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

Comparing trunk/oopsePaper/oopsePaper.tex (file contents):
Revision 1168 by mmeineke, Wed May 12 19:32:47 2004 UTC vs.
Revision 3389 by gezelter, Tue Apr 29 17:50:49 2008 UTC

# Line 1 | Line 1
1   \documentclass[11pt]{article}
2   \usepackage{amsmath}
3 < \usepackage{amssymb}
3 > \usepackage{amssymb}
4   \usepackage{endfloat}
5   \usepackage{listings}
6 < \usepackage{palatino}
6 > \usepackage{berkeley}
7   \usepackage{graphicx}
8   \usepackage[ref]{overcite}
9   \usepackage{setspace}
# Line 22 | Line 22
22          xleftmargin=0.5in, xrightmargin=0.5in,captionpos=b, %
23          abovecaptionskip=0.5cm, belowcaptionskip=0.5cm}
24   \renewcommand{\lstlistingname}{Scheme}
25 < \title{{\sc oopse}: An Open Source Object-Oriented Parallel Simulation
25 > \title{{\sc oopse}: An Object-Oriented Parallel Simulation
26   Engine for Molecular Dynamics}
27  
28   \author{Matthew A. Meineke, Charles F. Vardeman II, Teng Lin,\\
# Line 31 | Line 31 | Notre Dame, Indiana 46556}
31   University of Notre Dame\\
32   Notre Dame, Indiana 46556}
33  
34 < \date{\today}
34 > \date{September 20, 2004}
35   \maketitle
36  
37   \begin{abstract}
38 < We detail the capabilities of a new open-source parallel simulation
39 < progrm for MD ({\sc oopse}) that can work with  atom types that are missing from other popular packages.  In
40 < particular, {\sc oopse} is capable of performing efficient orientational
41 < dynamics on dipolar or rigid body systems, and it can handle simulations of metallic
42 < systems using the embedded atom method ({\sc eam}).
38 > {\sc oopse} is a new molecular dynamics simulation program which is
39 > capable of efficiently integrating equations of motion for atom types
40 > with orientational degrees of freedom (e.g. ``sticky'' atoms and point
41 > dipoles).  Transition metals can also be simulated using the embedded
42 > atom method ({\sc eam}) potential included in the code.  Parallel
43 > simulations are carried out using the force-based decomposition
44 > method.  Simulations are specified using a very simple C-based
45 > meta-data language.  A number of advanced integrators are included,
46 > and the basic integrator for orientational dynamics provides
47 > substantial improvements over older quaternion-based schemes.
48   \end{abstract}
49  
50   \section{\label{sec:intro}Introduction}
51  
52 < When choosing to simulate a chemical system with molecular dynamics,
53 < there are a variety of options available. For simple systems, one
54 < might consider writing one's own programming code. However, as systems
55 < grow larger and more complex, building and maintaining code for the
56 < simulations becomes a time consuming task. In such cases it is usually
57 < more convenient for a researcher to turn to pre-existing simulation
58 < packages. These packages, such as {\sc amber}\cite{pearlman:1995} and
59 < {\sc charmm}\cite{Brooks83}, provide powerful tools for researchers to
60 < conduct simulations of their systems without spending their time
61 < developing a code base to conduct their research. This then frees them
62 < to perhaps explore experimental analogues to their models.
52 > There are a number of excellent molecular dynamics packages available
53 > to the chemical physics
54 > community.\cite{Brooks83,MacKerell98,pearlman:1995,Gromacs,Gromacs3,DL_POLY,Tinker,Paradyn,namd,macromodel}
55 > All of these packages are stable, polished programs which solve many
56 > problems of interest.  Most are now capable of performing molecular
57 > dynamics simulations on parallel computers.  Some have source code
58 > which is freely available to the entire scientific community.  Few,
59 > however, are capable of efficiently integrating the equations of
60 > motion for atom types with orientational degrees of freedom
61 > (e.g. point dipoles, and ``sticky'' atoms).  And only one of the
62 > programs referenced can handle transition metal force fields like the
63 > Embedded Atom Method ({\sc eam}).  The direction our research program
64 > has taken us now involves the use of atoms with orientational degrees
65 > of freedom as well as transition metals.  Since these simulation
66 > methods may be of some use to other researchers, we have decided to
67 > release our program (and all related source code) to the scientific
68 > community.
69  
70 < Despite their utility, problems with these packages arise when
71 < researchers try to develop techniques or energetic models that the
72 < code was not originally designed to simulate. Examples of techniques
73 < and energetics not commonly implemented include; dipole-dipole
74 < interactions, rigid body dynamics, and metallic potentials. When faced
75 < with these obstacles, a researcher must either develop their own code
76 < or license and extend one of the commercial packages. What we have
77 < elected to do is develop a body of simulation code capable of
78 < implementing the types of models upon which our research is based.
70 > This paper communicates the algorithmic details of our program, which
71 > we have been calling the Object-Oriented Parallel Simulation Engine
72 > (i.e. {\sc oopse}).  We have structured this paper to first discuss
73 > the underlying concepts in this simulation package
74 > (Sec. \ref{oopseSec:IOfiles}).  The empirical energy functions
75 > implemented are discussed in Sec.~\ref{oopseSec:empiricalEnergy}.
76 > Sec.~\ref{oopseSec:mechanics} describes the various Molecular Dynamics
77 > algorithms {\sc oopse} implements in the integration of Hamilton's
78 > equations of motion.  Program design considerations for parallel
79 > computing are presented in
80 > Sec.~\ref{oopseSec:parallelization}. Concluding remarks are presented
81 > in Sec.~\ref{oopseSec:conclusion}.
82  
83 < In developing {\sc oopse}, we have adhered to the precepts of Open
70 < Source development, and are releasing our source code with a
71 < permissive license. It is our intent that by doing so, other
72 < researchers might benefit from our work, and add their own
73 < contributions to the package. The license under which {\sc oopse} is
74 < distributed allows any researcher to download and modify the source
75 < code for their own use. In this way further development of {\sc oopse}
76 < is not limited to only the models of interest to ourselves, but also
77 < those of the community of scientists who contribute back to the
78 < project.
83 > \section{\label{oopseSec:IOfiles}Concepts \& Files}
84  
85 < We have structured this paper to first discuss the empirical energy
86 < functions that {\sc oopse } implements in
87 < Sec.~\ref{oopseSec:empiricalEnergy}. Following that is a discussion of
88 < the various input and output files associated with the package
89 < (Sec.~\ref{oopseSec:IOfiles}). Sec.~\ref{oopseSec:mechanics}
90 < elucidates the various Molecular Dynamics algorithms {\sc oopse}
86 < implements in the integration of the Newtonian equations of
87 < motion.  Program design
88 < considerations are presented in Sec.~\ref{oopseSec:design}. And
89 < lastly, Sec.~\ref{oopseSec:conclusion} concludes the chapter.
85 > A simulation in {\sc oopse} is built using a few fundamental
86 > conceptual building blocks most of which are chemically intuitive.
87 > The basic unit of a simulation is an {\tt atom}.  The parameters
88 > describing an {\tt atom} have been generalized to make it as flexible
89 > as possible; this means that in addition to translational degrees of
90 > freedom, {\tt Atoms} may also have {\it orientational} degrees of freedom.
91  
92 < \section{\label{oopseSec:IOfiles}Concepts \& Files}
92 > The fundamental (static) properties of {\tt atoms} are defined by the
93 > {\tt forceField} chosen for the simulation.  The atomic properties
94 > specified by a {\tt forceField} might include (but are not limited to)
95 > charge, $\sigma$ and $\epsilon$ values for Lennard-Jones interactions,
96 > the strength of the dipole moment ($\mu$), the mass, and the moments
97 > of inertia.  Other more complicated properties of atoms might also be
98 > specified by the {\tt forceField}.
99  
100 < \subsection{{\sc bass} and Model Files}
100 > {\tt Atoms} can be grouped together in many ways.  A {\tt rigidBody}
101 > contains atoms that exert no forces on one another and which move as a
102 > single rigid unit.  A {\tt cutoffGroup} may contain atoms which
103 > function together as a (rigid {\it or} non-rigid) unit for potential
104 > energy calculations,
105 > \begin{equation}
106 > V_{ab} = s(r_{ab}) \sum_{i \in a} \sum_{j \in b} V_{ij}(r_{ij})
107 > \end{equation}
108 > Here, $a$ and $b$ are two {\tt cutoffGroups} containing multiple atoms
109 > ($a = \left\{i\right\}$ and $b = \left\{j\right\}$).  $s(r_{ab})$ is a
110 > generalized switching function which insures that the atoms in the two
111 > {\tt cutoffGroups} are treated identically as the two groups enter or
112 > leave an interaction region.
113  
114 < Every {\sc oopse} simulation begins with a Bizarre Atom Simulation
115 < Syntax ({\sc bass}) file. {\sc bass} is a script syntax that is parsed
116 < by {\sc oopse} at runtime. The {\sc bass} file allows for the user to
117 < completely describe the system they wish to simulate, as well as tailor
99 < {\sc oopse}'s behavior during the simulation. {\sc bass} files are
100 < denoted with the extension
101 < \texttt{.bass}, an example file is shown in
102 < Scheme~\ref{sch:bassExample}.
114 > {\tt Atoms} may also be grouped in more traditional ways into {\tt
115 > bonds}, {\tt bends}, and {\tt torsions}.  These groupings allow the
116 > correct choice of interaction parameters for short-range interactions
117 > to be chosen from the definitions in the {\tt forceField}.
118  
119 < \begin{lstlisting}[float,caption={[An example of a complete {\sc bass} file] An example showing a complete {\sc bass} file.},label={sch:bassExample}]
119 > All of these groups of {\tt atoms} are brought together in the {\tt
120 > molecule}, which is the fundamental structure for setting up and {\sc
121 > oopse} simulation.  {\tt Molecules} contain lists of {\tt atoms}
122 > followed by listings of the other atomic groupings ({\tt bonds}, {\tt
123 > bends}, {\tt torsions}, {\tt rigidBodies}, and {\tt cutoffGroups})
124 > which relate the atoms to one another.
125  
126 + Simulations often involve heterogeneous collections of molecules.  To
127 + specify a mixture of {\tt molecule} types, {\sc oopse} uses {\tt
128 + components}.  Even simulations containing only one type of molecule
129 + must specify a single {\tt component}.
130 +
131 + Starting a simulation requires two types of information: {\it
132 + meta-data}, which describes the types of objects present in the
133 + simulation, and {\it configuration} information, which describes the
134 + initial state of these objects.  The meta-data is given to {\sc oopse}
135 + using a C-based syntax that is parsed at the beginning of the
136 + simulation.  Configuration information is specified using an extended
137 + XYZ file format.  Both the meta-data and configuration file formats
138 + are described in the following sections.
139 +
140 + \subsection{Meta-data Files}
141 +
142 + {\sc oopse} uses a C-based script syntax to parse the meta-data files
143 + at run time.  These files allow the user to completely describe the
144 + system they wish to simulate, as well as tailor {\sc oopse}'s behavior
145 + during the simulation.  Meta-data files are typically denoted with the
146 + extension {\tt .md} (which can stand for Meta-Data or Molecular
147 + Dynamics or Molecule Definition depending on the user's mood). An
148 + example meta-data file is shown in Scheme~\ref{sch:mdExample}.
149 +
150 + \begin{lstlisting}[float,caption={[An example of a complete meta-data
151 + file] An example showing a complete meta-data
152 + file.},label={sch:mdExample}]
153 +
154   molecule{
155    name = "Ar";
156    nAtoms = 1;
# Line 118 | Line 166 | initialConfig = "./argon.init";
166    nMol = 108;
167   }
168  
169 < initialConfig = "./argon.init";
169 > initialConfig = "./argon.in";
170  
171   forceField = "LJ";
172   ensemble = "NVE"; // specify the simulation ensemble
# Line 129 | Line 177 | Within the \texttt{.bass} file it is necessary to prov
177  
178   \end{lstlisting}
179  
180 < Within the \texttt{.bass} file it is necessary to provide a complete
180 > Within the meta-data file it is necessary to provide a complete
181   description of the molecule before it is actually placed in the
182 < simulation. The {\sc bass} syntax was originally developed with this
183 < goal in mind, and allows for the specification of all the atoms in a
184 < molecular prototype, as well as any bonds, bends, or torsions. These
185 < descriptions can become lengthy for complex molecules, and it would be
186 < inconvenient to duplicate the simulation at the beginning of each {\sc
187 < bass} script. Addressing this issue {\sc bass} allows for the
188 < inclusion of model files at the top of a \texttt{.bass} file. These
189 < model files, denoted with the \texttt{.mdl} extension, allow the user
190 < to describe a molecular prototype once, then simply include it into
143 < each simulation containing that molecule. Returning to the example in
144 < Scheme~\ref{sch:bassExample}, the \texttt{.mdl} file's contents would
145 < be Scheme~\ref{sch:mdlExample}, and the new \texttt{.bass} file would
146 < become Scheme~\ref{sch:bassExPrime}.
