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

Comparing trunk/openmdDocs/openmdDoc.tex (file contents):
Revision 4102 by gezelter, Fri Apr 18 18:39:38 2014 UTC vs.
Revision 4103 by gezelter, Tue Apr 22 18:19:31 2014 UTC

# Line 779 | 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:forceFields}Force Fields}
782 > \chapter{\label{chapter:forceFields}Force Fields}
783  
784   Like many molecular simulation packages, {\sc OpenMD} splits the
785   potential energy into the short-ranged (bonded) portion and a
# Line 793 | Line 793 | section.
793   are defined by the force field which is selected in the MetaData
794   section.
795  
796 < \section{\label{section:shortRange}The basic interactions}
796 > \section{\label{section:divisionOfLabor}Separation into Internal and
797 >  Cross interactions}
798  
799 < The energy function for a system composed of $N$ molecules is
800 < traditionally written
799 > The classical potential energy function for a system composed of $N$
800 > molecules is traditionally written
801   \begin{equation}
802   V = \sum^{N}_{I=1} V^{I}_{\text{Internal}}
803          + \sum^{N-1}_{I=1} \sum_{J>I} V^{IJ}_{\text{Cross}},
804   \label{eq:totalPotential}
805   \end{equation}
806 < where $V^{IJ}_{\text{Cross}}$ contains all intermolecular interactions
807 < between molecules $I$ and $J$, and $V^{I}_{\text{Internal}}$ is the
808 < internal potential of molecule $I$:
809 < \begin{align*}
810 < V^{I}_{\text{Internal}} =  &
811 < \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.
806 > where $V^{I}_{\text{Internal}}$ contains all of the terms internal to
807 > molecule $I$ (e.g. bonding, bending, torsional, and inversion terms)
808 > and $V^{IJ}_{\text{Cross}}$ contains all intermolecular interactions
809 > between molecules $I$ and $J$.  For large molecules, the internal
810 > potential may also include some non-bonded terms like electrostatic or
811 > van der Waals interactions.
812  
813   The types of atoms being simulated, as well as the specific functional
814   forms and parameters of the intra-molecular functions and the
# Line 900 | Line 875 | end TorsionTypes
875   end TorsionTypes
876   \end{lstlisting}
877  
878 < \subsection{\label{section:ffOptions}The Options block}
878 > \section{\label{section:ffOptions}The Options block}
879  
880   The Options block defines properties governing how the force field
881   interactions are carried out.  This block is delineated with the text
# Line 918 | Line 893 | begin Options
893   Name                      = "alkane"       // any string
894   vdWtype                   = "Lennard-Jones"
895   DistanceMixingRule        = "arithmetic"   // can also be "geometric" or "cubic"
896 < DistanceType              = "sigma"        // can also be Rmin
896 > DistanceType              = "sigma"        // can also be "Rmin"
897   EnergyMixingRule          = "geometric"    // can also be "arithmetic" or "hhg"
898   EnergyUnitScaling         = 1.0
899   MetallicEnergyUnitScaling = 1.0
# Line 937 | Line 912 | end Options
912   end Options
913   \end{lstlisting}
914  
915 < \subsection{\label{section:ffBase}The BaseAtomTypes block}
915 > \section{\label{section:ffBase}The BaseAtomTypes block}
916  
917   An AtomType the primary data structure that OpenMD uses to store
918   static data about an atom.  Things that belong to AtomType are
# Line 985 | Line 960 | end BaseAtomTypes
960   end BaseAtomTypes
961   \end{lstlisting}
962  
963 < \subsection{\label{section:ffAtom}The AtomTypes block}
963 > \section{\label{section:ffAtom}The AtomTypes block}
964  
965   AtomTypes inherit most properties from BaseAtomTypes, but can override
966   their lower-level properties as well.  Scheme \ref{sch:atomTypesBlock}
# Line 1019 | Line 994 | end AtomTypes
994   end AtomTypes
995   \end{lstlisting}
996  
997 < \subsection{\label{section:ffDirectionalAtom}The DirectionalAtomTypes
997 > \section{\label{section:ffDirectionalAtom}The DirectionalAtomTypes
998    block}
999   DirectionalAtoms have orientational degrees of freedom as well as
1000   translation, so moving these atoms requires information about the
# Line 1054 | Line 1029 | calculation of temperatures to reflect this.
1029   of freedom (three translations and two rotations), and will alter
1030   calculation of temperatures to reflect this.
1031  
1032 < \subsection{\label{section::ffAtomProperties}AtomType properties}
1033 < \subsubsection{\label{section:ffLJ}The LennardJonesAtomTypes block}
1032 > \section{\label{section::ffAtomProperties}AtomType properties}
1033 > \subsection{\label{section:ffLJ}The LennardJonesAtomTypes block}
1034   One of the most basic interatomic interactions implemented in {\sc
1035    OpenMD} is the Lennard-Jones potential, which mimics the van der
1036   Waals interaction at long distances and uses an empirical repulsion at
# Line 1113 | Line 1088 | end LennardJonesAtomTypes
1088   end LennardJonesAtomTypes
1089   \end{lstlisting}
1090  
1091 < \subsubsection{\label{section:ffCharge}The ChargeAtomTypes block}
1091 > \subsection{\label{section:ffCharge}The ChargeAtomTypes block}
1092  
1093   In molecular simulations, proper accumulation of the electrostatic
1094   interactions is essential and is one of the most
# Line 1153 | Line 1128 | end ChargeAtomTypes
1128   end ChargeAtomTypes
1129   \end{lstlisting}
1130  
1131 < \subsubsection{\label{section:ffMultipole}The MultipoleAtomTypes
1131 > \subsection{\label{section:ffMultipole}The MultipoleAtomTypes
1132    block}
1133   For complex charge distributions that are centered on single sites, it
1134   is convenient to write the total electrostatic potential in terms of
# Line 1169 | Line 1144 | given by $C_{\bf a}$, $D_{{\bf a}\alpha}$, and $Q_{{\b
1144   \end{equation}
1145   Here, the point charge, dipole, and quadrupole for site $\bf a$ are
1146   given by $C_{\bf a}$, $D_{{\bf a}\alpha}$, and $Q_{{\bf
1147 <    a}\alpha\beta}$, respectively.  These are the primitive
1147 >    a}\alpha\beta}$, respectively.  These are the {\it primitive}
1148   multipoles.  If the site is representing a distribution of charges,
1149   these can be expressed as,
1150   \begin{align}
# Line 1234 | Line 1209 | and Quadrupole moments in units of Debye \AA.
1209   in units of degrees.  Dipole moments are entered in units of Debye,
1210   and Quadrupole moments in units of Debye \AA.
