise and poor convergence of the relevant
3600 + correlation functions. Traditional Non-equilibrium Molecular Dynamics
3601 + (NEMD) methods impose a gradient (e.g. thermal or momentum) on a
3602 + simulation.  However, the resulting flux is often difficult to
3603 + measure. Furthermore, problems arise for NEMD simulations of
3604 + heterogeneous systems, such as phase-phase boundaries or interfaces,
3605 + where the type of gradient to enforce at the boundary between
3606 + materials is unclear.
3607 +
3608 + {\it Reverse} Non-Equilibrium Molecular Dynamics (RNEMD) methods adopt
3609 + a different approach in that an unphysical {\it flux} is imposed
3610 + between different regions or ``slabs'' of the simulation box.  The
3611 + response of the system is to develop a temperature or momentum {\it
3612 +  gradient} between the two regions. Since the amount of the applied
3613 + flux is known exactly, and the measurement of gradient is generally
3614 + less complicated, imposed-flux methods typically take shorter
3615 + simulation times to obtain converged results for transport properties.
3616 +
3617 + \begin{figure}
3618 + \includegraphics[width=\linewidth]{rnemdDemo}
3619 + \caption{The (VSS) RNEMD approach imposes unphysical transfer of both
3620 +  linear momentum and kinetic energy between a ``hot'' slab and a
3621 +  ``cold'' slab in the simulation box.  The system responds to this
3622 +  imposed flux by generating both momentum and temperature gradients.
3623 +  The slope of the gradients can then be used to compute transport
3624 +  properties (e.g. shear viscosity and thermal conductivity).}
3625 + \label{rnemdDemo}
3626 + \end{figure}
3627 +
3628 + \section{\label{section:algo}Three algorithms for carrying out RNEMD simulations}
3629 + \subsection{\label{subsection:swapping}The swapping algorithm}
3630 + The original ``swapping'' approaches by M\"{u}ller-Plathe {\it et
3631 +  al.}\cite{ISI:000080382700030,MullerPlathe:1997xw} can be understood
3632 + as a sequence of imaginary elastic collisions between particles in
3633 + opposite slabs.  In each collision, the entire momentum vectors of
3634 + both particles may be exchanged to generate a thermal
3635 + flux. Alternatively, a single component of the momentum vectors may be
3636 + exchanged to generate a shear flux.  This algorithm turns out to be
3637 + quite useful in many simulations. However, the M\"{u}ller-Plathe
3638 + swapping approach perturbs the system away from ideal
3639 + Maxwell-Boltzmann distributions, and this may leads to undesirable
3640 + side-effects when the applied flux becomes large.\cite{Maginn:2010}
3641 + This limits the applicability of the swapping algorithm, so in OpenMD,
3642 + we have implemented two additional algorithms for RNEMD in addition to the
3643 + original swapping approach.
3644 +
3645 + \subsection{\label{subsection:nivs}Non-Isotropic Velocity Scaling (NIVS)}
3646 + Instead of having momentum exchange between {\it individual particles}
3647 + in each slab, the NIVS algorithm applies velocity scaling to all of
3648 + the selected particles in both slabs.\cite{kuang:164101} A combination of linear
3649 + momentum, kinetic energy, and flux constraint equations governs the
3650 + amount of velocity scaling performed at each step. Interested readers
3651 + should consult ref. \citealp{kuang:164101} for further details on the
3652 + methodology.
3653 +
3654 + NIVS has been shown to be very effective at producing thermal
3655 + gradients and for computing thermal conductivities, particularly for
3656 + heterogeneous interfaces.  Although the NIVS algorithm can also be
3657 + applied to impose a directional momentum flux, thermal anisotropy was
3658 + observed in relatively high flux simulations, and the method is not
3659 + suitable for imposing a shear flux or for computing shear viscosities.
3660 +
3661 + \subsection{\label{subsection:vss}Velocity Shearing and Scaling (VSS)}
3662 + The third RNEMD algorithm implemented in OpenMD utilizes a series of
3663 + simultaneous velocity shearing and scaling exchanges between the two
3664 + slabs.\cite{2012MolPh.110..691K}  This method results in a set of simpler equations to satisfy
3665 + the conservation constraints while creating a desired flux between the
3666 + two slabs.
3667 +
3668 + The VSS approach is versatile in that it may be used to implement both
3669 + thermal and shear transport either separately or simultaneously.