182 > simulation. {\sc oopse}'s meta-data syntax was originally developed
183 > with this goal in mind, and allows for the use of {\it include files}
184 > to specify all atoms in a molecular prototype, as well as any bonds,
185 > bends, or torsions.  Include files allow the user to describe a
186 > molecular prototype once, then simply include it into each simulation
187 > containing that molecule. Returning to the example in
188 > Scheme~\ref{sch:mdExample}, the include file's contents would be
189 > Scheme~\ref{sch:mdIncludeExample}, and the new meta-data file would
190 > become Scheme~\ref{sch:mdExPrime}.
191  
192 < \begin{lstlisting}[float,caption={An example \texttt{.mdl} file.},label={sch:mdlExample}]
192 > \begin{lstlisting}[float,caption={An example molecule definition in an
193 > include file.},label={sch:mdIncludeExample}]
194  
195   molecule{
196    name = "Ar";
# Line 158 | Line 203 | molecule{
203  
204   \end{lstlisting}
205  
206 < \begin{lstlisting}[float,caption={Revised {\sc bass} example.},label={sch:bassExPrime}]
206 > \begin{lstlisting}[float,caption={Revised meta-data example.},label={sch:mdExPrime}]
207  
208 < #include "argon.mdl"
208 > #include "argon.md"
209  
210   nComponents = 1;
211   component{
# Line 168 | Line 213 | initialConfig = "./argon.init";
213    nMol = 108;
214   }
215  
216 < initialConfig = "./argon.init";
216 > initialConfig = "./argon.in";
217  
218   forceField = "LJ";
219   ensemble = "NVE";
# Line 179 | Line 224 | statusTime = 50;
224  
225   \end{lstlisting}
226  
227 < \subsection{\label{oopseSec:atomsMolecules}Atoms, Molecules and Rigid Bodies}
227 > \subsection{\label{oopseSec:atomsMolecules}Atoms, Molecules, and other
228 > ways of grouping atoms}
229  
230 < The basic unit of an {\sc oopse} simulation is the atom. The
231 < parameters describing the atom are generalized to make the atom as
232 < flexible a representation as possible. They may represent specific
233 < atoms of an element, or be used for collections of atoms such as
234 < methyl and carbonyl groups. The atoms are also capable of having
235 < directional components associated with them (\emph{e.g.}~permanent
236 < dipoles). Charges, permanent dipoles, and Lennard-Jones parameters for
237 < a given atom type are set in the force field parameter files.
238 <
239 < Atoms can be collected into secondary structures such as rigid bodies
240 < or molecules. The molecule is a way for {\sc oopse} to keep track of
241 < the atoms in a simulation in logical manner. Molecular units store the
196 < identities of all the atoms and rigid bodies associated with
197 < themselves, and are responsible for the evaluation of their own
198 < internal interactions (\emph{i.e.}~bonds, bends, and torsions). Scheme
199 < \ref{sch:mdlExample} shows how one creates a molecule in a ``model'' or
200 < \texttt{.mdl} file. The position of the atoms given in the
201 < declaration are relative to the origin of the molecule, and is used
202 < when creating a system containing the molecule.
230 > As mentioned above, the fundamental unit for an {\sc oopse} simulation
231 > is the {\tt atom}.  Atoms can be collected into secondary structures
232 > such as {\tt rigidBodies}, {\tt cutoffGroups}, or {\tt molecules}. The
233 > {\tt molecule} is a way for {\sc oopse} to keep track of the atoms in
234 > a simulation in logical manner. Molecular units store the identities
235 > of all the atoms and rigid bodies associated with themselves, and they
236 > are responsible for the evaluation of their own internal interactions
237 > (\emph{i.e.}~bonds, bends, and torsions). Scheme
238 > \ref{sch:mdIncludeExample} shows how one creates a molecule in an
239 > included meta-data file. The positions of the atoms given in the
240 > declaration are relative to the origin of the molecule, and the origin
241 > is used when creating a system containing the molecule.
242  
243 < As stated previously, one of the features that sets {\sc oopse} apart
244 < from most of the current molecular simulation packages is the ability
245 < to handle rigid body dynamics. Rigid bodies are non-spherical
246 < particles or collections of particles that have a constant internal
243 > One of the features that sets {\sc oopse} apart from most of the
244 > current molecular simulation packages is the ability to handle rigid
245 > body dynamics. Rigid bodies are non-spherical particles or collections
246 > of particles (e.g. $\mbox{C}_{60}$) that have a constant internal
247   potential and move collectively.\cite{Goldstein01} They are not
248   included in most simulation packages because of the algorithmic
249 < complexity involved in propagating orientational degrees of
250 < freedom. Until recently, integrators which propagate orientational
251 < motion have had energy conservation problems when compared to  those available for translational
252 < motion.
249 > complexity involved in propagating orientational degrees of freedom.
250 > Integrators which propagate orientational motion with an acceptable
251 > level of energy conservation for molecular dynamics are relatively
252 > new inventions.  
253  
254   Moving a rigid body involves determination of both the force and
255   torque applied by the surroundings, which directly affect the
# Line 220 | Line 259 | the rigid body. The torque on rigid body $i$ is
259   the rigid body is simply the sum of these external forces.
260   Accumulation of the total torque on the rigid body is more complex
261   than the force because the torque is applied to the center of mass of
262 < the rigid body. The torque on rigid body $i$ is
262 > the rigid body. The space-fixed torque on rigid body $i$ is
263   \begin{equation}
264   \boldsymbol{\tau}_i=
265          \sum_{a}\biggl[(\mathbf{r}_{ia}-\mathbf{r}_i)\times \mathbf{f}_{ia}
# Line 242 | Line 281 | arithmetic operations involving the four quaternions (
281   $\psi$.\cite{Goldstein01} In order to avoid numerical instabilities
282   inherent in using the Euler angles, the four parameter ``quaternion''
283   scheme is often used. The elements of $\mathsf{A}$ can be expressed as
284 < arithmetic operations involving the four quaternions ($q_0, q_1, q_2,$
285 < and $q_3$).\cite{allen87:csl} Use of quaternions also leads to
284 > arithmetic operations involving the four quaternions ($q_w, q_x, q_y,$
285 > and $q_z$).\cite{allen87:csl} Use of quaternions also leads to
286   performance enhancements, particularly for very small
287   systems.\cite{Evans77}
288  
289 < {\sc oopse} utilizes a relatively new scheme that propagates the
290 < entire nine parameter rotation matrix. Further discussion
291 < on this choice can be found in Sec.~\ref{oopseSec:integrate}. An example
292 < definition of a rigid body can be seen in Scheme
289 > Rather than use one of the previously stated methods, {\sc oopse}
290 > utilizes a relatively new scheme that propagates the entire nine
291 > parameter rotation matrix. Further discussion on this choice can be
292 > found in Sec.~\ref{oopseSec:integrate}. An example definition of a
293 > rigid body can be seen in Scheme
294   \ref{sch:rigidBody}.
295  
296 < \begin{lstlisting}[float,caption={[Defining rigid bodies]A sample definition of a molecule containing a rigid body},label={sch:rigidBody}]
296 > \begin{lstlisting}[float,caption={[Defining rigid bodies]A sample
297 > definition of a molecule containing a rigid body and a cutoff
298 > group},label={sch:rigidBody}]
299   molecule{
300    name = "TIP3P";
301    nAtoms = 3;
# Line 275 | Line 317 | molecule{
317      nMembers = 3;
318      members(0, 1, 2);
319    }
320 +
321 +  nCutoffGroups = 1;
322 +  cutoffGroup[0]{
323 +    nMembers = 3;
324 +    members(0, 1, 2);
325 +  }
326   }
327   \end{lstlisting}
328  
329 < \subsection{\label{sec:miscConcepts}Putting a Script Together}
329 > \subsection{\label{sec:miscConcepts}Creating a Metadata File}
330  
331 < The actual creation of a {\sc bass} script requires several key components. The first  part of the script needs to be the declaration of all of the molecule prototypes used in the simulation. This is typically done through the inclusion of {\tt .mdl} files. Only the molecules actually present in the simulation need to be declared, however {\sc bass} allows for the declaration of more molecules than are needed. This gives the user the ability to build up a library of commonly used molecules into a single {\tt .mdl} file.
331 > The actual creation of a metadata file requires several key
332 > components. The first part of the file needs to be the declaration of
333 > all of the molecule prototypes used in the simulation. This is
334 > typically done through included meta-data files. Only the molecules
335 > actually present in the simulation need to be declared; however, {\sc
336 > oopse} allows for the declaration of more molecules than are
337 > needed. This gives the user the ability to build up a library of
338 > commonly used molecules into a single include file.
339  
340 < Once all prototypes are declared, the ordering of the rest of the script is less stringent. Typically, the next to follow the molecular prototypes are the component statements. These statements specify which molecules are present within the simulation. The number of components must first be declared before the first component block statement (an example is seen in Sch.~\ref{sch:bassExPrime}).  The component blocks tell {\sc oopse} the number of molecules that will be in the simulation, and the order in which the components blocks are declared sets the ordering of the real atoms within the simulation as well as in the output files.
340 > Once all prototypes are declared, the ordering of the rest of the
341 > script is less stringent.  The molecular composition of the simulation
342 > is specified with {\tt component} statements. Each different type of
343 > molecule present in the simulation is considered a separate
344 > component. The number of components must be declared before the first
345 > component block statement (an example is shown in
346 > Sch.~\ref{sch:mdExPrime}).  The component blocks tell {\sc oopse} the
347 > number of molecules that will be in the simulation, and the order in
348 > which the components blocks are declared sets the ordering of the real
349 > atoms in the configuration file as well as in the output files. The
350 > remainder of the script then sets the various simulation parameters
351 > for the system of interest.
352  
353 < The remainder of the script then sets the various simulation parameters for the system of interest. The required set of parameters that must be present in all simulations is given in Table~\ref{table:reqParams}.  The {\tt ensemble} statement is responsible for selecting the integration method used for the calculation of the equations of motion. An in depth discussion of the various methods available in {\sc oopse} can be found in Sec.~\ref{oopseSec:mechanics}. The {\tt forceField} statement is important for the selection of which forces will be used in the course of the simulation. {\sc oopse} supports several force fields, as outlined in Sec.~\ref{oopseSec:empericalEnergy}. The force fields are interchangeable between simulations, with the only requirement being that all atoms needed by the simulation are defined within the selected force field. The time step between force evaluations is set with the {\tt dt} parameter, and {\tt runTime} will set the time length of the simulation. Note, that {\tt runTime} is an absolute time, meaning if the simulation is started at t = 10.0~ns with a {\tt runTime} of 25.0~ns, the simulation will only run for an additional 15.0~ns. The final required parameter, is the {\tt initialConfig} statement. This will set the initial coordinates for the system, as well as the initial time if the {\tt useInitalTime = true;} flag is given. The format of the file specified in {\tt initialConfig}, is given in Sec.~\ref{oopseSec:coordFiles}. Additional parameters are summarized in Table~\ref{table:genParams}.
353 > The required set of parameters that must be present in all simulations
354 > is given in Table~\ref{table:reqParams}.  Since the user can use {\sc
355 > oopse} to perform energy minimizations as well as molecular dynamics
356 > simulations, one of the {\tt minimizer} or {\tt ensemble} keywords
357 > must be present.  The {\tt ensemble} keyword is responsible for
358 > selecting the integration method used for the calculation of the
359 > equations of motion. An in depth discussion of the various methods
360 > available in {\sc oopse} can be found in
361 > Sec.~\ref{oopseSec:mechanics}.  The {\tt minimizer} keyword selects
362 > which minimization method to use, and more details on the choices of
363 > minimizer parameters can be found in
364 > Sec.~\ref{oopseSec:minimizer}. The {\tt forceField} statement is
365 > important for the selection of which forces will be used in the course
366 > of the simulation. {\sc oopse} supports several force fields, as
367 > outlined in Sec.~\ref{oopseSec:empiricalEnergy}. The force fields are
368 > interchangeable between simulations, with the only requirement being
369 > that all atoms needed by the simulation are defined within the
370 > selected force field.
371  
372 + For molecular dynamics simulations, the time step between force
373 + evaluations is set with the {\tt dt} parameter, and {\tt runTime} will
374 + set the time length of the simulation. Note, that {\tt runTime} is an
375 + absolute time, meaning if the simulation is started at t = 10.0~ns
376 + with a {\tt runTime} of 25.0~ns, the simulation will only run for an
377 + additional 15.0~ns.  
378 +
379 + For energy minimizations, it is not necessary to specify {\tt dt} or
380 + {\tt runTime}.