1211  
1212 < \subsubsection{\label{section:ffGB}The FluctuatingChargeAtomTypes  block}
1213 < \subsubsection{\label{section:ffPol}The PolarizableAtomTypes block}
1239 < \subsubsection{\label{section:ffGB}The GayBerneAtomTypes block}
1212 > \subsection{\label{section:ffGB}The FluctuatingChargeAtomTypes  block}
1213 > %\subsubsection{\label{section:ffPol}The PolarizableAtomTypes block}
1214  
1215 + \subsection{\label{section:ffGB}The GayBerneAtomTypes block}
1216 +
1217   The Gay-Berne potential has been widely used in the liquid crystal
1218 < community to describe this anisotropic phase
1218 > community to describe anisotropic phase
1219   behavior.~\cite{Gay:1981yu,Berne:1972pb,Kushick:1976xy,Luckhurst:1990fy,Cleaver:1996rt}
1220   The form of the Gay-Berne potential implemented in OpenMD was
1221   generalized by Cleaver {\it et al.} and is appropriate for dissimilar
1222 < uniaxial ellipsoids.\cite{Cleaver:1996rt}  The potential is constructed in the
1223 < familiar form of the Lennard-Jones function using
1222 > uniaxial ellipsoids.\cite{Cleaver:1996rt} The potential is constructed
1223 > in the familiar form of the Lennard-Jones function using
1224   orientation-dependent $\sigma$ and $\epsilon$ parameters,
1225   \begin{equation*}
1226   V_{ij}({{\bf \hat u}_i}, {{\bf \hat u}_j}, {{\bf \hat
# Line 1302 | Line 1278 | end GayBerneAtomTypes                  
1278   end GayBerneAtomTypes                  
1279   \end{lstlisting}
1280  
1281 < \subsubsection{\label{section:ffSticky}The StickyAtomTypes block}
1281 > \subsection{\label{section:ffSticky}The StickyAtomTypes block}
1282  
1283   One of the solvents that can be simulated by {\sc OpenMD} is the
1284   extended Soft Sticky Dipole (SSD/E) water model.\cite{fennell04} The
# Line 1412 | Line 1388 | end StickyAtomTypes
1388   end StickyAtomTypes
1389   \end{lstlisting}
1390  
1391 < \subsection{\label{section::ffMetals}Metallic Atom Types}
1391 > \section{\label{section::ffMetals}Metallic Atom Types}
1392  
1393   {\sc OpenMD} implements a number of related potentials that describe
1394   bonding in transition metals. These potentials have an attractive
# Line 1441 | Line 1417 | repulsive interaction between atoms $i$ and $j$.
1417   The pairwise portion of the potential, $\phi_{ij}$, is usually a
1418   repulsive interaction between atoms $i$ and $j$.
1419  
1420 < \subsubsection{\label{section:ffEAM}The EAMAtomTypes block}
1420 > \subsection{\label{section:ffEAM}The EAMAtomTypes block}
1421   The Embedded Atom Method ({\sc eam}) is one of the most widely-used
1422   potentials for transition
1423   metals.~\cite{Finnis84,Ercolessi88,Chen90,Qi99,Ercolessi02,Daw84,FBD86,johnson89,Lu97}
# Line 1507 | Line 1483 | end EAMAtomTypes
1483   end EAMAtomTypes
1484   \end{lstlisting}
1485  
1486 + \subsection{\label{section:ffSC}The SuttonChenAtomTypes block}
1487  
1511 \subsubsection{\label{section:ffSC}The SuttonChenAtomTypes block}
1512
1488   The Sutton-Chen ({\sc sc})~\cite{Chen90} potential has been used to
1489   study a wide range of phenomena in metals.  Although it has the same
1490 < basic form as the {\sc eam} potential, the Sutton-Chen model takes on
1491 < a simpler form,
1490 > basic form as the {\sc eam} potential, the Sutton-Chen model requires
1491 > a simpler set of parameters,
1492   \begin{equation}
1493   \label{eq:SCP1}
1494   U_{tot}=\sum _{i}\left[ \frac{1}{2}\sum _{j\neq
# Line 1563 | Line 1538 | end SCAtomTypes
1538   end SCAtomTypes
1539   \end{lstlisting}
1540  
1541 < \subsection{\label{section::ffShortRange}Short Range Interactions}
1542 < \subsubsection{\label{section:ffBond}The BondTypes block}
1543 < \subsubsection{\label{section:ffBend}The BendTypes block}
1544 < 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:
1541 > \section{\label{section::ffShortRange}Short Range Interactions}
1542 > The internal structure of a molecule is usually specified in terms of
1543 > a set of ``bonded'' terms in the potential energy function for
1544 > molecule $I$,
1545   \begin{align*}
1546 < k_0 &= c_1 + c_3, \\
1547 < k_1 &= c_1 - 3c_3, \\
1548 < k_2 &= 2 c_2, \\
1549 < k_3 &= 4c_3.
1546 > V^{I}_{\text{Internal}} =  &
1547 > \sum_{r_{ij} \in I} V_{\text{bond}}(r_{ij})
1548 > + \sum_{\theta_{ijk} \in I} V_{\text{bend}}(\theta_{ijk})
1549 > + \sum_{\phi_{ijkl} \in I} V_{\text{torsion}}(\phi_{ijkl})
1550 > + \sum_{\omega_{ijkl} \in I} V_{\text{inversion}}(\omega_{ijkl}) \\
1551 > & + \sum_{i \in I} \sum_{(j>i+4) \in I}
1552 > \biggl[ V_{\text{dispersion}}(r_{ij}) +  V_{\text{electrostatic}}
1553 > (\mathbf{r}_{ij},\boldsymbol{\Omega}_{i},\boldsymbol{\Omega}_{j})
1554 > \biggr].
1555 > \label{eq:internalPotential}
1556   \end{align*}
1557 < By recasting the potential as a power series, repeated trigonometric
1558 < evaluations are avoided during the calculation of the potential
1559 < energy.
1557 > Here $V_{\text{bond}}, V_{\text{bend}},
1558 > V_{\text{torsion}},\mathrm{~and~} V_{\text{inversion}}$ represent the
1559 > bond, bend, torsion, and inversion potentials for all
1560 > topologically-connected sets of atoms within the molecule.  Bonds are
1561 > the primary way of specifying how the atoms are connected together to
1562 > form the molecule (i.e. they define the molecular topology).  The
1563 > other short-range interactions may be specified explicitly in the
1564 > molecule definition, or they may be deduced from bonding information.
1565 > For example, bends can be implicitly deduced from two bonds which
1566 > share an atom.  Torsions can be deduced from two bends that share a
1567 > bond.  Inversion potentials are utilized primarily to enforce
1568 > planarity around $sp^2$-hybridized sites, and these are specified with
1569 > central atoms and satellites (or an atom with bonds to exactly three
1570 > satellites). Non-bonded interactions are usually excluded for atom
1571 > pairs that are involved in the same bond, bend, or torsion, but all
1572 > other atom pairs are included in the standard non-bonded interactions.
1573  
1574 < \subsubsection{\label{section:ffInversion}The InversionTypes block}
1575 < \subsection{\label{section::ffLongRange}Long Range Interactions}
1576 < \subsubsection{\label{section:ffNBinteraction}The NonBondedInteraction block}
1574 > Bond lengths, angles, and torsions (dihedrals) are ``natural''
1575 > coordinates to treat molecular motion, as it is usually in these
1576 > coordinates that most chemists understand the behavior of molecules.