3670 + Perturbations of velocities away from the ideal Maxwell-Boltzmann
3671 + distributions are minimal, and thermal anisotropy is kept to a
3672 + minimum.  This ability to generate simultaneous thermal and shear
3673 + fluxes has been utilized to map out the shear viscosity of SPC/E water
3674 + over a wide range of temperatures (90~K) just with a single simulation.
3675 + VSS-RNEMD also allows the directional momentum flux to have
3676 + arbitary directions, which could aid in the study of anisotropic solid
3677 + surfaces in contact with liquid environments.
3678 +
3679 + \section{\label{section:usingRNEMD}Using OpenMD to perform a RNEMD simulation}
3680 + \subsection{\label{section:rnemdParams} What the user needs to specify}
3681 + To carry out a RNEMD simulation,
3682 + a user must specify a number of parameters in the MetaData (.md) file.
3683 + Because the RNEMD methods have a large number of parameters, these
3684 + must be enclosed in a {\it separate} {\tt RNEMD\{...\}} block.  The most important
3685 + parameters to specify are the {\tt useRNEMD}, {\tt fluxType} and flux
3686 + parameters. Most other parameters (summarized in table
3687 + \ref{table:rnemd}) have reasonable default values.  {\tt fluxType}
3688 + sets up the kind of exchange that will be carried out between the two
3689 + slabs (either Kinetic Energy ({\tt KE}) or momentum ({\tt Px, Py, Pz,
3690 +  Pvector}), or some combination of these).  The flux is specified
3691 + with the use of three possible parameters: {\tt kineticFlux} for
3692 + kinetic energy exchange, as well as {\tt momentumFlux} or {\tt
3693 +  momentumFluxVector} for simulations with directional exchange.
3694 +
3695 + \subsection{\label{section:rnemdResults} Processing the results}
3696 + OpenMD will generate a {\tt .rnemd}
3697 + file with the same prefix as the original {\tt .md} file.  This file
3698 + contains a running average of properties of interest computed within a
3699 + set of bins that divide the simulation cell along the $z$-axis.  The
3700 + first column of the {\tt .rnemd} file is the $z$ coordinate of the
3701 + center of each bin, while following columns may contain the average
3702 + temperature, velocity, or density within each bin.  The output format
3703 + in the {\tt .rnemd} file can be altered with the {\tt outputFields},
3704 + {\tt outputBins}, and {\tt outputFileName} parameters.  A report at
3705 + the top of the {\tt .rnemd} file contains the current exchange totals
3706 + as well as the average flux applied during the simulation.  Using the
3707 + slope of the temperature or velocity gradient obtaine from the {\tt
3708 +  .rnemd} file along with the applied flux, the user can very simply
3709 + arrive at estimates of thermal conductivities ($\lambda$),
3710 + \begin{equation}
3711 + J_z = -\lambda \frac{\partial T}{\partial z},
3712 + \end{equation}
3713 + and shear viscosities ($\eta$),
3714 + \begin{equation}
3715 + j_z(p_x) = -\eta \frac{\partial \langle v_x \rangle}{\partial z}.
3716 + \end{equation}
3717 + Here, the quantities on the left hand side are the actual flux values
3718 + (in the header of the {\tt .rnemd} file), while the slopes are
3719 + obtained from linear fits to the gradients observed in the {\tt
3720 +  .rnemd} file.
3721  
3722 + More complicated simulations (including interfaces) require a bit more
3723 + care.  Here the second derivative may be required to compute the
3724 + interfacial thermal conductance,
3725 + \begin{align}
3726 +  G^\prime &= \left(\nabla\lambda \cdot \mathbf{\hat{n}}\right)_{z_0} \\
3727 +  &= \frac{\partial}{\partial z}\left(-\frac{J_z}{
3728 +      \left(\frac{\partial T}{\partial z}\right)}\right)_{z_0} \\
3729 +  &= J_z\left(\frac{\partial^2 T}{\partial z^2}\right)_{z_0} \Big/
3730 +  \left(\frac{\partial T}{\partial z}\right)_{z_0}^2.
3731 +  \label{derivativeG}
3732 + \end{align}
3733 + where $z_0$ is the location of the interface between two materials and
3734 + $\mathbf{\hat{n}}$ is a unit vector normal to the interface.  We
3735 + suggest that users interested in interfacial conductance consult
3736 + reference \citealp{kuang:AuThl} for other approaches to computing $G$.