381 +
382 + The final required parameter is the {\tt initialConfig}
383 + statement. This will set the initial coordinates for the system, as
384 + well as the initial time if the {\tt useInitalTime} flag is set to
385 + {\tt true}. The format of the file specified in {\tt initialConfig},
386 + is given in Sec.~\ref{oopseSec:coordFiles}. Additional parameters are
387 + summarized in Table~\ref{table:genParams}.
388 +
389 + It is important to note the fundamental units in all files which are
390 + read and written by {\sc oopse}.  Energies are in $\mbox{kcal
391 + mol}^{-1}$, distances are in $\mbox{\AA}$, times are in $\mbox{fs}$,
392 + translational velocities are in $\mbox{\AA~fs}^{-1}$, and masses are
393 + in $\mbox{amu}$.  Orientational degrees of freedom are described using
394 + quaternions (unitless, but $q_w^2 + q_x^2 + q_y^2 + q_z^2 = 1$),
395 + body-fixed angular momenta ($\mbox{amu \AA}^{2} \mbox{radians
396 + fs}^{-1}$), and body-fixed moments of inertia ($\mbox{amu \AA}^{2}$).
397 +
398   \begin{table}
399 < \caption{The Global Keywords: Required Parameters}
399 > \caption{Meta-data Keywords: Required Parameters}
400   \label{table:reqParams}
401   \begin{center}
402   % Note when adding or removing columns, the \hsize numbers must add up to the total number
# Line 300 | Line 409 | The remainder of the script then sets the various simu
409  
410   {\bf keyword} & {\bf units} & {\bf use} & {\bf remarks} \\ \hline
411  
412 < {\tt forceField} & string & Sets the force field. & Possible force fields are "DUFF", "LJ", and "EAM". \\
304 < {\tt ensemble} & string & Sets the ensemble. & Possible ensembles are "NVE", "NVT", "NPTi", "NPTf", and "NPTxyz".\\
305 < {\tt dt} & fs & Sets the time step. & Selection of {\tt dt} should be small enough to sample the fastest motion of the simulation. \\
412 > {\tt forceField} & string & Sets the force field. & Possible force fields are DUFF, LJ, and EAM. \\
413   {\tt nComponents} & integer & Sets the number of components. & Needs to appear before the first {\tt Component} block. \\
414   {\tt initialConfig} & string & Sets the file containing the initial configuration. & Can point to any file containing the configuration in the correct order. \\
415 < {\tt runTime} & fs & Sets the time at which the simulation should end. & This is an absolute time, and will end the simulation when the current time meets or exceeds the {\tt runTime}. \\
415 > {\tt minimizer}& string & Chooses a minimizer & Possible minimizers
416 > are SD and CG. Either {\tt ensemble} or {\tt minimizer} must be specified. \\
417 > {\tt ensemble} & string & Sets the ensemble. & Possible ensembles are
418 > NVE, NVT, NPTi, NPTf, and NPTxyz.  Either {\tt ensemble}
419 > or {\tt minimizer} must be specified. \\
420 > {\tt dt} & fs & Sets the time step. & Selection of {\tt dt} should be
421 > small enough to sample the fastest motion of the simulation. ({\tt
422 > dt} is required for molecular dynamics simulations)\\
423 > {\tt runTime} & fs & Sets the time at which the simulation should
424 > end. & This is an absolute time, and will end the simulation when the
425 > current time meets or exceeds the {\tt runTime}. ({\tt runTime} is
426 > required for molecular dynamics simulations)\\
427  
310
428   \end{tabularx}
429   \end{center}
430   \end{table}
431  
432   \begin{table}
433 < \caption{The Global Keywords: General Parameters}
433 > \caption{Meta-data Keywords: General Parameters}
434   \label{table:genParams}
435   \begin{center}
436   % Note when adding or removing columns, the \hsize numbers must add up to the total number
# Line 326 | Line 443 | The remainder of the script then sets the various simu
443  
444   {\bf keyword} & {\bf units} & {\bf use} & {\bf remarks} \\ \hline
445  
446 < {\tt finalConfig} & string & Option to set the name of the final output file. & Useful when stringing simulations together. Defaults to the {\tt .bass} file with an {\tt .eor} extension. \\
447 < {\tt useInitialTime} & logical & Sets whether the initial time is taken from the {\tt .init} file. & Useful when recovering a simulation from a crashed processor. Default is false. \\
448 < {\tt sampleTime} & fs & Sets the frequency at which the {\tt .dump} file is written. & Default sets the frequency to the {\tt runTime}. \\
449 < {\tt statusTime} & fs & Sets the frequency at which the {\tt .stat} file is written. & Defaults sets the frequency to the {\tt sampleTime}. \\
450 < {\tt LJrcut} & $\mbox{\AA}$ & Manually sets the Lennard-Jones cutoff. & Defaults to $2.5\sigma_L$, where $\sigma_L$ is the largest LJ $\sigma$ in the simulation. \\
451 < {\tt electrostaticCutoffRadius}& & & \\
452 <      & $\mbox{\AA}$ & Manually sets the cutoff used by the electrostatic potentials. & Defaults to $15\mbox{\AA}$ \\
453 < {\tt electrostaticSkinThickness} & & & \\
454 <     & $\mbox{\AA}$  & Manually sets the skin thickness for the electrostatic switching function. & Defaults to 5~\% of the {\tt electrostaticSkinThickness}. \\
455 < {\tt useReactionField} & logical & Turns the reaction field correction on/off. & Default is "false". \\
446 > {\tt finalConfig} & string & Sets the name of the final
447 > output file. & Useful when stringing simulations together. Defaults to
448 > the root name of the initial meta-data file but with an {\tt .eor}
449 > extension. \\
450 > {\tt useInitialTime} & logical & Sets whether the initial time is taken from the {\tt .in} file. & Useful when recovering a simulation from a crashed processor. Default is false. \\
451 > {\tt sampleTime} & fs & Sets the frequency at which the {\tt .dump}
452 > file is written. & The default is equal to the {\tt runTime}. \\
453 > {\tt statusTime} & fs & Sets the frequency at which the {\tt .stat}
454 > file is written. & The default is equal to the {\tt sampleTime}. \\
455 > {\tt cutoffRadius} & $\mbox{\AA}$ & Manually sets the cutoff radius &
456 > Defaults to $15\mbox{\AA}$ for systems containing charges or dipoles or to $2.5
457 > \sigma_{L}$, where $\sigma_{L}$ is the largest LJ $\sigma$ in the
458 > simulation. \\
459 > {\tt switchingRadius} & $\mbox{\AA}$  & Manually sets the inner radius for the switching function. & Defaults to 95~\% of the {\tt cutoffRadius}. \\
460 > {\tt useReactionField} & logical & Turns the reaction field correction on/off. & Default is false. \\
461   {\tt dielectric} & unitless & Sets the dielectric constant for reaction field. & If {\tt useReactionField} is true, then {\tt dielectric} must be set. \\
462   {\tt usePeriodicBoundaryConditions} & & & \\
463 <        & logical & Turns periodic boundary conditions on/off. & Default is "true". \\
464 < {\tt seed } & integer & Sets the seed value for the random number generator. & The seed needs to be at least 9 digits long. The default is to take the seed from the CPU clock.
463 >        & logical & Turns periodic boundary conditions on/off. & Default is true. \\
464 > {\tt seed } & integer & Sets the seed value for the random number
465 > generator. & The seed needs to be at least 9 digits long. The default
466 > is to take the seed from the CPU clock. \\
467 > {\tt forceFieldVariant} & string & Sets the name of the variant of the
468 > force field.  & {\sc eam} has three variants: {\tt u3}, {\tt u6}, and
469 > {\tt VC}.
470  
471   \end{tabularx}
472   \end{center}
473   \end{table}
474  
475  
349
476   \subsection{\label{oopseSec:coordFiles}Coordinate Files}
477  
478   The standard format for storage of a systems coordinates is a modified
479   xyz-file syntax, the exact details of which can be seen in
480   Scheme~\ref{sch:dumpFormat}. As all bonding and molecular information
481 < is stored in the \texttt{.bass} and \texttt{.mdl} files, the
482 < coordinate files are simply the complete set of coordinates for each
483 < atom at a given simulation time. One important note, although the
484 < simulation propagates the complete rotation matrix, directional
485 < entities are written out using quanternions, to save space in the
486 < output files.
481 > is stored in the meta-data files, the coordinate files contain only
482 > the coordinates of the objects which move independently during the
483 > simulation.  It is important to note that {\it not all atoms} are
484 > capable of independent motion.  Atoms which are part of rigid bodies
485 > are not ``integrable objects'' in the equations of motion; the rigid
486 > bodies themselves are the integrable objects.  Therefore, the
487 > coordinate file contains coordinates of all the {\tt
488 > integrableObjects} in the system.  For systems without rigid bodies,
489 > this is simply the coordinates of all the atoms.
490  
491 < \begin{lstlisting}[float,caption={[The format of the coordinate files]Shows the format of the coordinate files. The fist line is the number of atoms. The second line begins with the time stamp followed by the three $\mathsf{H}$ column vectors. It is important to note, that for extended system ensembles, additional information pertinent to the integrators may be stored on this line as well. The next lines are the atomic coordinates for all atoms in the system. First is the name followed by position, velocity, quanternions, and lastly, body fixed angular momentum.},label=sch:dumpFormat]
491 > It is important to note that although the simulation propagates the
492 > complete rotation matrix, directional entities are written out using
493 > quaternions to save space in the output files.  All objects (atoms,
494 > orientational atoms, and rigid bodies) are given quaternions and
495 > angular momenta in coordinate files which are output by OOPSE, but it
496 > is not necessary for the user to specify the quaternions or angular
497 > momenta for atoms without orientational degrees of freedom.
498  
499 < nAtoms
499 > \begin{lstlisting}[float,caption={[The format of the coordinate
500 > files] An example of the format of the coordinate files. The fist line
501 > is the number of {\tt integrableObjects} (freely-moving atoms and
502 > rigid bodies). The second line begins with the time stamp followed by
503 > the three $\mathsf{H}$ column vectors. It is important to note that
504 > for extended system ensembles, additional information pertinent to the
505 > integrators may be stored on this line as well. The next lines are the
506 > coordinates for all integrable objects in the system.  The name of the
507 > integrable object is followed by position, velocity, quaternions, and
508 > lastly, body fixed angular momentum.},label=sch:dumpFormat]
509 >
510 > nIntegrable
511   time; Hxx Hyx Hzx; Hxy Hyy Hzy; Hxz Hyz Hzz;
512 < Name1 x y z vx vy vz q0 q1 q2 q3 jx jy jz
513 < Name2 x y z vx vy vz q0 q1 q2 q3 jx jy jz
512 > Name1 x y z vx vy vz qw qx qy qz jx jy jz
513 > Name2 x y z vx vy vz qw qx qy qz jx jy jz
514   etc...
515  
516   \end{lstlisting}
517  
518 + The {\tt name} field for atoms is simply the atom type as specified in
519 + the meta-data file.  The {\tt name} field for a rigid body is
520 + specified as {\tt MOLTYPE\_RB\_N}, to specify that this is {\tt
521 + rigidBody} N in a molecule of type MOLTYPE.  In simulations with rigid
522 + body models of water, a sample coordinate line might be:
523  
524 < There are three major files used by {\sc oopse} written in the
525 < coordinate format, they are as follows: the initialization file
526 < (\texttt{.init}), the simulation trajectory file (\texttt{.dump}), and
376 < the final coordinates of the simulation (\texttt{.eor}). The initialization file is
377 < necessary for {\sc oopse} to start the simulation with the proper
378 < coordinates, and is generated before the simulation run. The
379 < trajectory file is created at the beginning of the simulation, and is
380 < used to store snapshots of the simulation at regular intervals. The
381 < first frame is a duplication of the
382 < \texttt{.init} file, and each subsequent frame is appended to the file
383 < at an interval specified in the \texttt{.bass} file with the
384 < \texttt{sampleTime} flag. The final coordinate file is the end of run file. The
385 < \texttt{.eor} file stores the final configuration of the system for a
386 < given simulation. The file is updated at the same time as the
387 < \texttt{.dump} file, however, it only contains the most recent
388 < frame. In this way, an \texttt{.eor} file may be used as the
389 < initialization file to a second simulation in order to continue a
390 < simulation or recover one from a processor that has crashed during the
391 < course of the run.
524 > \begin{tt}
525 > TIP3P\_RB\_0  x y z vx vy vz qw qx qy qz jx jy jz
526 > \end{tt}
527  
528 < \subsection{\label{oopseSec:initCoords}Generation of Initial Coordinates}
528 > which tells the program that the rigid body representing a TIP3P
529 > molecule (rigid body \# 0) is listed on that line.