1577 > The bond lengths and angles are often considered ``hard'' degrees of
1578 > freedom.  That is, we can't deform them very much without a
1579 > significant energetic penalty.  On the other hand, dihedral angles or
1580 > torsions are ``soft'' and typically undergo significant deformation
1581 > under normal conditions.
1582  
1583 + \subsection{\label{section:ffBond}The BondTypes block}
1584  
1585 + Bonds are the primary way to specify how the atoms are connected
1586 + together to form the molecule.  In general, bonds exert strong
1587 + restoring forces to keep the molecule compact.  The bond energy
1588 + functions are relatively simple functions of the distance between two
1589 + atomic sites,
1590 + \begin{equation}
1591 + b = \left| \vec{r}_{ij} \right| = \sqrt{(x_j - x_i)^2 + (y_j - y_i)^2
1592 +  + (z_j - z_i)^2}.
1593 + \end{equation}
1594 + All BondTypes must specify two AtomType names ($i$ and $j$) that
1595 + describe when that bond should be applied, as well as an equilibrium
1596 + bond length, $b_{ij}^0$, in units of \AA. The most common forms for
1597 + bonding potentials are {\tt Harmonic} bonds,
1598 + \begin{equation}
1599 + V_{\text{bond}}(b) = \frac{k_{ij}}{2} \left(b -
1600 +  b_{ij}^0 \right)^2
1601 + \end{equation}
1602 + and {\tt Morse} bonds,
1603 + \begin{equation}
1604 + V_{\text{bond}}(b) = D_{ij} \left[ 1 - e^{-\beta_{ij} (b - b_{ij}^0)} \right]^2
1605 + \end{equation}
1606  
1607 < (see Fig.~\ref{fig:lipidModel}), The parameters for $k_{\theta}$ and
1608 < $\theta_0$ are borrowed from those in TraPPE.\cite{Siepmann1998}
1607 > OpenMD can also simulate some less common types of bond potentials,
1608 > including {\tt Fixed} bonds (which are constrained to be at a
1609 > specified bond length),
1610 > \begin{equation}
1611 > V_{\text{bond}}(b) = 0.
1612 > \end{equation}
1613 > {\tt Cubic} bonds include terms to model anharmonicity,
1614 > \begin{equation}
1615 > V_{\text{bond}}(b) =  K_3 (b -  b_{ij}^0)^3 + K_2 (b - b_{ij}^0)^2 + K_1 (b -  b_{ij}^0) + K_0,
1616 > \end{equation}
1617 > and {\tt Quartic} bonds, include another term in the Taylor
1618 > expansion around $b_{ij}^0$,
1619 > \begin{equation}
1620 > V_{\text{bond}}(b) = K_4 (b -  b_{ij}^0)^4 +  K_3 (b -  b_{ij}^0)^3 +
1621 > K_2 (b - b_{ij}^0)^2 + K_1 (b -  b_{ij}^0) + K_0,
1622 > \end{equation}
1623 > can also be simulated.  Note that the factor of $1/2$ that is included
1624 > in the {\tt Harmonic} bond type force constant is {\it not} present in
1625 > either the {\tt Cubic} or {\tt Quartic} bond types.
1626  
1627 < 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:
1627 > {\tt Polynomial} bonds which can have any number of terms,
1628   \begin{equation}
1629 < V_{\mathrm{long-range}} = \sum_{a} \sum_{b>a} s(r_{ab}) \sum_{i \in a} \sum_{j \in b} V_{ij}(r_{ij}),
1629 > V_{\text{bond}}(b) = \sum_n K_n (b -  b_{ij}^0)^n.
1630   \end{equation}
1631 < where $r_{ab}$ is the distance between the centers of mass of the two
1632 < cutoff groups ($a$ and $b$).
1631 > can also be specified by giving a sequence of integer ($n$) and force
1632 > constant ($K_n$) pairs.
1633  
1634 < The sums over $a$ and $b$ are over the cutoff groups that are present
1635 < in the simulation.  Atoms which are not explicitly defined as members
1636 < of a {\tt cutoffGroup} are treated as a group consisting of only one
1637 < atom.  The switching function, $s(r)$ is the standard cubic switching
1638 < function,
1634 > The order of terms in the BondTypes block is:
1635 > \begin{itemize}
1636 > \item {\tt AtomType} 1
1637 > \item {\tt AtomType} 2
1638 > \item {\tt BondType} (one of {\tt Harmonic}, {\tt Morse}, {\tt Fixed}, {\tt
1639 >        Cubic}, {\tt Quartic}, or {\tt Polynomial})
1640 > \item $b_{ij}^0$, the equilibrium bond length in \AA
1641 > \item any other parameters required by the {\tt BondType}
1642 > \end{itemize}
1643 >
1644 > \begin{lstlisting}[caption={[An example of a BondTypes block.] A
1645 > simple example of a BondTypes block.  Distances ($b_0$)
1646 > are given in \AA\ and force constants are given in
1647 > units so that when multiplied by the correct power of distance they
1648 > return energies in kcal/mol.  For example $k$ for a Harmonic bond is
1649 > in units of kcal/mol/\AA$^2$.},
1650 > label={sch:BondTypes}]
1651 > begin BondTypes
1652 > //Atom1 Atom2   Harmonic        b0        k (kcal/mol/A^2)
1653 > CH2     CH2     Harmonic        1.526     260
1654 > //Atom1 Atom2   Morse           b0        D       beta (A^-1)
1655 > CN      NC      Morse           1.157437  212.95  2.5802
1656 > //Atom1 Atom2   Fixed           b0
1657 > CT      HC      Fixed           1.09
1658 > //Atom1 Atom2   Cubic           b0        K3      K2      K1      K0
1659 > //Atom1 Atom2   Quartic         b0        K4      K3      K2      K1      K0
1660 > //Atom1 Atom2   Polynomial      b0        n       Kn      [m      Km]
1661 > end BondTypes
1662 > \end{lstlisting}
1663 >
1664 > There are advantages and disadvantages of all of the different types
1665 > of bonds, but specific simulation tasks may call for specific
1666 > behaviors.
1667 >
1668 > \subsection{\label{section:ffBend}The BendTypes block}
1669 > The equilibrium geometries and energy functions for bending motions in
1670 > a molecule are strongly dependent on the bonding environment of the
1671 > central atomic site.  For example, different types of hybridized
1672 > carbon centers require different bending angles and force constants to
1673 > describe the local geometry.