3737 + Users interested in {\it friction coefficients} at heterogeneous
3738 + interfaces may also find reference \citealp{2012MolPh.110..691K}
3739 + useful.
3740  
3741 + \newpage
3742 +
3743 + \begin{longtable}[c]{JKLM}
3744 + \caption{Meta-data Keywords: Parameters for RNEMD simulations}\\
3745 + \multicolumn{4}{c}{The following keywords must be enclosed inside a {\tt RNEMD\{...\}} block.}
3746 + \\ \hline
3747 + {\bf keyword} & {\bf units} & {\bf use} & {\bf remarks}  \\ \hline
3748 + \endhead
3749 + \hline
3750 + \endfoot
3751 + {\tt useRNEMD} & logical & perform RNEMD? & default is ``false'' \\
3752 + {\tt objectSelection} & string & see section \ref{section:syntax}
3753 + for selection syntax & default is ``select all'' \\
3754 + {\tt method} & string & exchange method & one of the following:
3755 + {\tt Swap, NIVS,} or {\tt VSS}  (default is {\tt VSS}) \\
3756 + {\tt fluxType} & string & what is being exchanged between slabs? & one
3757 + of the following: {\tt KE, Px, Py, Pz, Pvector, KE+Px, KE+Py, KE+Pvector} \\
3758 + {\tt kineticFlux} & kcal mol$^{-1}$ \AA$^{-2}$ fs$^{-1}$ & specify the kinetic energy flux &  \\
3759 + {\tt momentumFlux} & amu \AA$^{-1}$ fs$^{-2}$ & specify the momentum flux & \\
3760 + {\tt momentumFluxVector} & amu \AA$^{-1}$ fs$^{-2}$ & specify the momentum flux when
3761 + {\tt Pvector} is part of the exchange & Vector3d input\\
3762 + {\tt exchangeTime} & fs & how often to perform the exchange & default is 100 fs\\
3763 +
3764 + {\tt slabWidth} & $\mbox{\AA}$ & width of the two exchange slabs & default is $\mathsf{H}_{zz} / 10.0$ \\
3765 + {\tt slabAcenter} & $\mbox{\AA}$ & center of the end slab & default is 0 \\
3766 + {\tt slabBcenter} & $\mbox{\AA}$ & center of the middle slab & default is $\mathsf{H}_{zz} / 2$ \\
3767 + {\tt outputFileName} & string & file name for output histograms & default is the same prefix as the
3768 + .md file, but with the {\tt .rnemd} extension \\
3769 + {\tt outputBins} & int & number of $z$-bins in the output histogram &
3770 + default is 20 \\
3771 + {\tt outputFields} & string & columns to print in the {\tt .rnemd}
3772 + file where each column is separated by a pipe ($\mid$) symbol. & Allowed column names are: {\sc z, temperature, velocity, density} \\
3773 + \label{table:rnemd}
3774 + \end{longtable}
3775 +
3776   \chapter{\label{section:minimizer}Energy Minimization}
3777  
3778 < As one of the basic procedures of molecular modeling, energy
3083 < minimization is used to identify local configurations that are stable
3778 > Energy minimization is used to identify local configurations that are stable
3779   points on the potential energy surface. There is a vast literature on
3780   energy minimization algorithms have been developed to search for the
3781   global energy minimum as well as to find local structures which are
# Line 3207 | Line 3902 | diagram of the class heirarchy:
3902   \begin{figure}
3903   \centering
3904   \includegraphics[width=3in]{heirarchy.pdf}
3905 < \caption[Class heirarchy for StuntDoubles in {\sc OpenMD}-4]{ \\ The
3906 < class heirarchy of StuntDoubles in {\sc OpenMD}-4. The selection
3905 > \caption[Class heirarchy for StuntDoubles in {\sc OpenMD}]{ \\ The
3906 > class heirarchy of StuntDoubles in {\sc OpenMD}. The selection
3907   syntax allows the user to select any of the objects that are descended
3908   from a StuntDouble.}
3909   \label{fig:heirarchy}
# Line 3388 | Line 4083 | VMD. The options available for Dump2XYZ are as follows
4083    -z & {\tt -{}-zconstraint}  &                replace the atom types of zconstraint molecules  (default=off) \\
4084    -r & {\tt -{}-rigidbody}  &                  add a pseudo COM atom to rigidbody  (default=off) \\
4085    -t & {\tt -{}-watertype} &                   replace the atom type of water model (default=on) \\
4086 <  -b & {\tt -{}-basetype}  &                   using base atom type  (default=off) \\
4086 >  -b & {\tt -{}-basetype}  &                   using base atom type
4087 >  (default=off) \\
4088 >  -v& {\tt -{}-velocities}             & Print velocities in xyz file  (default=off)\\