530  
531 < As was stated in Sec.~\ref{oopseSec:coordFiles}, an initialization
532 < file is needed to provide the starting coordinates for a
533 < simulation.  Several helper programs are provided with {\sc oopse} to illustrate possible build routes. However, as each simulation is different, system creation is left to the end user. The {\tt .init} file must list the atoms in the correct order or {\sc oopse} will give an atom mismatch error.
531 > There are three files used by {\sc oopse} which are written in the
532 > coordinate format.  They are: the initial coordinate file
533 > (\texttt{.in}), the simulation trajectory file (\texttt{.dump}), and
534 > the final coordinates or ``end-of-run'' for the simulation
535 > (\texttt{.eor}). The initial coordinate file is necessary for {\sc
536 > oopse} to start the simulation with the proper coordinates, and this
537 > file must be generated by the user before the simulation run. The
538 > trajectory (or ``dump'') file is updated during simulation and is used
539 > to store snapshots of the coordinates at regular intervals. The first
540 > frame is a duplication of the
541 > \texttt{.in} file, and each subsequent frame is appended to the file
542 > at an interval specified in the meta-data file with the
543 > \texttt{sampleTime} flag. The final coordinate file is the
544 > ``end-of-run'' file.  The \texttt{.eor} file stores the final
545 > configuration of the system for a given simulation. The file is
546 > updated at the same time as the \texttt{.dump} file, but it only
547 > contains the most recent frame. In this way, an \texttt{.eor} file may
548 > be used to initialize a second simulation should it be necessary to
549 > recover from a crash or power outage.
550  
551 < The correct ordering of the atoms relies on the ordering of atoms and molecules within the model and {\sc bass} scripts. {\sc oopse} expects the order to comply with the following guidelines:
551 > \subsection{\label{oopseSec:initCoords}Generation of Initial Coordinates}
552 >
553 > As was stated in Sec.~\ref{oopseSec:coordFiles}, an initial coordinate
554 > file is needed to provide the starting coordinates for a simulation.
555 > Since each simulation is different, system creation is left to the end
556 > user; however, we have included a few sample programs which make some
557 > specialized structures.  The {\tt .in} file must list the integrable
558 > objects in the correct order.  The ordering of the integrable objects
559 > relies on the ordering of molecules within the meta-data file. {\sc
560 > oopse} expects the order to comply with the following guidelines:
561   \begin{enumerate}
562 < \item All of the molecules of the first declared component are given before proceeding to the molecules of the second component, and so on for all declared components.
563 < \item The ordering of the atoms for each molecule follows the order declared in the molecule's declaration within the model file.
562 > \item All of the molecules of the first declared component are given
563 > before proceeding to the molecules of the second component, and so on
564 > for all subsequently declared components.
565 > \item The ordering of the atoms for each molecule follows the order
566 > declared in the molecule's declaration within the model file.
567 > \item Only atoms which are not members of a {\tt rigidBody} are
568 > included.
569 > \item Rigid Body coordinates for a molecule are listed immediately
570 > after the the other atoms in a molecule.  Some molecules may be
571 > entirely rigid, in which case, only the rigid body coordinates are
572 > given.
573   \end{enumerate}
574 < An example is given in Scheme~\ref{sch:initEx1} resulting in the {\tt .init} file shown in Scheme~\ref{sch:initEx2}.
574 > An example is given in the meta-data file in Scheme~\ref{sch:initEx1}
575 > which results in the {\tt .in} file shown in Scheme~\ref{sch:initEx2}.
576  
577 < \begin{lstlisting}[float,caption={This scheme illustrates the declaration of the $\text{I}_2$ molecule and the HCl molecule. The two molecules are then included into a simulation.}, label=sch:initEx1]
577 > \begin{lstlisting}[float,caption={Example declaration of the
578 > $\text{I}_2$ molecule and the HCl molecule. The two molecules are then
579 > included into a simulation.}, label=sch:initEx1]
580  
581   molecule{
582    name = "I2";
# Line 445 | Line 618 | initialConfig = "mixture.init";
618    nMol = 1;
619   }
620  
621 < initialConfig = "mixture.init";
621 > initialConfig = "mixture.in";
622  
623   \end{lstlisting}
624  
625 < \begin{lstlisting}[float,caption={This is the contents of the {\tt mixture.init} file matching the declarations in Scheme~\ref{sch:initEx1}. Note that even though $\text{I}_2$ is declared before HCl, the {\tt .init} file follows the order in which the components were included.},label=sch:initEx2]
625 > \begin{lstlisting}[float,caption={The contents of the {\tt
626 > mixture.in} file matching the declarations in
627 > Scheme~\ref{sch:initEx1}. Note that even though $\text{I}_2$ is
628 > declared before HCl, the {\tt .in} file follows the order {\it in
629 > which the components were included}.},label=sch:initEx2]
630  
631   10
632   0.0;  10.0  0.0  0.0;  0.0  10.0  0.0;  0.0  0.0  10.0;
# Line 471 | Line 648 | instantaneous temperature, volume, pressure, etc. It i
648  
649   The last output file generated by {\sc oopse} is the statistics
650   file. This file records such statistical quantities as the
651 < instantaneous temperature, volume, pressure, etc. It is written out
652 < with the frequency specified in the \texttt{.bass} file with the
651 > instantaneous temperature (in $K$), volume (in $\mbox{\AA}^{3}$),
652 > pressure (in $\mbox{atm}$), etc. It is written out with the frequency
653 > specified in the meta-data file with the
654   \texttt{statusTime} keyword. The file allows the user to observe the
655   system variables as a function of simulation time while the simulation
656   is in progress. One useful function the statistics file serves is to
657 < monitor the conserved quantity of a given simulation ensemble, this
658 < allows the user to observe the stability of the integrator. The
657 > monitor the conserved quantity of a given simulation ensemble,
658 > allowing the user to gauge the stability of the integrator. The
659   statistics file is denoted with the \texttt{.stat} file extension.
660  
661 + \section{\label{oopseSec:empiricalEnergy}The Empirical Energy
662 + Functions}
663  
664 < \section{\label{oopseSec:empiricalEnergy}The Empirical Energy Functions}
664 > Like many simulation packages, {\sc oopse} splits the potential energy
665 > into the short-ranged (bonded) portion and a long-range (non-bonded)
666 > potential,
667 > \begin{equation}
668 > V = V_{\mathrm{short-range}} + V_{\mathrm{long-range}}.
669 > \end{equation}
670 > The short-ranged portion includes the explicit bonds, bends, and
671 > torsions which have been defined in the meta-data file for the
672 > molecules which are present in the simulation.  The functional forms and
673 > parameters for these interactions are defined by the force field which
674 > is chosen.
675  
676 < \
676 > Calculating the long-range (non-bonded) potential involves a sum over
677 > all pairs of atoms (except for those atoms which are involved in a
678 > bond, bend, or torsion with each other).  If done poorly, calculating
679 > the the long-range interactions for $N$ atoms would involve $N(N-1)/2$
680 > evaluations of atomic distances.  To reduce the number of distance
681 > evaluations between pairs of atoms, {\sc oopse} uses a switched cutoff
682 > with Verlet neighbor lists.\cite{allen87:csl} It is well known that
683 > neutral groups which contain charges will exhibit pathological forces
684 > unless the cutoff is applied to the neutral groups evenly instead of
685 > to the individual atoms.\cite{leach01:mm} {\sc oopse} allows users to
686 > specify cutoff groups which may contain an arbitrary number of atoms
687 > in the molecule.  Atoms in a cutoff group are treated as a single unit
688 > for the evaluation of the switching function:
689 > \begin{equation}
690 > V_{\mathrm{long-range}} = \sum_{a} \sum_{b>a} s(r_{ab}) \sum_{i \in a} \sum_{j \in b} V_{ij}(r_{ij}),
691 > \end{equation}
692 > where $r_{ab}$ is the distance between the centers of mass of the two
693 > cutoff groups ($a$ and $b$).
694 >
695 > The sums over $a$ and $b$ are over the cutoff groups that are present
696 > in the simulation.  Atoms which are not explicitly defined as members
697 > of a {\tt cutoffGroup} are treated as a group consisting of only one
698 > atom.  The switching function, $s(r)$ is the standard cubic switching
699 > function,
700 > \begin{equation}
701 > S(r) =
702 >        \begin{cases}
703 >        1 & \text{if $r \le r_{\text{sw}}$},\\
704 >        \frac{(r_{\text{cut}} + 2r - 3r_{\text{sw}})(r_{\text{cut}} - r)^2}
705 >        {(r_{\text{cut}} - r_{\text{sw}})^2}
706 >        & \text{if $r_{\text{sw}} < r \le r_{\text{cut}}$}, \\
707 >        0 & \text{if $r > r_{\text{cut}}$.}
708 >        \end{cases}
709 > \label{eq:dipoleSwitching}
710 > \end{equation}
711 > Here, $r_{\text{sw}}$ is the {\tt switchingRadius}, or the distance
712 > beyond which interactions are reduced, and $r_{\text{cut}}$ is the
713 > {\tt cutoffRadius}, or the distance at which interactions are
714 > truncated.
715 >
716 > Users of {\sc oopse} do not need to specify the {\tt cutoffRadius} or
717 > {\tt switchingRadius}.  In simulations containing only Lennard-Jones
718 > atoms, the cutoff radius has a default value of $2.5\sigma_{ii}$,
719 > where $\sigma_{ii}$ is the largest Lennard-Jones length parameter
720 > present in the simulation.  In simulations containing charged or
721 > dipolar atoms, the default cutoff radius is $15 \mbox{\AA}$.  
722 >
723 > The {\tt switchingRadius} is set to a default value of 95\% of the
724 > {\tt cutoffRadius}.  In the special case of a simulation containing
725 > {\it only} Lennard-Jones atoms, the default switching radius takes the
726 > same value as the cutoff radius, and {\sc oopse} will use a shifted
727 > potential to remove discontinuities in the potential at the cutoff.
728 > Both radii may be specified in the meta-data file.
729 >
730 > Force fields can be added to {\sc oopse}, although it comes with a few
731 > simple examples (Lennard-Jones, {\sc duff}, {\sc water}, and {\sc
732 > eam}) which are explained in the following sections.
733 >
734   \subsection{\label{sec:LJPot}The Lennard Jones Force Field}
735  
736   The most basic force field implemented in {\sc oopse} is the
737 < Lennard-Jones force field, which mimics the van der Waals interaction at
738 < long distances, and uses an empirical repulsion at short
737 > Lennard-Jones force field, which mimics the van der Waals interaction
738 > at long distances and uses an empirical repulsion at short
739   distances. The Lennard-Jones potential is given by:
740   \begin{equation}
741   V_{\text{LJ}}(r_{ij}) =
# Line 501 | Line 748 | $\epsilon_{ij}$ scales the well depth of the potential
748   where $r_{ij}$ is the distance between particles $i$ and $j$,
749   $\sigma_{ij}$ scales the length of the interaction, and
750   $\epsilon_{ij}$ scales the well depth of the potential. Scheme
751 < \ref{sch:LJFF} gives an example \texttt{.bass} file that
751 > \ref{sch:LJFF} gives an example meta-data file that
752   sets up a system of 108 Ar particles to be simulated using the
753   Lennard-Jones force field.
754  
755 < \begin{lstlisting}[float,caption={[Invocation of the Lennard-Jones force field] A sample system using the Lennard-Jones force field.},label={sch:LJFF}]
755 > \begin{lstlisting}[float,caption={[Invocation of the Lennard-Jones
756 > force field] A sample meta-data file for a small Lennard-Jones
757 > simulation.},label={sch:LJFF}]
758  
759 < #include "argon.mdl"
759 > #include "argon.md"
760  
761   nComponents = 1;
762   component{
# Line 515 | Line 764 | initialConfig = "./argon.init";
764    nMol = 108;
765   }
766  
767 < initialConfig = "./argon.init";
767 > initialConfig = "./argon.in";
768  
769   forceField = "LJ";
770   \end{lstlisting}
771  
523 Because this potential is calculated between all pairs, the force
524 evaluation can become computationally expensive for large systems. To
525 keep the pair evaluations to a manageable number, {\sc oopse} employs
526 a cut-off radius.\cite{allen87:csl} The cutoff radius can either be
527 specified in the \texttt{.bass} file, or left as its default value of
528 $2.5\sigma_{ii}$, where $\sigma_{ii}$ is the largest Lennard-Jones
529 length parameter present in the simulation. Truncating the calculation
530 at $r_{\text{cut}}$ introduces a discontinuity into the potential
531 energy and the force. To offset this discontinuity in the potential,
532 the energy value at $r_{\text{cut}}$ is subtracted from the
533 potential. This causes the potential to go to zero smoothly at the
534 cut-off radius, and preserves conservation of energy in integrating
535 the equations of motion. There still remains a discontinuity in the derivative (the forces), however, this does not significantly affect the dynamics.