1674 >
1675 > The bending potential energy functions used in most force fields are
1676 > often simple functions of the angle between two bonds,
1677   \begin{equation}
1678 < S(r) =
1679 <        \begin{cases}
1680 <        1 & \text{if $r \le r_{\text{sw}}$},\\
1681 <        \frac{(r_{\text{cut}} + 2r - 3r_{\text{sw}})(r_{\text{cut}} - r)^2}
1682 <        {(r_{\text{cut}} - r_{\text{sw}})^3}
1683 <        & \text{if $r_{\text{sw}} < r \le r_{\text{cut}}$}, \\
1684 <        0 & \text{if $r > r_{\text{cut}}$.}
1685 <        \end{cases}
1686 < \label{eq:dipoleSwitching}
1678 > \theta_{ijk} = \cos^{-1} \left(\frac{\vec{r}_{ij} \cdot
1679 >    \vec{r}_{jk}}{\left| \vec{r}_{ij} \right| \left| \vec{r}_{ij}
1680 >    \right|} \right)
1681 > \end{equation}
1682 > Here atom $j$ is the central atom that is bonded to two partners $i$
1683 > and $k$.
1684 >
1685 > All BendTypes must specify three AtomType names ($i$, $j$ and $k$)
1686 > that describe when that bend potential should be applied, as well as
1687 > an equilibrium bending angle, $\theta_{ijk}^0$, in units of
1688 > degrees. The most common forms for bending potentials are {\tt
1689 >  Harmonic} bends,
1690 > \begin{equation}
1691 > V_{\text{bend}}(\theta_{ijk}) = \frac{k_{ijk}}{2}( \theta_{ijk} - \theta_{ijk}^0
1692 > )^2, \label{eq:bendPot}
1693   \end{equation}
1694 < Here, $r_{\text{sw}}$ is the {\tt switchingRadius}, or the distance
1695 < beyond which interactions are reduced, and $r_{\text{cut}}$ is the
1696 < {\tt cutoffRadius}, or the distance at which interactions are
1697 < truncated.
1694 > where $k_{ijk}$ is the force constant which determines the strength of
1695 > the harmonic bend. {\tt UreyBradley} bends utilize an additional 1-3
1696 > bond-type interaction in addition to the harmonic bending potential:
1697 > \begin{equation}
1698 >  V_{\text{bend}}(\vec{r}_i , \vec{r}_j, \vec{r}_k)
1699 >  =\frac{k_{ijk}}{2}( \theta_{ijk} - \theta_{ijk}^0)^2
1700 >  + \frac{k_{ub}}{2}( r_{ik} - s_0 )^2. \label{eq:ubBend}
1701 > \end{equation}
1702  
1703 < Users of {\sc OpenMD} do not need to specify the {\tt cutoffRadius} or
1704 < {\tt switchingRadius}.  In simulations containing only Lennard-Jones
1705 < atoms, the cutoff radius has a default value of $2.5\sigma_{ii}$,
1706 < where $\sigma_{ii}$ is the largest Lennard-Jones length parameter
1707 < present in the simulation.  In simulations containing charged or
1708 < dipolar atoms, the default cutoff radius is $15 \mbox{\AA}$.  
1703 > A {\tt Cosine} bend is a variant on the harmonic bend which utilizes
1704 > the cosine of the angle instead of the angle itself,
1705 > \begin{equation}
1706 > V_{\text{bend}}(\theta_{ijk}) = \frac{k_{ijk}}{2}( \cos\theta_{ijk} -
1707 > \cos \theta_{ijk}^0 )^2. \label{eq:cosBend}
1708 > \end{equation}
1709  
1710 < The {\tt switchingRadius} is set to a default value of 95\% of the
1711 < {\tt cutoffRadius}.  In the special case of a simulation containing
1712 < {\it only} Lennard-Jones atoms, the default switching radius takes the
1713 < same value as the cutoff radius, and {\sc OpenMD} will use a shifted
1714 < potential to remove discontinuities in the potential at the cutoff.
1715 < Both radii may be specified in the meta-data file.
1710 > OpenMD can also simulate some less common types of bend potentials,
1711 > including {\tt Cubic} bends, which include terms to model anharmonicity,
1712 > \begin{equation}
1713 > V_{\text{bend}}(\theta_{ijk}) =  K_3 (\theta_{ijk} -  \theta_{ijk}^0)^3 + K_2 (\theta_{ijk} -  \theta_{ijk}^0)^2 + K_1 (\theta_{ijk} -  \theta_{ijk}^0) + K_0,
1714 > \end{equation}
1715 > and {\tt Quartic} bends, which include another term in the Taylor
1716 > expansion around $\theta_{ijk}^0$,
1717 > \begin{equation}
1718 >  V_{\text{bend}}(\theta_{ijk}) = K_4 (\theta_{ijk} -  \theta_{ijk}^0)^4 +  K_3 (\theta_{ijk} -  \theta_{ijk}^0)^3 +
1719 >  K_2 (\theta_{ijk} -  \theta_{ijk}^0)^2 + K_1 (\theta_{ijk} -
1720 >  \theta_{ijk}^0) + K_0,
1721 > \end{equation}
1722 > can also be simulated.  Note that the factor of $1/2$ that is included
1723 > in the {\tt Harmonic} bend type force constant is {\it not} present in
1724 > either the {\tt Cubic} or {\tt Quartic} bend types.
1725  
1726 < Force fields can be added to {\sc OpenMD}, although it comes with a few
1727 < simple examples (Lennard-Jones, {\sc duff}, {\sc water}, and {\sc
1728 < eam}) which are explained in the following sections.
1726 > {\tt Polynomial} bends which can have any number of terms,
1727 > \begin{equation}
1728 > V_{\text{bend}}(\theta_{ijk}) = \sum_n K_n (\theta_{ijk} -  \theta_{ijk}^0)^n.
1729 > \end{equation}
1730 > can also be specified by giving a sequence of integer ($n$) and force
1731 > constant ($K_n$) pairs.
1732  
1733 < \section{\label{sec:LJPot}The Lennard Jones Force Field}
1733 > The order of terms in the BendTypes block is:
1734 > \begin{itemize}
1735 > \item {\tt AtomType} 1
1736 > \item {\tt AtomType} 2 (this is the central atom)
1737 > \item {\tt AtomType} 3
1738 > \item {\tt BendType} (one of {\tt Harmonic}, {\tt UreyBradley}, {\tt
1739 >    Cosine}, {\tt Cubic}, {\tt Quartic}, or {\tt Polynomial})
1740 > \item $\theta_{ijk}^0$, the equilibrium bending angle in degrees.
1741 > \item any other parameters required by the {\tt BendType}
1742 > \end{itemize}
1743  
1744 < Scheme
1745 < \ref{sch:LJFF} gives an example meta-data file that
1746 < sets up a system of 108 Ar particles to be simulated using the
1747 < Lennard-Jones force field.