4089 >  -f& {\tt -{}-forces}                 & Print forces xyz file  (default=off)\\
4090 >  -u& {\tt -{}-vectors}                & Print vectors (dipoles, etc) in xyz file  
4091 >                                  (default=off)\\
4092 >  -c& {\tt -{}-charges}                & Print charges in xyz file  (default=off)\\
4093 >  -e& {\tt -{}-efield}                 & Print electric field vector in xyz file  
4094 >                                  (default=off)\\
4095       & {\tt -{}-repeatX=INT}  &                 The number of images to repeat in the x direction  (default=`0') \\
4096       & {\tt -{}-repeatY=INT} &                 The number of images to repeat in the y direction  (default=`0') \\
4097       &  {\tt -{}-repeatZ=INT}  &                The number of images to repeat in the z direction  (default=`0') \\
# Line 3480 | Line 4183 | The options available for {\tt StaticProps} are as fol
4183      & {\tt -{}-sele1=selection script}   & select the first StuntDouble set \\
4184      & {\tt -{}-sele2=selection script}   & select the second StuntDouble set \\
4185      & {\tt -{}-sele3=selection script}   & select the third StuntDouble set \\
4186 <    & {\tt -{}-refsele=selection script} & select reference (can only be used with {\tt -{}-gxyz}) \\
4186 >    & {\tt -{}-refsele=selection script} & select reference (can only
4187 >    be used with {\tt -{}-gxyz}) \\
4188 >    & {\tt -{}-comsele=selection script}
4189 >                               & select stunt doubles for center-of-mass
4190 >                                  reference point\\
4191 >    & {\tt -{}-seleoffset=INT}        & global index offset for a second object (used
4192 >                                  to define a vector between sites in molecule)\\
4193 >
4194      & {\tt -{}-molname=STRING}           & molecule name \\
4195      & {\tt -{}-begin=INT}                & begin internal index \\
4196      & {\tt -{}-end=INT}                  & end internal index \\
4197 +    & {\tt -{}-radius=DOUBLE}            & nanoparticle radius\\
4198   \hline
4199   \multicolumn{3}{|l|}{One option from the following group of options is required:} \\
4200   \hline
4201 <    &  {\tt -{}-gofr}                    &  $g(r)$ \\
4202 <    &  {\tt -{}-r\_theta}                 &  $g(r, \cos(\theta))$ \\
4203 <    &  {\tt -{}-r\_omega}                 &  $g(r, \cos(\omega))$ \\
4204 <    &  {\tt -{}-theta\_omega}             &  $g(\cos(\theta), \cos(\omega))$ \\
4205 <    &  {\tt -{}-gxyz}                    &  $g(x, y, z)$ \\
4206 <    &  {\tt -{}-p2}                      &  $P_2$ order parameter ({\tt -{}-sele1} and {\tt -{}-sele2} must be specified) \\
4201 >    & {\tt -{}-bo}          & bond order parameter ({\tt -{}-rcut} must be specified) \\
4202 >    & {\tt -{}-bor}         & bond order parameter as a function of
4203 >    radius  ({\tt -{}-rcut} must be specified) \\
4204 >    & {\tt -{}-bad}         & $N(\theta)$ bond angle density within ({\tt -{}-rcut} must be specified) \\
4205 >    & {\tt -{}-count}       & count of molecules matching selection
4206 >    criteria (and associated statistics) \\
4207 >  -g&  {\tt -{}-gofr}                    &  $g(r)$ \\
4208 >    &  {\tt -{}-gofz}                    &  $g(z)$ \\
4209 >    &  {\tt -{}-r\_theta}                &  $g(r, \cos(\theta))$ \\
4210 >    &  {\tt -{}-r\_omega}                &  $g(r, \cos(\omega))$ \\
4211 >    &  {\tt -{}-r\_z}                    &  $g(r, z)$ \\
4212 >    &  {\tt -{}-theta\_omega}            &  $g(\cos(\theta), \cos(\omega))$ \\
4213 >    &  {\tt -{}-gxyz}                    &  $g(x, y, z)$ \\
4214 >    &  {\tt -{}-twodgofr}                & 2D $g(r)$ (Slab width {\tt -{}-dz} must be specified)\\
4215 >  -p&  {\tt -{}-p2}                      &  $P_2$ order parameter  ({\tt -{}-sele1} must be specified, {\tt -{}-sele2} is optional) \\
4216 >    &  {\tt -{}-rp2}                     &  Ripple order parameter ({\tt -{}-sele1} and {\tt -{}-sele2} must be specified) \\