536
772   Interactions between dissimilar particles requires the generation of
773 < cross term parameters for $\sigma$ and $\epsilon$. These are
774 < calculated through the Lorentz-Berthelot mixing
773 > cross term parameters for $\sigma$ and $\epsilon$. These parameters
774 > are determined using the Lorentz-Berthelot mixing
775   rules:\cite{allen87:csl}
776   \begin{equation}
777   \sigma_{ij} = \frac{1}{2}[\sigma_{ii} + \sigma_{jj}],
# Line 551 | Line 786 | simulate lipid bilayers. The simulations require a mod
786   \subsection{\label{oopseSec:DUFF}Dipolar Unified-Atom Force Field}
787  
788   The dipolar unified-atom force field ({\sc duff}) was developed to
789 < simulate lipid bilayers. The simulations require a model capable of
790 < forming bilayers, while still being sufficiently computationally
791 < efficient to allow large systems ($\sim$100's of phospholipids,
792 < $\sim$1000's of waters) to be simulated for long times
793 < ($\sim$10's of nanoseconds).
794 <
795 < With this goal in mind, {\sc duff} has no point
796 < charges. Charge-neutral distributions were replaced with dipoles,
797 < while most atoms and groups of atoms were reduced to Lennard-Jones
798 < interaction sites. This simplification cuts the length scale of long
799 < range interactions from $\frac{1}{r}$ to $\frac{1}{r^3}$, removing the need for the computationally expensive Ewald sum. Instead, we Verlet neighbor-lists and cutoff radii are used for the dipolar interactions, or a reaction field is added to mimic longer range interactions.
789 > simulate lipid bilayers. These types of simulations require a model
790 > capable of forming bilayers, while still being sufficiently
791 > computationally efficient to allow large systems ($\sim$100's of
792 > phospholipids, $\sim$1000's of waters) to be simulated for long times
793 > ($\sim$10's of nanoseconds). With this goal in mind, {\sc duff} has no
794 > point charges. Charge-neutral distributions are replaced with dipoles,
795 > while most atoms and groups of atoms are reduced to Lennard-Jones
796 > interaction sites. This simplification reduces the length scale of
797 > long range interactions from $\frac{1}{r}$ to $\frac{1}{r^3}$,
798 > removing the need for the computationally expensive Ewald
799 > sum. Instead, Verlet neighbor-lists and cutoff radii are used for the
800 > dipolar interactions, and, if desired, a reaction field may be added
801 > to mimic longer range interactions.
802  
803   As an example, lipid head-groups in {\sc duff} are represented as
804 < point dipole interaction sites. By placing a dipole at the head
805 < group's center of mass, our model mimics the charge separation found
806 < in common phospholipid head groups such as
807 < phosphatidylcholine.\cite{Cevc87} Additionally, a large Lennard-Jones
808 < site is located at the pseudoatom's center of mass. The model is
809 < illustrated by the red atom in Fig.~\ref{oopseFig:lipidModel}. The
810 < water model we use to complement the dipoles of the lipids is our
811 < reparameterization\cite{fennell04} of the soft sticky dipole (SSD) model of Ichiye
804 > point dipole interaction sites.  Placing a dipole at the head group's
805 > center of mass mimics the charge separation found in common
806 > phospholipid head groups such as phosphatidylcholine.\cite{Cevc87}
807 > Additionally, a large Lennard-Jones site is located at the
808 > pseudoatom's center of mass. The model is illustrated by the red atom
809 > in Fig.~\ref{oopseFig:lipidModel}. The water model we use to
810 > complement the dipoles of the lipids is a
811 > reparameterization\cite{fennell04} of the soft sticky dipole (SSD)
812 > model of Ichiye
813   \emph{et al.}\cite{liu96:new_model}
814  
815   \begin{figure}
816   \centering
817 < \includegraphics[width=\linewidth]{twoChainFig.pdf}
818 < \caption[A representation of a lipid model in {\sc duff}]{A representation of the lipid model. $\phi$ is the torsion angle, $\theta$ %
819 < is the bend angle, and $\mu$ is the dipole moment of the head group.}
817 > \includegraphics[width=\linewidth]{lipidModel.eps}
818 > \caption[A representation of a lipid model in {\sc duff}]{A
819 > representation of the lipid model. $\phi$ is the torsion angle,
820 > $\theta$ is the bend angle, and $\mu$ is the dipole moment of the head
821 > group.}
822   \label{oopseFig:lipidModel}
823   \end{figure}
824  
825 < We have used a set of scalable parameters to model the alkyl groups
826 < with Lennard-Jones sites. For this, we have borrowed parameters from
827 < the TraPPE force field of Siepmann
828 < \emph{et al}.\cite{Siepmann1998} TraPPE is a unified-atom
829 < representation of n-alkanes, which is parametrized against phase
830 < equilibria using Gibbs ensemble Monte Carlo simulation
831 < techniques.\cite{Siepmann1998} One of the advantages of TraPPE is that
832 < it generalizes the types of atoms in an alkyl chain to keep the number
833 < of pseudoatoms to a minimum; the parameters for a unified atom such as
834 < $\text{CH}_2$ do not change depending on what species are bonded to
595 < it.
825 > A set of scalable parameters has been used to model the alkyl groups
826 > with Lennard-Jones sites. For this, parameters from the TraPPE force
827 > field of Siepmann \emph{et al.}\cite{Siepmann1998} have been
828 > utilized. TraPPE is a unified-atom representation of n-alkanes which
829 > is parametrized against phase equilibria using Gibbs ensemble Monte
830 > Carlo simulation techniques.\cite{Siepmann1998} One of the advantages
831 > of TraPPE is that it generalizes the types of atoms in an alkyl chain
832 > to keep the number of pseudoatoms to a minimum; thus, the parameters
833 > for a unified atom such as $\text{CH}_2$ do not change depending on
834 > what species are bonded to it.
835  
836 < TraPPE and {\sc duff} also constrain all bonds to be of fixed length. Typically,
837 < bond vibrations are the fastest motions in a molecular dynamic
838 < simulation. Small time steps between force evaluations must be used to
839 < ensure adequate energy conservation in the bond degrees of freedom. By
840 < constraining the bond lengths, larger time steps may be used when
841 < integrating the equations of motion. A simulation using {\sc duff} is
842 < illustrated in Scheme \ref{sch:DUFF}.
836 > As is required by TraPPE, {\sc duff} also constrains all bonds to be
837 > of fixed length. Typically, bond vibrations are the fastest motions in
838 > a molecular dynamic simulation.  With these vibrations present, small
839 > time steps between force evaluations must be used to ensure adequate
840 > energy conservation in the bond degrees of freedom. By constraining
841 > the bond lengths, larger time steps may be used when integrating the
842 > equations of motion. A simulation using {\sc duff} is illustrated in
843 > Scheme \ref{sch:DUFF}.
844  
845 < \begin{lstlisting}[float,caption={[Invocation of {\sc duff}]A portion of a \texttt{.bass} file showing a simulation utilizing {\sc duff}},label={sch:DUFF}]
845 > \begin{lstlisting}[float,caption={[Invocation of {\sc duff}]A portion
846 > of a meta-data file showing a simulation utilizing {\sc
847 > duff}},label={sch:DUFF}]  
848  
849 < #include "water.mdl"
850 < #include "lipid.mdl"
849 > #include "water.md"
850 > #include "lipid.md"
851  
852   nComponents = 2;
853   component{
# Line 618 | Line 860 | initialConfig = "bilayer.init";
860    nMol = 1936;
861   }
862  
863 < initialConfig = "bilayer.init";
863 > initialConfig = "bilayer.in";
864  
865   forceField = "DUFF";
866  
# Line 644 | Line 886 | within the molecule $I$, and $V_{\text{torsion}}$ is t
886   \label{eq:internalPotential}
887   \end{equation}
888   Here $V_{\text{bend}}$ is the bend potential for all 1, 3 bonded pairs
889 < within the molecule $I$, and $V_{\text{torsion}}$ is the torsion potential
890 < for all 1, 4 bonded pairs. The pairwise portions of the internal
891 < potential are excluded for atom pairs that are involved in the same bond, bend, or torsion. All other atom pairs within the molecule are subject to the LJ pair potential.
889 > within the molecule $I$, and $V_{\text{torsion}}$ is the torsion
890 > potential for all 1, 4 bonded pairs.  The pairwise portions of the
891 > non-bonded interactions are excluded for atom pairs that are involved
892 > in the smae bond, bend, or torsion. All other atom pairs within a
893 > molecule are subject to the LJ pair potential.
894  
651
895   The bend potential of a molecule is represented by the following function:
896   \begin{equation}
897 < V_{\text{bend}}(\theta_{ijk}) = k_{\theta}( \theta_{ijk} - \theta_0 )^2, \label{eq:bendPot}
897 > V_{\text{bend}}(\theta_{ijk}) = k_{\theta}( \theta_{ijk} - \theta_0
898 > )^2, \label{eq:bendPot}
899   \end{equation}
900   where $\theta_{ijk}$ is the angle defined by atoms $i$, $j$, and $k$
901   (see Fig.~\ref{oopseFig:lipidModel}), $\theta_0$ is the equilibrium
# Line 691 | Line 935 | evaluations are avoided during the calculation of the
935   k_3 &= 4c_3.
936   \end{align*}
937   By recasting the potential as a power series, repeated trigonometric
938 < evaluations are avoided during the calculation of the potential energy.
938 > evaluations are avoided during the calculation of the potential
939 > energy.
940  
941  
942 < The cross potential between molecules $I$ and $J$, $V^{IJ}_{\text{Cross}}$, is
943 < as follows:
942 > The cross potential between molecules $I$ and $J$,
943 > $V^{IJ}_{\text{Cross}}$, is as follows:
944   \begin{equation}
945   V^{IJ}_{\text{Cross}} =
946          \sum_{i \in I} \sum_{j \in J}
# Line 725 | Line 970 | respectively. $|\mu_i|$ is the magnitude of the dipole
970   Here $\mathbf{r}_{ij}$ is the vector starting at atom $i$ pointing
971   towards $j$, and $\boldsymbol{\Omega}_i$ and $\boldsymbol{\Omega}_j$
972   are the orientational degrees of freedom for atoms $i$ and $j$
973 < respectively. $|\mu_i|$ is the magnitude of the dipole moment of atom
974 < $i$, $\boldsymbol{\hat{u}}_i$ is the standard unit orientation vector
975 < of $\boldsymbol{\Omega}_i$, and $\boldsymbol{\hat{r}}_{ij}$ is the
976 < unit vector pointing along $\mathbf{r}_{ij}$
973 > respectively. The magnitude of the dipole moment of atom $i$ is
974 > $|\mu_i|$, $\boldsymbol{\hat{u}}_i$ is the standard unit orientation
975 > vector of $\boldsymbol{\Omega}_i$, and $\boldsymbol{\hat{r}}_{ij}$ is
976 > the unit vector pointing along $\mathbf{r}_{ij}$
977   ($\boldsymbol{\hat{r}}_{ij}=\mathbf{r}_{ij}/|\mathbf{r}_{ij}|$).
978  
979 < To improve computational efficiency of the dipole-dipole interactions,
980 < {\sc oopse} employs an electrostatic cutoff radius. This parameter can
736 < be set in the \texttt{.bass} file, and controls the length scale over
737 < which dipole interactions are felt. To compensate for the
738 < discontinuity in the potential and the forces at the cutoff radius, we
739 < have implemented a switching function to smoothly scale the
740 < dipole-dipole interaction at the cutoff.
741 < \begin{equation}
742 < S(r_{ij}) =
743 <        \begin{cases}
744 <        1 & \text{if $r_{ij} \le r_t$},\\
745 <        \frac{(r_{\text{cut}} + 2r_{ij} - 3r_t)(r_{\text{cut}} - r_{ij})^2}
746 <        {(r_{\text{cut}} - r_t)^2}
747 <        & \text{if $r_t < r_{ij} \le r_{\text{cut}}$}, \\
748 <        0 & \text{if $r_{ij} > r_{\text{cut}}$.}
749 <        \end{cases}
750 < \label{eq:dipoleSwitching}
751 < \end{equation}
752 < Here $S(r_{ij})$ scales the potential at a given $r_{ij}$, and $r_t$
753 < is the taper radius some given thickness less than the electrostatic
754 < cutoff. The switching thickness can be set in the \texttt{.bass} file.