1744 > \begin{lstlisting}[caption={[An example of a BendTypes block.] A
1745 > simple example of a BendTypes block.  By convention, equilibrium angles
1746 > ($\theta_0$) are given in degrees but force constants are given in
1747 > units so that when multiplied by the correct power of angle (in
1748 > radians) they return energies in kcal/mol.  For example $k$ for a
1749 > Harmonic bend is in units of kcal/mol/radians$^2$.},
1750 > label={sch:BendTypes}]
1751 > begin BendTypes
1752 > //Atom1 Atom2   Atom3   Harmonic      theta0(deg) Ktheta(kcal/mol/radians^2)
1753 > CT      CT      CT      Harmonic      109.5        80.000000
1754 > CH2     CH      CH2     Harmonic      112.0       117.68
1755 > CH3     CH2     SH      Harmonic       96.0        67.220
1756 > //UreyBradley
1757 > //Atom1 Atom2   Atom3   UreyBradley   theta0      Ktheta  s0  Kub
1758 > //Cosine
1759 > //Atom1 Atom2   Atom3   Cosine        theta0      Ktheta(kcal/mol)
1760 > //Cubic
1761 > //Atom1 Atom2   Atom3   Cubic         theta0      K3      K2  K1   K0
1762 > //Quartic
1763 > //Atom1 Atom2   Atom3   Quartic       theta0      K4      K3  K2   K1   K0
1764 > //Polynomial
1765 > //Atom1 Atom2   Atom3   Polynomial    theta0      n       Kn  [m   Km]
1766 > end BendTypes
1767 > \end{lstlisting}
1768  
1769 < \begin{lstlisting}[float,caption={[Invocation of the Lennard-Jones
1770 < force field] A sample startup file for a small Lennard-Jones
1771 < simulation.},label={sch:LJFF}]
1772 < <OpenMD>
1692 <  <MetaData>
1693 < #include "argon.md"
1769 > Note that the parameters for a particular bend type are the same for
1770 > any bending triplet of the same atomic types (in the same or reversed
1771 > order).  Changing the AtomType in the Atom2 position will change the
1772 > matched bend types in the force field.
1773  
1774 < component{
1775 <  type = "Ar";
1776 <  nMol = 108;
1777 < }
1774 > \subsection{\label{section:ffTorsion}The TorsionTypes block}
1775 > The torsion potential for rotation of groups around a central bond can
1776 > often be represented with various cosine functions.  For two
1777 > tetrahedral ($sp^3$) carbons connected by a single bond, the torsion
1778 > potential might be
1779 > \begin{equation*}
1780 > V_{\text{torsion}} = \frac{v}{2} \left[ 1 + \cos( 3 \phi ) \right]
1781 > \end{equation*}
1782 > where $v$ is the barrier for going from {\em staggered} $\rightarrow$
1783 > {\em eclipsed} conformations, while for $sp^2$ carbons connected by a
1784 > double bond, the potential might be
1785 > \begin{equation*}
1786 > V_{\text{torsion}} = \frac{w}{2} \left[ 1 - \cos( 2 \phi ) \right]
1787 > \end{equation*}
1788 > where $w$ is the barrier for going from  {\em cis} $\rightarrow$ {\em
1789 >  trans} conformations.
1790  
1791 < forceField = "LJ";
1792 <  </MetaData>
1793 <  <Snapshot>     // not shown in this scheme
1794 <  </Snapshot>
1795 < </OpenMD>
1796 < \end{lstlisting}
1791 > A general torsion potentials can be represented as a cosine series of
1792 > the form:
1793 > \begin{equation}
1794 > V_{\text{torsion}}(\phi_{ijkl}) = c_1[1 + \cos \phi_{ijkl}]
1795 >        + c_2[1 - \cos(2\phi_{ijkl})]
1796 >        + c_3[1 + \cos(3\phi_{ijkl})],
1797 > \label{eq:origTorsionPot}
1798 > \end{equation}
1799 > where the angle $\phi_{ijkl}$ is defined
1800 > \begin{equation}
1801 > \cos\phi_{ijkl} = (\hat{\mathbf{r}}_{ij} \times \hat{\mathbf{r}}_{jk}) \cdot
1802 >        (\hat{\mathbf{r}}_{jk} \times \hat{\mathbf{r}}_{kl}).
1803 > \label{eq:torsPhi}
1804 > \end{equation}
1805 > Here, $\hat{\mathbf{r}}_{\alpha\beta}$ are the set of unit bond
1806 > vectors between atoms $i$, $j$, $k$, and $l$.
1807 >
1808 > For computational efficiency, OpenMD recasts torsion potential in the
1809 > method of {\sc charmm},\cite{Brooks83} in which the angle series is
1810 > converted to a power series of the form:
1811 > \begin{equation}
1812 >  V_{\text{torsion}}(\phi_{ijkl}) =  
1813 >  k_3 \cos^3 \phi_{ijkl} + k_2 \cos^2 \phi_{ijkl} + k_1 \cos \phi_{ijkl} + k_0,
1814 > \label{eq:torsionPot}
1815 > \end{equation}
1816 > where:
1817 > \begin{align*}
1818 > k_0 &= c_1 + 2 c_2 + c_3, \\
1819 > k_1 &= c_1 - 3c_3, \\
1820 > k_2 &= - 2 c_2, \\
1821 > k_3 &= 4 c_3.
1822 > \end{align*}
1823 > By recasting the potential as a power series, repeated trigonometric
1824 > evaluations are avoided during the calculation of the potential
1825 > energy.
1826  
1827 + Using this framework, OpenMD implements a variety of different
1828 + potential energy functions for torsions:
1829 + \begin{itemize}
1830 + \item {\tt Cubic}:
1831 + \begin{equation*}
1832 +  V_{\text{torsion}}(\phi) =  
1833 +  k_3 \cos^3 \phi + k_2 \cos^2 \phi + k_1 \cos \phi + k_0,
1834 + \end{equation*}
1835 + \item {\tt Quartic}:
1836 + \begin{equation*}
1837 +  V_{\text{torsion}}(\phi) =  k_4 \cos^4 \phi +
1838 +  k_3 \cos^3 \phi + k_2 \cos^2 \phi + k_1 \cos \phi + k_0,
1839 + \end{equation*}
1840 + \item {\tt Polynomial}:
1841 + \begin{equation*}
1842 + V_{\text{torsion}}(\phi) =  \sum_n k_n \cos^n \phi ,
1843 + \end{equation*}
1844 + \item {\tt Charmm}:
1845 + \begin{equation*}
1846 + V_{\text{torsion}}(\phi) = \sum_n K_n \left( 1 + cos(n
1847 +  \phi - \delta_n) \right),
1848 + \end{equation*}
1849 + \item {\tt Opls}:
1850 + \begin{equation*}
1851 +  V_{\text{torsion}}(\phi) =  \frac{1}{2} \left(v_1 (1 + \cos \phi) \right)
1852 +    + v_2 (1 - \cos 2 \phi) +  v_3 (1 + \cos 3 \phi),
1853 + \end{equation*}
1854 + \item {\tt Trappe}:\cite{Siepmann1998}
1855 + \begin{equation*}
1856 +  V_{\text{torsion}}(\phi) =  c_0 + c_1 (1 + \cos \phi) + c_2 (1 - \cos 2 \phi)  +
1857 +  c_3 (1 + \cos 3 \phi),
1858 + \end{equation*}
1859 + \item {\tt Harmonic}:
1860 + \begin{equation*}
1861 + V_{\text{torsion}}(\phi) =  \frac{d_0}{2} \left(\phi - \phi^0\right).
1862 + \end{equation*}
1863 + \end{itemize}
1864  
1865 < \section{\label{section:DUFF}Dipolar Unified-Atom Force Field}
1865 > Most torsion types don't require specific angle information in the
1866 > parameters since they are typically expressed in cosine polynomials.