4217      &  {\tt -{}-scd}                     &  $S_{CD}$ order parameter(either {\tt -{}-sele1}, {\tt -{}-sele2}, {\tt -{}-sele3} are specified or {\tt -{}-molname}, {\tt -{}-begin}, {\tt -{}-end} are specified) \\
4218 <    &  {\tt -{}-density}                 &  density plot ({\tt -{}-sele1} must be specified) \\
4219 <    &  {\tt -{}-slab\_density}           &  slab density ({\tt -{}-sele1} must be specified)
4218 >  -d&  {\tt -{}-density}                 &  density plot \\
4219 >    &  {\tt -{}-slab\_density}           &  slab density \\
4220 >    &  {\tt -{}-p\_angle}                & $p(\cos(\theta))$ ($\theta$
4221 >    is the angle between molecular axis and radial vector from origin\\
4222 >    &  {\tt -{}-hxy}                     & Calculates the undulation  spectrum, $h(x,y)$, of an interface \\
4223 >    &  {\tt -{}-rho\_r}                  & $\rho(r)$\\
4224 >    &  {\tt -{}-angle\_r}                &  $\theta(r)$ (spatially resolves the
4225 >    angle between the molecular axis and the radial vector from the
4226 >    origin)\\
4227 >    &  {\tt -{}-hullvol}                 & hull volume of nanoparticle\\
4228 >    &  {\tt -{}-rodlength}               & length of nanorod\\
4229 >  -Q&  {\tt -{}-tet\_param}              & tetrahedrality order parameter ($Q$)\\
4230 >    &  {\tt -{}-tet\_param\_z}           & spatially-resolved tetrahedrality order
4231 >                                   parameter $Q(z)$\\
4232 >    &  {\tt -{}-rnemdz}                  & slab-resolved RNEMD statistics (temperature,
4233 >                                  density, velocity)\\
4234 >    &  {\tt -{}-rnemdr}                  & shell-resolved RNEMD statistics (temperature,
4235 >                                  density, angular velocity)
4236   \end{longtable}
4237  
4238   \subsection{\label{section:DynamicProps}DynamicProps}
# Line 3536 | Line 4273 | The options available for DynamicProps are as follows:
4273    -o& {\tt -{}-output=filename}        & output file name \\
4274      & {\tt -{}-sele1=selection script} & select first StuntDouble set \\
4275      & {\tt -{}-sele2=selection script} & select second StuntDouble set (if sele2 is not set, use script from sele1) \\
4276 +    & {\tt -{}-order=INT}              & Lengendre Polynomial Order\\
4277 +  -z& {\tt -{}-nzbins=INT}             & Number of $z$ bins (default=`100`)\\
4278 +  -m& {\tt -{}-memory=memory specification}
4279 +                                &Available memory  
4280 +                                  (default=`2G`)\\
4281   \hline
4282   \multicolumn{3}{|l|}{One option from the following group of options is required:} \\
4283   \hline
4284 <  -r& {\tt -{}-rcorr}                  & compute mean square displacement \\
4285 <  -v& {\tt -{}-vcorr}                  & compute velocity correlation function \\
4286 <  -d& {\tt -{}-dcorr}                  & compute dipole correlation function
4284 >  -s& {\tt -{}-selecorr}               & selection correlation function \\
4285 >  -r& {\tt -{}-rcorr}                  & compute mean squared displacement \\
4286 >  -v& {\tt -{}-vcorr}                  & velocity autocorrelation function \\
4287 >  -d& {\tt -{}-dcorr}                  & dipole correlation function \\
4288 >  -l& {\tt -{}-lcorr}                  & Lengendre correlation function \\
4289 >    & {\tt -{}-lcorrZ}                 & Lengendre correlation function binned by $z$ \\
4290 >    & {\tt -{}-cohZ}                   & Lengendre correlation function for OH bond vectors binned by $z$\\
4291 >  -M& {\tt -{}-sdcorr}                 & System dipole correlation function\\
4292 >    & {\tt -{}-r\_rcorr}               & Radial mean squared displacement\\
4293 >    & {\tt -{}-thetacorr}              & Angular mean squared displacement\\
4294 >    & {\tt -{}-drcorr}                 & Directional mean squared displacement for particles with unit vectors\\
4295 >    & {\tt -{}-helfandEcorr}           & Helfand moment for thermal conductvity\\
4296 >  -p& {\tt -{}-momentum}               & Helfand momentum for viscosity\\
4297 >    & {\tt -{}-stresscorr}             & Stress tensor correlation function