979 > \subsubsection{\label{oopseSec:SSD}The {\sc duff} Water Models: SSD/E
980 > and SSD/RF}
981  
756 \subsubsection{\label{oopseSec:SSD}The {\sc duff} Water Models: SSD/E and SSD/RF}
757
982   In the interest of computational efficiency, the default solvent used
983   by {\sc oopse} is the extended Soft Sticky Dipole (SSD/E) water
984   model.\cite{fennell04} The original SSD was developed by Ichiye
# Line 813 | Line 1037 | Since SSD/E is a single-point {\it dipolar} model, the
1037   can be found in the original SSD
1038   articles.\cite{liu96:new_model,liu96:monte_carlo,chandra99:ssd_md,Ichiye03}
1039  
1040 + \begin{figure}
1041 + \centering
1042 + \includegraphics[width=\linewidth]{waterAngle.eps}
1043 + \caption[Coordinate definition for the SSD/E water model]{Coordinates
1044 + for the interaction between two SSD/E water molecules.  $\theta_{ij}$
1045 + is the angle that $r_{ij}$ makes with the $\hat{z}$ vector in the
1046 + body-fixed frame for molecule $i$.  The $\hat{z}$ vector bisects the
1047 + HOH angle in each water molecule. }
1048 + \label{oopseFig:ssd}
1049 + \end{figure}
1050 +
1051 +
1052   Since SSD/E is a single-point {\it dipolar} model, the force
1053   calculations are simplified significantly relative to the standard
1054   {\it charged} multi-point models. In the original Monte Carlo
1055   simulations using this model, Ichiye {\it et al.} reported that using
1056   SSD decreased computer time by a factor of 6-7 compared to other
1057 < models.\cite{liu96:new_model} What is most impressive is that these savings
1058 < did not come at the expense of accurate depiction of the liquid state
1059 < properties.  Indeed, SSD/E maintains reasonable agreement with the Head-Gordon
1060 < diffraction data for the structural features of liquid
1061 < water.\cite{hura00,liu96:new_model} Additionally, the dynamical properties
1062 < exhibited by SSD/E agree with experiment better than those of more
1063 < computationally expensive models (like TIP3P and
1064 < SPC/E).\cite{chandra99:ssd_md} The combination of speed and accurate depiction
1065 < of solvent properties makes SSD/E a very attractive model for the
1066 < simulation of large scale biochemical simulations.
1057 > models.\cite{liu96:new_model} What is most impressive is that these
1058 > savings did not come at the expense of accurate depiction of the
1059 > liquid state properties.  Indeed, SSD/E maintains reasonable agreement
1060 > with the Head-Gordon diffraction data for the structural features of
1061 > liquid water.\cite{hura00,liu96:new_model} Additionally, the dynamical
1062 > properties exhibited by SSD/E agree with experiment better than those
1063 > of more computationally expensive models (like TIP3P and
1064 > SPC/E).\cite{chandra99:ssd_md} The combination of speed and accurate
1065 > depiction of solvent properties makes SSD/E a very attractive model
1066 > for the simulation of large scale biochemical simulations.
1067  
1068   Recent constant pressure simulations revealed issues in the original
1069   SSD model that led to lower than expected densities at all target
# Line 836 | Line 1072 | model (an SSD variant  parameterized for reaction fiel
1072   exhibits improved liquid structure and transport behavior. If the use
1073   of a reaction field long-range interaction correction is desired, it
1074   is recommended that the parameters be modified to those of the SSD/RF
1075 < model (an SSD variant  parameterized for reaction field). Solvent parameters can be easily modified in an accompanying
1076 < \texttt{.bass} file as illustrated in the scheme below. A table of the
1077 < parameter values and the drawbacks and benefits of the different
1078 < density corrected SSD models can be found in
1079 < reference~\cite{fennell04}.
1075 > model (an SSD variant parameterized for reaction field). These solvent
1076 > parameters are listed and can be easily modified in the {\sc duff}
1077 > force field file ({\tt DUFF.frc}).  A table of the parameter values
1078 > and the drawbacks and benefits of the different density corrected SSD
1079 > models can be found in reference~\citen{fennell04}.
1080  
1081 < \begin{lstlisting}[float,caption={[A simulation of {\sc ssd} water]A portion of a \texttt{.bass} file showing a simulation including {\sc ssd} water.},label={sch:ssd}]
1081 > \subsection{\label{oopseSec:WATER}The {\sc water} Force Field}
1082  
1083 < #include "water.mdl"
1084 <
1085 < nComponents = 1;
1086 < component{
1087 <  type = "SSD_water";
1088 <  nMol = 864;
1089 < }
1090 <
1091 < initialConfig = "liquidWater.init";
1092 <
1093 < forceField = "DUFF";
1094 <
1095 < /*
1096 < * The following two flags set the cutoff
1097 < * radius for the electrostatic forces
1098 < * as well as the skin thickness of the switching
1099 < * function.
1100 < */
1101 <
1102 < electrostaticCutoffRadius  = 9.2;
867 < electrostaticSkinThickness = 1.38;
868 <
869 < \end{lstlisting}
870 <
1083 > In addition to the {\sc duff} force field's solvent description, a
1084 > separate {\sc water} force field has been included for simulating most
1085 > of the common rigid-body water models. This force field includes the
1086 > simple and point-dipolar models (SSD, SSD1, SSD/E, SSD/RF, and DPD
1087 > water), as well as the common charge-based models (SPC, SPC/E, TIP3P,
1088 > TIP4P, and
1089 > TIP5P).\cite{liu96:new_model,Ichiye03,fennell04,Marrink01,Berendsen81,Berendsen87,Jorgensen83,Mahoney00}
1090 > In order to handle these models, charge-charge interactions were
1091 > included in the force-loop:
1092 > \begin{equation}
1093 > V_{\text{charge}}(r_{ij}) = \sum_{ij}\frac{q_iq_je^2}{r_{ij}},
1094 > \end{equation}
1095 > where $q$ represents the charge on particle $i$ or $j$, and $e$ is the
1096 > charge of an electron in Coulombs.  As with the other pair
1097 > interactions, charges can be simulated with a pure cutoff or a
1098 > reaction field.  The {\sc water} force field can be easily expanded
1099 > through modification of the {\sc water} force field file ({\tt
1100 > WATER.frc}). By adding atom types and inserting the appropriate
1101 > parameters, it is possible to extend the force field to handle rigid
1102 > molecules other than water.
1103  
1104   \subsection{\label{oopseSec:eam}Embedded Atom Method}
1105  
1106 < {\sc oopse} implements a potential that
1107 < describes bonding transition metal
1108 < systems\cite{Finnis84,Ercolessi88,Chen90,Qi99,Ercolessi02} and has attractive interaction which models  ``Embedding''
1109 < a positively charged metal ion in the electron density due to the
1106 > {\sc oopse} implements a potential that describes bonding in
1107 > transition metal
1108 > systems.~\cite{Finnis84,Ercolessi88,Chen90,Qi99,Ercolessi02} This
1109 > potential has an attractive interaction which models ``Embedding'' a
1110 > positively charged pseudo-atom core in the electron density due to the
1111   free valance ``sea'' of electrons created by the surrounding atoms in
1112 < the system. A mostly-repulsive pairwise part of the potential
1113 < describes the interaction of the positively charged metal core ions
1114 < with one another. A particular potential description called the
1115 < Embedded Atom Method\cite{Daw84,FBD86,johnson89,Lu97}({\sc eam}) that has
1116 < particularly wide adoption has been selected for inclusion in {\sc oopse}. A
1117 < good review of {\sc eam} and other metallic potential formulations was written
1118 < by Voter.\cite{voter}
1112 > the system.  A pairwise part of the potential (which is primarily
1113 > repulsive) describes the interaction of the positively charged metal
1114 > core ions with one another.  The Embedded Atom Method ({\sc
1115 > eam})~\cite{Daw84,FBD86,johnson89,Lu97} has been widely adopted in the
1116 > materials science community and has been included in {\sc oopse}. A
1117 > good review of {\sc eam} and other formulations of metallic potentials
1118 > was given by Voter.\cite{Voter:95}
1119  
1120   The {\sc eam} potential has the form:
1121 < \begin{eqnarray}
1122 < V & = & \sum_{i} F_{i}\left[\rho_{i}\right] + \sum_{i} \sum_{j \neq i}
1123 < \phi_{ij}({\bf r}_{ij}),  \\
1124 < \rho_{i}  & = & \sum_{j \neq i} f_{j}({\bf r}_{ij}),
1125 < \end{eqnarray}
893 < where $F_{i} $ is the embedding function that equates the energy
1121 > \begin{equation}
1122 > V  =  \sum_{i} F_{i}\left[\rho_{i}\right] + \sum_{i} \sum_{j \neq i}
1123 > \phi_{ij}({\bf r}_{ij})
1124 > \end{equation}
1125 > where $F_{i} $ is an embedding functional that approximates the energy
1126   required to embed a positively-charged core ion $i$ into a linear
1127   superposition of spherically averaged atomic electron densities given
1128 < by $\rho_{i}$.  $\phi_{ij}$ is a primarily repulsive pairwise
1129 < interaction between atoms $i$ and $j$. In the original formulation of
1130 < {\sc eam}\cite{Daw84}, $\phi_{ij}$ was an entirely repulsive term,
1131 < however in later refinements to {\sc eam} have shown that non-uniqueness
1132 < between $F$ and $\phi$ allow for more general forms for
1133 < $\phi$.\cite{Daw89} There is a cutoff distance, $r_{cut}$, which
1134 < limits the summations in the {\sc eam} equation to the few dozen atoms
1128 > by $\rho_{i}$,
1129 > \begin{equation}
1130 > \rho_{i}   =  \sum_{j \neq i} f_{j}({\bf r}_{ij}),
1131 > \end{equation}
1132 > Since the density at site $i$ ($\rho_i$) must be computed before the
1133 > embedding functional can be evaluated, {\sc eam} and the related
1134 > transition metal potentials require two loops through the atom pairs
1135 > to compute the inter-atomic forces.
1136 >
1137 > The pairwise portion of the potential, $\phi_{ij}$, is a primarily
1138 > repulsive interaction between atoms $i$ and $j$. In the original
1139 > formulation of {\sc eam}\cite{Daw84}, $\phi_{ij}$ was an entirely
1140 > repulsive term; however later refinements to {\sc eam} allowed for
1141 > more general forms for $\phi$.\cite{Daw89} The effective cutoff
1142 > distance, $r_{{\text cut}}$ is the distance at which the values of
1143 > $f(r)$ and $\phi(r)$ drop to zero for all atoms present in the
1144 > simulation.  In practice, this distance is fairly small, limiting the
1145 > summations in the {\sc eam} equation to the few dozen atoms
1146   surrounding atom $i$ for both the density $\rho$ and pairwise $\phi$
1147 < interactions. Foiles \emph{et al}.~fit {\sc eam} potentials for the fcc
1148 < metals Cu, Ag, Au, Ni, Pd, Pt and alloys of these metals.\cite{FBD86}
1149 < These fits are included in {\sc oopse}.
1147 > interactions.
1148 >
1149 > In computing forces for alloys, mixing rules as outlined by
1150 > Johnson~\cite{johnson89} are used to compute the heterogenous pair
1151 > potential,
1152 > \begin{eqnarray}
1153 > \label{eq:johnson}
1154 > \phi_{ab}(r)=\frac{1}{2}\left(
1155 > \frac{f_{b}(r)}{f_{a}(r)}\phi_{aa}(r)+
1156 > \frac{f_{a}(r)}{f_{b}(r)}\phi_{bb}(r)
1157 > \right).
1158 > \end{eqnarray}
1159 > No mixing rule is needed for the densities, since the density at site
1160 > $i$ is simply the linear sum of density contributions of all the other
1161 > atoms.
1162 >
1163 > The {\sc eam} force field has a number of variants in the literature.
1164 > Foiles, Baskes and Daw fit {\sc eam} potentials for Cu, Ag, Au, Ni,
1165 > Pd, Pt and alloys of these metals.\cite{FBD86} These fits are included
1166 > in {\sc oopse} as the {\tt u3} variant of the {\sc eam} force field.
1167 > Voter and Chen reparamaterized a set of {\sc eam} functions which do a
1168 > better job of predicting melting points.\cite{Voter:87} These
1169 > functions are included in {\sc oopse} as the {\tt VC} variant of the
1170 > {\sc eam} force field.  An additional set of functions (the
1171 > ``Universal 6'' functions) are included in {\sc oopse} as the {\tt u6}
1172 > variant of {\sc eam}.  In general, to specify one variant of a force
1173 > field, the user would need a single line in the meta-data file to
1174 > select the {\tt forceFieldVariant}.  For example, to specify the
1175 > Voter-Chen variant of the {\sc eam} force field, the user would add
1176 > the {\tt forceFieldVariant = "VC";} line to the meta-data file.
1177 >
1178 > The potential files used by the {\sc eam} force field are in the
1179 > standard {\tt funcfl} format, which is the format utilized by a number
1180 > of other codes (e.g. ParaDyn~\cite{Paradyn}, {\sc dynamo 86}).  It
1181 > should be noted that the energy units in these files are in eV, not
1182 > $\mbox{kcal mol}^{-1}$ as in the rest of the {\sc oopse} force field
1183 > files.  