1867 > {\tt Charmm} and {\tt Harmonic} torsions are a bit different.  {\tt
1868 >  Charmm} torsion types require a set of phase angles, $\delta_n$ that
1869 > are expressed in degrees, and associated periodicity numbers, $n$.
1870 > {\tt Harmonic} torsions have an equilibrium torsion angle, $\phi_0$
1871 > that is measured in degrees, while $d_0$ has units of
1872 > kcal/mol/degrees$^2$.  All other torsion parameters are measured in
1873 > units of kcal/mol.
1874  
1875 < The dipolar unified-atom force field ({\sc duff}) was developed to
1876 < simulate lipid bilayers. These types of simulations require a model
1877 < capable of forming bilayers, while still being sufficiently
1878 < computationally efficient to allow large systems ($\sim$100's of
1879 < phospholipids, $\sim$1000's of waters) to be simulated for long times
1880 < ($\sim$10's of nanoseconds). With this goal in mind, {\sc duff} has no
1881 < point charges. Charge-neutral distributions are replaced with dipoles,
1882 < while most atoms and groups of atoms are reduced to Lennard-Jones
1883 < interaction sites. This simplification reduces the length scale of
1884 < long range interactions from $\frac{1}{r}$ to $\frac{1}{r^3}$,
1885 < removing the need for the computationally expensive Ewald
1886 < sum. Instead, Verlet neighbor-lists and cutoff radii are used for the
1887 < dipolar interactions, and, if desired, a reaction field may be added
1888 < to mimic longer range interactions.
1875 > \begin{lstlisting}[caption={[An example of a TorsionTypes block.] A
1876 > simple example of a TorsionTypes block.  Energy constants are given in
1877 > kcal / mol, and when required by the form, $\delta$ angles are given
1878 > in degrees.},
1879 > label={sch:TorsionTypes}]
1880 > begin TorsionTypes
1881 > //Cubic
1882 > //Atom1 Atom2   Atom3   Atom4   Cubic   k3       k2        k1      k0  
1883 > CH2     CH2     CH2     CH2     Cubic   5.9602   -0.2568   -3.802  2.1586
1884 > CH2     CH      CH      CH2     Cubic   3.3254   -0.4215   -1.686  1.1661
1885 > //Trappe
1886 > //Atom1 Atom2   Atom3   Atom4   Trappe  c0       c1        c2      c3
1887 > CH3     CH2     CH2     SH      Trappe  0.10507  -0.10342  0.03668 0.60874    
1888 > //Charmm
1889 > //Atom1 Atom2   Atom3   Atom4   Charmm  Kchi     n    delta  [Kchi n delta]
1890 > CT      CT      CT      C       Charmm  0.156    3    0.00
1891 > OH      CT      CT      OH      Charmm  0.144    3    0.00    1.175 2  0
1892 > HC      CT      CM      CM      Charmm  1.150    1    0.00    0.38  3 180
1893 > //Quartic
1894 > //Atom1 Atom2   Atom3   Atom4   Quartic          k4    k3    k2    k1    k0
1895 > //Polynomial
1896 > //Atom1 Atom2   Atom3   Atom4   Polynomial  n Kn     [m  Km]
1897 > S       CH2     CH2     C       Polynomial  0 2.218   1  2.905  2 -3.136  3 -0.7313  4 6.272  5 -7.528
1898 > end TorsionTypes
1899 > \end{lstlisting}
1900  
1901 < As an example, lipid head-groups in {\sc duff} are represented as
1902 < point dipole interaction sites.  Placing a dipole at the head group's
1903 < 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}
1901 > Note that the parameters for a particular torsion type are the same
1902 > for any torsional quartet of the same atomic types (in the same or
1903 > reversed order).
1904  
1905 < \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}
1905 > \subsection{\label{section:ffInversion}The InversionTypes block}
1906  
1907 < A set of scalable parameters has been used to model the alkyl groups
1908 < with Lennard-Jones sites. For this, parameters from the TraPPE force
1909 < field of Siepmann \emph{et al.}\cite{Siepmann1998} have been
1910 < 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.
1907 > Inversion potentials are often added to force fields to enforce
1908 > planarity around $sp^2$-hybridized carbons or to correct vibrational
1909 > frequencies for umbrella-like vibrational modes for central atoms
1910 > bonded to exactly three satellite groups.
1911  
1912 < As is required by TraPPE, {\sc duff} also constrains all bonds to be
1913 < of fixed length. Typically, bond vibrations are the fastest motions in
1914 < a molecular dynamic simulation.  With these vibrations present, small
1915 < time steps between force evaluations must be used to ensure adequate
1916 < energy conservation in the bond degrees of freedom. By constraining
1917 < the bond lengths, larger time steps may be used when integrating the
1918 < equations of motion. A simulation using {\sc duff} is illustrated in
1919 < Scheme \ref{sch:DUFF}.
1912 > In OpenMD's version of an inversion, the central atom is entered {\it
1913 >  first} in each line in the {\tt InversionTypes} block. Note that
1914 > this is quite different than how other codes treat Improper torsional
1915 > potentials to mimic inversion behavior.  In some other widely-used
1916 > simulation packages, the central atom is treated as atom 3 in a
1917 > standard torsion form:
1918 > \begin{itemize}
1919 >  \item OpenMD:  I - (J - K - L)  (e.g. I is $sp^2$ hybridized carbon)
1920 >  \item AMBER:   I - J - K - L   (e.g. K is $sp^2$ hybridized carbon)
1921 > \end{itemize}
1922  
1923 < \begin{lstlisting}[float,caption={[Invocation of {\sc duff}]A portion
1924 < of a startup file showing a simulation utilizing {\sc
1925 < duff}},label={sch:DUFF}]  
1926 < <OpenMD>
1927 <  <MetaData>
1928 < #include "water.md"
1929 < #include "lipid.md"
1923 > The inversion angle itself is defined as:
1924 > \begin{equation}
1925 > \cos\omega_{i-jkl} = \left(\hat{\mathbf{r}}_{il} \times
1926 >  \hat{\mathbf{r}}_{ij}\right)\cdot\left( \hat{\mathbf{r}}_{il} \times
1927 >  \hat{\mathbf{r}}_{ik}\right)
1928 > \end{equation}
1929 > Here, $\hat{\mathbf{r}}_{\alpha\beta}$ are the set of unit bond
1930 > vectors between the central atoms $i$, and the satellite atoms $j$,
1931 > $k$, and $l$.  Note that other definitions of inversion angles are
1932 > possible, so users are encouraged to be particularly careful when
1933 > converting other force field files for use with OpenMD.