4298   \end{longtable}
4299  
4300   \chapter{\label{section:PreparingInput} Preparing Input Configurations}
# Line 3608 | Line 4361 | expect the the input specifier on the command line.
4361   to {\tt atom2md}, but they use a specific input format and do not
4362   expect the the input specifier on the command line.
4363  
4364 +
4365   \section{\label{section:SimpleBuilder}SimpleBuilder}
4366  
4367   {\tt SimpleBuilder} creates simple lattice structures.  It requires an
# Line 3634 | Line 4388 | The options available for SimpleBuilder are as follows
4388      &  {\tt -{}-nz=INT}            &  number of unit cells in z
4389   \end{longtable}
4390  
4391 + \section{\label{section:icosahedralBuilder}icosahedralBuilder}
4392 +
4393 + {\tt icosahedralBuilder} creates single-component geometric solids
4394 + that can be useful in simulating nanostructures.  Like the other
4395 + builders, it requires an initial, but skeletal {\sc OpenMD} file to
4396 + specify the component that is to be placed on the lattice.  The total
4397 + number of placed molecules will be shown at the top of the
4398 + configuration file that is generated, and that number may not match
4399 + the original meta-data file, so a new meta-data file is also generated
4400 + which matches the lattice structure.
4401 +
4402 + The options available for icosahedralBuilder are as follows:
4403 + \begin{longtable}[c]{|EFG|}
4404 + \caption{icosahedralBuilder Command-line Options}
4405 + \\ \hline
4406 + {\bf option} & {\bf verbose option} & {\bf behavior} \\ \hline
4407 + \endhead
4408 + \hline
4409 + \endfoot
4410 +  -h& {\tt -{}-help}               & Print help and exit\\
4411 +  -V& {\tt -{}-version}            & Print version and exit\\
4412 +  -o& {\tt -{}-output=STRING}      & Output file name\\
4413 +  -n& {\tt -{}-shells=INT}         & Nanoparticle shells\\
4414 +  -d& {\tt -{}-latticeConstant=DOUBLE} & Lattice spacing in Angstroms for cubic lattice.\\
4415 +  -c& {\tt -{}-columnAtoms=INT}        & Number of atoms along central
4416 +  column (Decahedron only)\\
4417 +  -t& {\tt -{}-twinAtoms=INT}          & Number of atoms along twin
4418 +  boundary (Decahedron only) \\
4419 +  -p& {\tt -{}-truncatedPlanes=INT}   & Number of truncated planes (Curling-stone Decahedron only)\\
4420 + \hline
4421 + \multicolumn{3}{|l|}{One option from the following group of options is required:} \\
4422 + \hline
4423 +   & {\tt -{}-ico}    & Create an Icosahedral cluster \\
4424 +   & {\tt -{}-deca}   & Create a regualar Decahedral cluster\\
4425 +   & {\tt -{}-ino}    & Create an Ino Decahedral cluster\\
4426 +   & {\tt -{}-marks}  & Create a Marks Decahedral cluster\\
4427 +   & {\tt -{}-stone}  & Create a Curling-stone Decahedral cluster
4428 + \end{longtable}
4429 +
4430 +
4431   \section{\label{section:Hydro}Hydro}
4432   {\tt Hydro} generates resistance tensor ({\tt .diff}) files which are
4433   required when using the Langevin integrator using complex rigid
# Line 3667 | Line 4461 | hydrodynamic calculations will not be performed (defau
4461   \end{longtable}
4462  
4463  
4464 +
4465 +
4466 +
4467   \chapter{\label{section:parallelization} Parallel Simulation Implementation}
4468  
4469   Although processor power is continually improving, it is still
# Line 3750 | Line 4547 | DMR-0079647.
4547   DMR-0079647.
4548  
4549  
4550 < \bibliographystyle{jcc}
4550 > \bibliographystyle{aip}
4551   \bibliography{openmdDoc}
4552  
4553   \end{document}

Diff Legend

Removed lines
+ Added lines
< Changed lines
> Changed lines