1184  
1185   \subsection{\label{oopseSec:pbc}Periodic Boundary Conditions}
1186  
1187   \newcommand{\roundme}{\operatorname{round}}
1188  
1189 < \textit{Periodic boundary conditions} are widely used to simulate bulk properties with a relatively small number of particles. The
1190 < simulation box is replicated throughout space to form an infinite
1189 > \textit{Periodic boundary conditions} are widely used to simulate bulk
1190 > properties with a relatively small number of particles. In this method
1191 > the simulation box is replicated throughout space to form an infinite
1192   lattice.  During the simulation, when a particle moves in the primary
1193   cell, its image in other cells move in exactly the same direction with
1194   exactly the same orientation. Thus, as a particle leaves the primary
1195   cell, one of its images will enter through the opposite face. If the
1196   simulation box is large enough to avoid ``feeling'' the symmetries of
1197   the periodic lattice, surface effects can be ignored. The available
1198 < periodic cells in OOPSE are cubic, orthorhombic and parallelepiped. We
1199 < use a $3 \times 3$ matrix, $\mathsf{H}$, to describe the shape and
1200 < size of the simulation box. $\mathsf{H}$ is defined:
1198 > periodic cells in {\sc oopse} are cubic, orthorhombic and
1199 > parallelepiped.  {\sc oopse} use a $3 \times 3$ matrix, $\mathsf{H}$,
1200 > to describe the shape and size of the simulation box. $\mathsf{H}$ is
1201 > defined:
1202   \begin{equation}
1203   \mathsf{H} = ( \mathbf{h}_x, \mathbf{h}_y, \mathbf{h}_z ),
1204   \end{equation}
# Line 936 | Line 1215 | directions. To find the minimum image of a vector $\ma
1215   \end{align}
1216   The vector $\mathbf{s}$ is now a vector expressed as the number of box
1217   lengths in the $\mathbf{h}_x$, $\mathbf{h}_y$, and $\mathbf{h}_z$
1218 < directions. To find the minimum image of a vector $\mathbf{r}$, we
1219 < first convert it to its corresponding vector in box space, and then,
1220 < cast each element to lie in the range $[-0.5,0.5]$:
1218 > directions. To find the minimum image of a vector $\mathbf{r}$, {\sc
1219 > oopse} first converts it to its corresponding vector in box space, and
1220 > then casts each element to lie in the range $[-0.5,0.5]$:
1221   \begin{equation}
1222   s_{i}^{\prime}=s_{i}-\roundme(s_{i}),
1223   \end{equation}
# Line 954 | Line 1233 | than $x$.  For example, $\roundme(3.6)=4$, $\roundme(3
1233   Here $\lfloor x \rfloor$ is the floor operator, and gives the largest
1234   integer value that is not greater than $x$, and $\lceil x \rceil$ is
1235   the ceiling operator, and gives the smallest integer that is not less
1236 < than $x$.  For example, $\roundme(3.6)=4$, $\roundme(3.1)=3$,
958 < $\roundme(-3.6)=-4$, $\roundme(-3.1)=-3$.
1236 > than $x$.
1237  
1238 < Finally, we obtain the minimum image coordinates $\mathbf{r}^{\prime}$ by
1239 < transforming back to real space,
1238 > Finally, the minimum image coordinates $\mathbf{r}^{\prime}$ are
1239 > obtained by transforming back to real space,
1240   \begin{equation}
1241   \mathbf{r}^{\prime}=\mathsf{H}^{-1}\mathbf{s}^{\prime}.%
1242   \end{equation}
1243   In this way, particles are allowed to diffuse freely in $\mathbf{r}$,
1244 < but their minimum images, $\mathbf{r}^{\prime}$ are used to compute
1244 > but their minimum images, or $\mathbf{r}^{\prime}$, are used to compute
1245   the inter-atomic forces.
1246  
1247  
# Line 971 | Line 1249 | DLM method}
1249   \section{\label{oopseSec:mechanics}Mechanics}
1250  
1251   \subsection{\label{oopseSec:integrate}Integrating the Equations of Motion: the
1252 < DLM method}
1252 > {\sc dlm} method}
1253  
1254   The default method for integrating the equations of motion in {\sc
1255   oopse} is a velocity-Verlet version of the symplectic splitting method
1256   proposed by Dullweber, Leimkuhler and McLachlan
1257 < (DLM).\cite{Dullweber1997} When there are no directional atoms or
1257 > ({\sc dlm}).\cite{Dullweber1997} When there are no directional atoms or
1258   rigid bodies present in the simulation, this integrator becomes the
1259   standard velocity-Verlet integrator which is known to sample the
1260   microcanonical (NVE) ensemble.\cite{Frenkel1996}
1261  
1262   Previous integration methods for orientational motion have problems
1263 < that are avoided in the DLM method.  Direct propagation of the Euler
1263 > that are avoided in the {\sc dlm} method.  Direct propagation of the Euler
1264   angles has a known $1/\sin\theta$ divergence in the equations of
1265 < motion for $\phi$ and $\psi$,\cite{allen87:csl} leading to
1266 < numerical instabilities any time one of the directional atoms or rigid
1267 < bodies has an orientation near $\theta=0$ or $\theta=\pi$.  More
1268 < modern quaternion-based integration methods have relatively poor
1269 < energy conservation.  While quaternions work well for orientational
1270 < motion in other ensembles, the microcanonical ensemble has a
1271 < constant energy requirement that is quite sensitive to errors in the
1272 < equations of motion.  An earlier implementation of {\sc oopse}
1273 < utilized quaternions for propagation of rotational motion; however, a
1274 < detailed investigation showed that they resulted in a steady drift in
997 < the total energy, something that has been observed by
998 < Laird {\it et al.}\cite{Laird97}      
1265 > motion for $\phi$ and $\psi$,\cite{allen87:csl} leading to numerical
1266 > instabilities any time one of the directional atoms or rigid bodies
1267 > has an orientation near $\theta=0$ or $\theta=\pi$.  Quaternion-based
1268 > integration methods work well for propagating orientational motion;
1269 > however, energy conservation concerns arise when using the
1270 > microcanonical (NVE) ensemble.  An earlier implementation of {\sc
1271 > oopse} utilized quaternions for propagation of rotational motion;
1272 > however, a detailed investigation showed that they resulted in a
1273 > steady drift in the total energy, something that has been observed by
1274 > Laird {\it et al.}\cite{Laird97}
1275  
1276   The key difference in the integration method proposed by Dullweber
1277   \emph{et al.} is that the entire $3 \times 3$ rotation matrix is
# Line 1069 | Line 1345 | is equivalent to the more familiar body-fixed forms,
1345   represented by ${\bf j}$.  This equation of motion for angular momenta
1346   is equivalent to the more familiar body-fixed forms,
1347   \begin{eqnarray}
1348 < \dot{j_{x}} & = & \tau^b_x(t)  +
1349 < \left(\overleftrightarrow{\mathsf{I}}_{yy} - \overleftrightarrow{\mathsf{I}}_{zz} \right) j_y j_z, \\
1350 < \dot{j_{y}} & = & \tau^b_y(t) +
1351 < \left(\overleftrightarrow{\mathsf{I}}_{zz} - \overleftrightarrow{\mathsf{I}}_{xx} \right) j_z j_x,\\
1352 < \dot{j_{z}} & = & \tau^b_z(t) +
1353 < \left(\overleftrightarrow{\mathsf{I}}_{xx} - \overleftrightarrow{\mathsf{I}}_{yy} \right) j_x j_y,
1348 > \dot{j_{x}} & = & \tau^b_x(t)  -
1349 > \left(\overleftrightarrow{\mathsf{I}}_{yy}^{-1} - \overleftrightarrow{\mathsf{I}}_{zz}^{-1} \right) j_y j_z, \\
1350 > \dot{j_{y}} & = & \tau^b_y(t) -
1351 > \left(\overleftrightarrow{\mathsf{I}}_{zz}^{-1} - \overleftrightarrow{\mathsf{I}}_{xx}^{-1} \right) j_z j_x,\\
1352 > \dot{j_{z}} & = & \tau^b_z(t) -
1353 > \left(\overleftrightarrow{\mathsf{I}}_{xx}^{-1} - \overleftrightarrow{\mathsf{I}}_{yy}^{-1} \right) j_x j_y,
1354   \end{eqnarray}
1355   which utilize the body-fixed torques, ${\bf \tau}^b$. Torques are
1356   most easily derived in the space-fixed frame,
# Line 1092 | Line 1368 | The DLM method uses a Trotter factorization of the ori
1368   Here $\hat{\bf u}$ is a unit vector pointing along the principal axis
1369   of the particle in the space-fixed frame.
1370  
1371 < The DLM method uses a Trotter factorization of the orientational
1371 > The {\sc dlm} method uses a Trotter factorization of the orientational
1372   propagator.  This has three effects:
1373   \begin{enumerate}
1374   \item the integrator is area-preserving in phase space (i.e. it is
# Line 1186 | Line 1462 | The matrix rotations used in the DLM method end up bei
1462          + \frac{h}{2} {\bf \tau}^b(t + h) .
1463   \end{align*}
1464  
1465 < The matrix rotations used in the DLM method end up being more costly
1466 < computationally than the simpler arithmetic quaternion
1467 < propagation. With the same time step, a 1000-molecule water simulation
1468 < shows an average 7\% increase in computation time using the DLM method
1469 < in place of quaternions. This cost is more than justified when
1470 < comparing the energy conservation of the two methods as illustrated in
1471 < Fig.~\ref{timestep}.
1465 > The matrix rotations used in the {\sc dlm} method end up being more
1466 > costly computationally than the simpler arithmetic quaternion
1467 > propagation. With the same time step, a 1024-molecule water simulation
1468 > incurs an average 12\% increase in computation time using the {\sc
1469 > dlm} method in place of quaternions. This cost is more than justified
1470 > when comparing the energy conservation achieved by the two
1471 > methods. Figure ~\ref{quatdlm} provides a comparative analysis of the
1472 > {\sc dlm} method versus the traditional quaternion scheme.
1473  
1474   \begin{figure}
1475   \centering
1476 < \includegraphics[width=\linewidth]{timeStep.pdf}
1477 < \caption[Energy conservation for quaternion versus DLM dynamics]{Energy conservation using quaternion based integration versus
1478 < the method proposed by Dullweber \emph{et al.} with increasing time
1479 < step. For each time step, the dotted line is total energy using the
1480 < DLM integrator, and the solid line comes from the quaternion
1481 < integrator. The larger time step plots are shifted up from the true
1482 < energy baseline for clarity.}
1483 < \label{timestep}
1476 > \includegraphics[width=\linewidth]{quatvsdlm.eps}
1477 > \caption[Energy conservation analysis of the {\sc dlm} and quaternion
1478 > integration methods]{Analysis of the energy conservation of the {\sc
1479 > dlm} and quaternion integration methods.  $\delta \mathrm{E}_1$ is the
1480 > linear drift in energy over time and $\delta \mathrm{E}_0$ is the
1481 > standard deviation of energy fluctuations around this drift.  All
1482 > simulations were of a 1024-molecule simulation of SSD water at 298 K
1483 > starting from the same initial configuration. Note that the {\sc dlm}
1484 > method provides more than an order of magnitude improvement in both
1485 > the energy drift and the size of the energy fluctuations when compared
1486 > with the quaternion method at any given time step.  At time steps
1487 > larger than 4 fs, the quaternion scheme resulted in rapidly rising
1488 > energies which eventually lead to simulation failure.  Using the {\sc
1489 > dlm} method, time steps up to 8 fs can be taken before this behavior
1490 > is evident.}
1491 > \label{quatdlm}
1492   \end{figure}
1493  
1494 < In Fig.~\ref{timestep}, the resulting energy drift at various time
1495 < steps for both the DLM and quaternion integration schemes is
1496 < compared. All of the 1000 molecule water simulations started with the
1497 < same configuration, and the only difference was the method for
1498 < handling rotational motion. At time steps of 0.1 and 0.5 fs, both
1499 < methods for propagating molecule rotation conserve energy fairly well,
1500 < with the quaternion method showing a slight energy drift over time in
1501 < the 0.5 fs time step simulation. At time steps of 1 and 2 fs, the
1502 < energy conservation benefits of the DLM method are clearly
1503 < demonstrated. Thus, while maintaining the same degree of energy
1504 < conservation, one can take considerably longer time steps, leading to
1505 < an overall reduction in computation time.
1494 > In Fig.~\ref{quatdlm}, $\delta \mbox{E}_1$ is a measure of the linear
1495 > energy drift in units of $\mbox{kcal mol}^{-1}$ per particle over a
1496 > nanosecond of simulation time, and $\delta \mbox{E}_0$ is the standard
1497 > deviation of the energy fluctuations in units of $\mbox{kcal
1498 > mol}^{-1}$ per particle. In the top plot, it is apparent that the
1499 > energy drift is reduced by a significant amount (2 to 3 orders of
1500 > magnitude improvement at all tested time steps) by chosing the {\sc
1501 > dlm} method over the simple non-symplectic quaternion integration
1502 > method.  In addition to this improvement in energy drift, the
1503 > fluctuations in the total energy are also dampened by 1 to 2 orders of
1504 > magnitude by utilizing the {\sc dlm} method.