1934  
1935 < component{
1936 <  type = "simpleLipid_16";
1937 <  nMol = 60;
1938 < }
1939 <
1940 < component{
1941 <  type = "SSD_water";
1942 <  nMol = 1936;
1943 < }
1944 <
1945 < forceField = "DUFF";
1946 <  </MetaData>
1947 <  <Snapshot>     // not shown in this scheme
1948 <  </Snapshot>
1949 < </OpenMD>
1935 > There are many common ways to create planarity or umbrella behavior in
1936 > a potential energy function, and OpenMD implements a number of the
1937 > more common functions:
1938 > \begin{itemize}
1939 > \item {\tt ImproperCosine}:
1940 > \begin{equation*}
1941 > V_{\text{torsion}}(\omega) = \sum_n \frac{K_n}{2} \left( 1 + cos(n
1942 >  \omega - \delta_n) \right),
1943 > \end{equation*}
1944 > \item {\tt AmberImproper}:
1945 > \begin{equation*}
1946 >  V_{\text{torsion}}(\omega) =  \frac{v}{2} (1 - \cos\left(2 \left(\omega - \omega_0\right)\right),
1947 > \end{equation*}
1948 > \item {\tt Harmonic}:
1949 > \begin{equation*}
1950 > V_{\text{torsion}}(\omega) =  \frac{d}{2} \left(\omega - \omega_0\right).
1951 > \end{equation*}
1952 > \end{itemize}
1953 > \begin{lstlisting}[caption={[An example of an InversionTypes block.] A
1954 > simple example of a InversionTypes block.  Angles ($\delta_n$ and
1955 > $\omega_0$) angles are given in degrees, while energy parameters ($v,
1956 > K_n$) are given in kcal / mol.   The Harmonic Inversion type has a
1957 > force constant that must be given in kcal/mol/degrees$^2$.},
1958 > label={sch:InversionTypes}]
1959 > begin InversionTypes
1960 > //Harmonic
1961 > //Atom1 Atom2   Atom3   Atom4   Harmonic  d(kcal/mol/deg^2)  omega0
1962 > RCHar3  X       X       X       Harmonic  1.21876e-2         180.0
1963 > //AmberImproper
1964 > //Atom1 Atom2   Atom3   Atom4   AmberImproper   v(kcal/mol)
1965 > C       CT      N       O       AmberImproper   10.500000
1966 > CA      CA      CA      CT      AmberImproper   1.100000
1967 > //ImproperCosine
1968 > //Atom1 Atom2   Atom3   Atom4   ImproperCosine  Kn  n  delta_n  [Kn n delta_n]
1969 > end InversionTypes
1970   \end{lstlisting}
1971  
1972 + \section{\label{section::ffLongRange}Long Range Interactions}
1973  
1974 + Calculating the long-range (non-bonded) potential involves a sum over
1975 + all pairs of atoms (except for those atoms which are involved in a
1976 + bond, bend, or torsion with each other).  Many of these interactions
1977 + can be inferred from the AtomTypes,
1978  
1979 < The cross potential between molecules $I$ and $J$,
1980 < $V^{IJ}_{\text{Cross}}$, 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.
1979 > \subsection{\label{section:ffNBinteraction}The NonBondedInteractions
1980 >  block}
1981  
1982 + The user might want like to specify explicit or special interactions
1983 + that override the default non-bodned interactions that are inferred
1984 + from the AtomTypes.  To do this, OpenMD implements a
1985 + NonBondedInteractions block to allow the user to specify the following
1986 + (pair-wise) non-bonded interactions:
1987  
1988 < \section{\label{section:WATER}The {\sc water} Force Field}
1988 > \begin{itemize}
1989 > \item {\tt LennardJones}:
1990 > \begin{equation*}
1991 > V_{\text{NB}}(r) = 4 \epsilon_{ij} \left(
1992 >  \left(\frac{\sigma_{ij}}{r} \right)^{12} -
1993 >  \left(\frac{\sigma_{ij}}{r} \right)^{6} \right),
1994 > \end{equation*}
1995 > \item {\tt ShiftedMorse}:
1996 > \begin{equation*}
1997 > V_{\text{NB}}(r) = D_{ij} \left( e^{-2 \beta_{ij} (r -
1998 >     r^0)} - 2 e^{- \beta_{ij} (r -
1999 >     r^0)} \right),
2000 > \end{equation*}
2001 > \item {\tt RepulsiveMorse}:
2002 > \begin{equation*}
2003 > V_{\text{NB}}(r) = D_{ij} \left( e^{-2 \beta_{ij} (r -
2004 >     r^0)} \right),
2005 > \end{equation*}
2006 > \item {\tt RepulsivePower}:
2007 > \begin{equation*}
2008 >  V_{\text{NB}}(r) = \epsilon_{ij}
2009 >  \left(\frac{\sigma_{ij}}{r} \right)^{n_{ij}}.
2010 > \end{equation*}
2011 > \end{itemize}
2012  
2013 < In addition to the {\sc duff} force field's solvent description, a
2014 < separate {\sc water} force field has been included for simulating most
2015 < of the common rigid-body water models. This force field includes the
2016 < simple and point-dipolar models (SSD, SSD1, SSD/E, SSD/RF, and DPD
2017 < water), as well as the common charge-based models (SPC, SPC/E, TIP3P,
2018 < TIP4P, and
2019 < 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.
2013 > \begin{lstlisting}[caption={[An example of a NonBondedInteractions block.] A
2014 > simple example of a NonBondedInteractions block. Distances ($\sigma,
2015 > r_0$) are given in \AA, while energies ($\epsilon, D0$) are in
2016 > kcal/mol.  The Morse potentials have an additional parameter $\beta_0$
2017 > which is in units of \AA$^{-1}$.},
2018 > label={sch:InversionTypes}]
2019 > begin NonBondedInteractions
2020  
2021 + //Lennard-Jones
2022 + //Atom1 Atom2   LennardJones    sigma  epsilon
2023 + Au      CH3     LennardJones    3.54   0.2146
2024 + Au      CH2     LennardJones    3.54   0.1749
2025 + Au      CH      LennardJones    3.54   0.1749
2026 + Au      S       LennardJones    2.40   8.465
2027  
2028 < \section{\label{section:sc}The Sutton-Chen Force Field}
2028 > //Shifted Morse
2029 > //Atom1 Atom2   ShiftedMorse    r0     D0       beta0
2030 > Au      O_SPCE  ShiftedMorse    3.70   0.0424   0.769
2031  
2032 + //Repulsive Morse
2033 + //Atom1 Atom2   RepulsiveMorse  r0     D0       beta0
2034 + Au      H_SPCE  RepulsiveMorse  -1.00  0.00850  0.769
2035  
2036 < \section{\label{section:clay}The CLAY force field}
2037 <
2038 < The {\sc clay} force field is based on an ionic (nonbonded)
2039 < description of the metal-oxygen interactions associated with hydrated
2040 < phases. All atoms are represented as point charges and are allowed
2041 < complete translational freedom. Metal-oxygen interactions are based on
1848 < a simple Lennard-Jones potential combined with electrostatics. The
1849 < empirical parameters were optimized by Cygan {\it et
1850 < al.}\cite{Cygan04} on the basis of known mineral structures, and
1851 < partial atomic charges were derived from periodic DFT quantum chemical
1852 < calculations of simple oxide, hydroxide, and oxyhydroxide model
1853 < compounds with well-defined structures.
2036 > //Repulsive Power
2037 > //Atom1 Atom2   RepulsivePower   sigma    epsilon    n
2038 > Au      ON      RepulsivePower   3.47005  0.186208   11
2039 > Au      NO      RepulsivePower   3.53955  0.168629   11
2040 > end NonBondedInteractions
2041 > \end{lstlisting}
2042  
1855
2043   \section{\label{section:electrostatics}Electrostatics}
2044  
2045   To aid in performing simulations in more traditional force fields, we
# Line 1982 | Line 2169 | the shifted potential (eq. (\ref{eq:SPPot})) becomes
2169   \end{equation}
2170   the shifted potential (eq. (\ref{eq:SPPot})) becomes
2171   \begin{equation}
2172 < V_{\textrm{DSP}}(r) = q_iq_j\left(\frac{\textrm{erfc}\left(\alpha r\right)}{r}-\
1986 < frac{\textrm{erfc}\left(\alpha R_\textrm{c}\right)}{R_\textrm{c}}\right) \quad r
2172 > V_{\textrm{DSP}}(r) = q_iq_j\left(\frac{\textrm{erfc}\left(\alpha r\right)}{r}-\frac{\textrm{erfc}\left(\alpha R_\textrm{c}\right)}{R_\textrm{c}}\right) \quad r
2173   \leqslant R_\textrm{c},
2174   \label{eq:DSPPot}
2175   \end{equation}
# Line 2028 | Line 2214 | this reason, the default electrostatic summation metho
2214   {\sc OpenMD} is the DSF (Eq. \ref{eq:DSFPot}) with a damping parameter
2215   ($\alpha$) that is set algorithmically from the cutoff radius.
2216  
2217 +
2218 + \section{\label{section:cutoffGroups}Switching Functions}
2219 +
2220 + If done poorly, calculating the the long-range interactions for $N$
2221 + atoms would involve $N(N-1)/2$ evaluations of atomic distances.  To
2222 + reduce the number of distance evaluations between pairs of atoms, {\sc
2223 +  OpenMD} allows the use of switched cutoffs with Verlet neighbor
2224 + lists.\cite{Allen87} Neutral groups which contain charges will exhibit
2225 + pathological forces unless the cutoff is applied to the neutral groups
2226 + evenly instead of to the individual atoms.\cite{leach01:mm} {\sc
2227 +  OpenMD} allows users to specify cutoff groups which may contain an
2228 + arbitrary number of atoms in the molecule.  Atoms in a cutoff group
2229 + are treated as a single unit for the evaluation of the switching
2230 + function:
2231 + \begin{equation}
2232 + V_{\mathrm{long-range}} = \sum_{a} \sum_{b>a} s(r_{ab}) \sum_{i \in a} \sum_{j \in b} V_{ij}(r_{ij}),
2233 + \end{equation}
2234 + where $r_{ab}$ is the distance between the centers of mass of the two
2235 + cutoff groups ($a$ and $b$).
2236 +
2237 + The sums over $a$ and $b$ are over the cutoff groups that are present
2238 + in the simulation.  Atoms which are not explicitly defined as members
2239 + of a {\tt cutoffGroup} are treated as a group consisting of only one
2240 + atom.  The switching function, $s(r)$ is the standard cubic switching
2241 + function,
2242 + \begin{equation}
2243 + S(r) =
2244 +        \begin{cases}
2245 +        1 & \text{if $r \le r_{\text{sw}}$},\\
2246 +        \frac{(r_{\text{cut}} + 2r - 3r_{\text{sw}})(r_{\text{cut}} - r)^2}
2247 +        {(r_{\text{cut}} - r_{\text{sw}})^3}
2248 +        & \text{if $r_{\text{sw}} < r \le r_{\text{cut}}$}, \\
2249 +        0 & \text{if $r > r_{\text{cut}}$.}
2250 +        \end{cases}
2251 + \label{eq:dipoleSwitching}
2252 + \end{equation}
2253 + Here, $r_{\text{sw}}$ is the {\tt switchingRadius}, or the distance
2254 + beyond which interactions are reduced, and $r_{\text{cut}}$ is the
2255 + {\tt cutoffRadius}, or the distance at which interactions are
2256 + truncated.
2257 +
2258 + Users of {\sc OpenMD} do not need to specify the {\tt cutoffRadius} or
2259 + {\tt switchingRadius}.  In simulations containing only Lennard-Jones
2260 + atoms, the cutoff radius has a default value of $2.5\sigma_{ii}$,
2261 + where $\sigma_{ii}$ is the largest Lennard-Jones length parameter
2262 + present in the simulation.  In simulations containing charged or
2263 + dipolar atoms, the default cutoff radius is $15 \mbox{\AA}$.  
2264 +
2265 + The {\tt switchingRadius} is set to a default value of 95\% of the
2266 + {\tt cutoffRadius}.  In the special case of a simulation containing
2267 + {\it only} Lennard-Jones atoms, the default switching radius takes the
2268 + same value as the cutoff radius, and {\sc OpenMD} will use a shifted
2269 + potential to remove discontinuities in the potential at the cutoff.
2270 + Both radii may be specified in the meta-data file.
2271 +
2272 +
2273 + \section{\label{section:WATER}The {\sc water} Force Field}
2274 +
2275 + In addition to the {\sc duff} force field's solvent description, a
2276 + separate {\sc water} force field has been included for simulating most
2277 + of the common rigid-body water models. This force field includes the
2278 + simple and point-dipolar models (SSD, SSD1, SSD/E, SSD/RF, and DPD
2279 + water), as well as the common charge-based models (SPC, SPC/E, TIP3P,
2280 + TIP4P, and
2281 + TIP5P).\cite{liu96:new_model,Ichiye03,fennell04,Marrink01,Berendsen81,Berendsen87,Jorgensen83,Mahoney00}
2282 + In order to handle these models, charge-charge interactions were
2283 + included in the force-loop:
2284 + \begin{equation}
2285 + V_{\text{charge}}(r_{ij}) = \sum_{ij}\frac{q_iq_je^2}{r_{ij}},
2286 + \end{equation}
2287 + where $q$ represents the charge on particle $i$ or $j$, and $e$ is the
2288 + charge of an electron in Coulombs. The charge-charge interaction
2289 + support is rudimentary in the current version of {\sc OpenMD}.  As with
2290 + the other pair interactions, charges can be simulated with a pure
2291 + cutoff or a reaction field.  The various methods for performing the
2292 + Ewald summation have not yet been included.  The {\sc water} force
2293 + field can be easily expanded through modification of the {\sc water}
2294 + force field file ({\tt WATER.frc}). By adding atom types and inserting
2295 + the appropriate parameters, it is possible to extend the force field
2296 + to handle rigid molecules other than water.
2297 +
2298 +
2299 +
2300   \section{\label{section:pbc}Periodic Boundary Conditions}
2301  
2302   \newcommand{\roundme}{\operatorname{round}}

Diff Legend

Removed lines
+ Added lines
< Changed lines
> Changed lines