1505 >
1506 > Although the {\sc dlm} method is more computationally expensive than
1507 > the traditional quaternion scheme for propagating a single time step,
1508 > consideration of the computational cost for a long simulation with a
1509 > particular level of energy conservation is in order.  A plot of energy
1510 > drift versus computational cost was generated
1511 > (Fig.~\ref{cpuCost}). This figure provides an estimate of the CPU time
1512 > required under the two integration schemes for 1 nanosecond of
1513 > simulation time for the model 1024-molecule system.  By chosing a
1514 > desired energy drift value it is possible to determine the CPU time
1515 > required for both methods. If a $\delta \mbox{E}_1$ of $1 \times
1516 > 10^{-3} \mbox{kcal mol}^{-1}$ per particle is desired, a nanosecond of
1517 > simulation time will require ~19 hours of CPU time with the {\sc dlm}
1518 > integrator, while the quaternion scheme will require ~154 hours of CPU
1519 > time. This demonstrates the computational advantage of the integration
1520 > scheme utilized in {\sc oopse}.
1521 >
1522 > \begin{figure}
1523 > \centering
1524 > \includegraphics[width=\linewidth]{compCost.eps}
1525 > \caption[Energy drift as a function of required simulation run
1526 > time]{Energy drift as a function of required simulation run time.
1527 > $\delta \mathrm{E}_1$ is the linear drift in energy over time.
1528 > Simulations were performed on a single 2.5 GHz Pentium 4
1529 > processor. Simulation time comparisons can be made by tracing
1530 > horizontally from one curve to the other. For example, a simulation
1531 > that takes ~24 hours using the {\sc dlm} method will take roughly 210
1532 > hours using the simple quaternion method if the same degree of energy
1533 > conservation is desired.}
1534 > \label{cpuCost}
1535 > \end{figure}
1536  
1537   There is only one specific keyword relevant to the default integrator,
1538   and that is the time step for integrating the equations of motion.
1539  
1540   \begin{center}
1541   \begin{tabular}{llll}
1542 < {\bf variable} & {\bf {\tt .bass} keyword} & {\bf units} & {\bf
1542 > {\bf variable} & {\bf Meta-data keyword} & {\bf units} & {\bf
1543   default value} \\  
1544   $h$ & {\tt dt = 2.0;} & fs & none
1545   \end{tabular}
# Line 1234 | Line 1549 | integrator can selected with the {\tt ensemble} keywor
1549  
1550   {\sc oopse} implements a number of extended system integrators for
1551   sampling from other ensembles relevant to chemical physics.  The
1552 < integrator can selected with the {\tt ensemble} keyword in the
1553 < {\tt .bass} file:
1552 > integrator can be selected with the {\tt ensemble} keyword in the
1553 > meta-data file:
1554  
1555   \begin{center}
1556   \begin{tabular}{lll}
1557 < {\bf Integrator} & {\bf Ensemble} & {\bf {\tt .bass} line} \\
1557 > {\bf Integrator} & {\bf Ensemble} & {\bf Meta-data instruction} \\
1558   NVE & microcanonical & {\tt ensemble = NVE; } \\
1559   NVT & canonical & {\tt ensemble = NVT; } \\
1560   NPTi & isobaric-isothermal & {\tt ensemble = NPTi;} \\
# Line 1254 | Line 1569 | system, and has been shown to sample the canonical dis
1569   The relatively well-known Nos\'e-Hoover thermostat\cite{Hoover85} is
1570   implemented in {\sc oopse}'s NVT integrator.  This method couples an
1571   extra degree of freedom (the thermostat) to the kinetic energy of the
1572 < system, and has been shown to sample the canonical distribution in the
1573 < system degrees of freedom while conserving a quantity that is, to
1572 > system and it has been shown to sample the canonical distribution in
1573 > the system degrees of freedom while conserving a quantity that is, to
1574   within a constant, the Helmholtz free energy.\cite{melchionna93}
1575  
1576   NPT algorithms attempt to maintain constant pressure in the system by
# Line 1279 | Line 1594 | variables.
1594  
1595   \begin{center}
1596   \begin{tabular}{llll}
1597 < {\bf variable} & {\bf {\tt .bass} keyword} & {\bf units} & {\bf
1597 > {\bf variable} & {\bf Meta-data instruction} & {\bf units} & {\bf
1598   default value} \\  
1599   $T_{\mathrm{target}}$ & {\tt targetTemperature = 300;} &  K & none \\
1600   $P_{\mathrm{target}}$ & {\tt targetPressure = 1;} & atm & none \\
# Line 1325 | Line 1640 | f = 3 N + 3 N_{\mathrm{orient}} - N_{\mathrm{constrain
1640   \end{equation}
1641   Here, $f$ is the total number of degrees of freedom in the system,
1642   \begin{equation}
1643 < f = 3 N + 3 N_{\mathrm{orient}} - N_{\mathrm{constraints}},
1643 > f = 3 N + 2 N_{\mathrm{linear}} + 3 N_{\mathrm{non-linear}} - N_{\mathrm{constraints}},
1644   \end{equation}
1645   and $K$ is the total kinetic energy,
1646   \begin{equation}
1647   K = \sum_{i=1}^{N} \frac{1}{2} m_i {\bf v}_i^T \cdot {\bf v}_i +
1648 < \sum_{i=1}^{N_{\mathrm{orient}}}  \frac{1}{2} {\bf j}_i^T \cdot
1648 > \sum_{i=1}^{N_{\mathrm{linear}}+N_{\mathrm{non-linear}}}  \frac{1}{2} {\bf j}_i^T \cdot
1649   \overleftrightarrow{\mathsf{I}}_i^{-1} \cdot {\bf j}_i.
1650   \end{equation}
1651 + $N_{\mathrm{linear}}$ is the number of linear rotors (i.e. with two
1652 + non-zero moments of inertia), and $N_{\mathrm{non-linear}}$ is the
1653 + number of non-linear rotors (i.e. with three non-zero moments of
1654 + inertia).  
1655  
1656   In eq.(\ref{eq:nosehooverext}), $\tau_T$ is the time constant for
1657   relaxation of the temperature to the target value.  To set values for
1658   $\tau_T$ or $T_{\mathrm{target}}$ in a simulation, one would use the
1659 < {\tt tauThermostat} and {\tt targetTemperature} keywords in the {\tt
1660 < .bass} file.  The units for {\tt tauThermostat} are fs, and the units
1661 < for the {\tt targetTemperature} are degrees K.   The integration of
1662 < the equations of motion is carried out in a velocity-Verlet style 2
1659 > {\tt tauThermostat} and {\tt targetTemperature} keywords in the
1660 > meta-data file.  The units for {\tt tauThermostat} are fs, and the
1661 > units for the {\tt targetTemperature} are degrees K.   The integration
1662 > of the equations of motion is carried out in a velocity-Verlet style 2
1663   part algorithm:
1664  
1665   {\tt moveA:}
# Line 1370 | Line 1689 | the section on the DLM integrator.  Note that this ope
1689   Here $\mathrm{rotate}(h * {\bf j}
1690   \overleftrightarrow{\mathsf{I}}^{-1})$ is the same symplectic Trotter
1691   factorization of the three rotation operations that was discussed in
1692 < the section on the DLM integrator.  Note that this operation modifies
1692 > the section on the {\sc dlm} integrator.  Note that this operation modifies
1693   both the rotation matrix $\mathsf{A}$ and the angular momentum ${\bf
1694   j}$.  {\tt moveA} propagates velocities by a half time step, and
1695   positional degrees of freedom by a full time step.  The new positions
1696   (and orientations) are then used to calculate a new set of forces and
1697   torques in exactly the same way they are calculated in the {\tt
1698 < doForces} portion of the DLM integrator.
1698 > doForces} portion of the {\sc dlm} integrator.
1699  
1700   Once the forces and torques have been obtained at the new time step,
1701   the temperature, velocities, and the extended system variable can be
# Line 1402 | Line 1721 | Since ${\bf v}(t + h)$ and ${\bf j}(t + h)$ are requir
1721          \chi(t + h) \right) .
1722   \end{align*}
1723  
1724 < Since ${\bf v}(t + h)$ and ${\bf j}(t + h)$ are required to caclculate
1724 > Since ${\bf v}(t + h)$ and ${\bf j}(t + h)$ are required to calculate
1725   $T(t + h)$ as well as $\chi(t + h)$, they indirectly depend on their
1726   own values at time $t + h$.  {\tt moveB} is therefore done in an
1727   iterative fashion until $\chi(t + h)$ becomes self-consistent.  The
# Line 1430 | Line 1749 | To carry out isobaric-isothermal ensemble calculations
1749   \subsection{\label{sec:NPTi}Constant-pressure integration with
1750   isotropic box deformations (NPTi)}
1751  
1752 < To carry out isobaric-isothermal ensemble calculations {\sc oopse}
1752 > To carry out isobaric-isothermal ensemble calculations, {\sc oopse}
1753   implements the Melchionna modifications to the Nos\'e-Hoover-Andersen
1754 < equations of motion,\cite{melchionna93}
1754 > equations of motion.\cite{melchionna93} The equations of motion are
1755 > the same as NVT with the following exceptions:
1756  
1757   \begin{eqnarray}
1758   \dot{{\bf r}} & = & {\bf v} + \eta \left( {\bf r} - {\bf R}_0 \right), \\
1759   \dot{{\bf v}} & = & \frac{{\bf f}}{m} - (\eta + \chi) {\bf v}, \\
1440 \dot{\mathsf{A}} & = & \mathsf{A} \cdot
1441 \mbox{ skew}\left(\overleftrightarrow{I}^{-1} \cdot {\bf j}\right),\\
1442 \dot{{\bf j}} & = & {\bf j} \times \left( \overleftrightarrow{I}^{-1}
1443 \cdot {\bf j} \right) - \mbox{ rot}\left(\mathsf{A}^{T} \cdot \frac{\partial
1444 V}{\partial \mathsf{A}} \right) - \chi {\bf j}, \\
1445 \dot{\chi} & = & \frac{1}{\tau_{T}^2} \left(
1446 \frac{T}{T_{\mathrm{target}}} - 1 \right) ,\\
1760   \dot{\eta} & = & \frac{1}{\tau_{B}^2 f k_B T_{\mathrm{target}}} V \left( P -
1761   P_{\mathrm{target}} \right), \\
1762   \dot{\mathcal{V}} & = & 3 \mathcal{V} \eta .
# Line 1470 | Line 1783 | outer} product of the velocities denoted by the $\otim
1783   \overleftrightarrow{\mathsf{W}}(t).
1784   \end{equation}
1785   The kinetic contribution to the pressure tensor utilizes the {\it
1786 < outer} product of the velocities denoted by the $\otimes$ symbol.  The
1786 > outer} product of the velocities, denoted by the $\otimes$ symbol.  The
1787   stress tensor is calculated from another outer product of the
1788   inter-atomic separation vectors (${\bf r}_{ij} = {\bf r}_j - {\bf
1789   r}_i$) with the forces between the same two atoms,
# Line 1478 | Line 1791 | The instantaneous pressure is then simply obtained fro
1791   \overleftrightarrow{\mathsf{W}}(t) = \sum_{i} \sum_{j>i} {\bf r}_{ij}(t)
1792   \otimes {\bf f}_{ij}(t).
1793   \end{equation}
1794 + In systems containing cutoff groups, the stress tensor is computed
1795 + between the centers-of-mass of the cutoff groups:
1796 + \begin{equation}
1797 + \overleftrightarrow{\mathsf{W}}(t) = \sum_{a} \sum_{b} {\bf r}_{ab}(t)
1798 + \otimes {\bf f}_{ab}(t).
1799 + \end{equation}
1800 + where ${\bf r}_{ab}$ is the distance between the centers of mass, and
1801 + \begin{equation}
1802 + {\bf f}_{ab} = s(r_{ab}) \sum_{i \in a} \sum_{j \in b} {\bf f}_{ij} +
1803 + s^{\prime}(r_{ab}) \frac{{\bf r}_{ab}}{|r_{ab}|} \sum_{i \in a} \sum_{j
1804 + \in b} V_{ij}({\bf r}_{ij}).
1805 + \end{equation}
1806 +
1807   The instantaneous pressure is then simply obtained from the trace of
1808 < the Pressure tensor,
1808 > the pressure tensor,
1809   \begin{equation}
1810 < P(t) = \frac{1}{3} \mathrm{Tr} \left( \overleftrightarrow{\mathsf{P}}(t).
1811 < \right)
1810 > P(t) = \frac{1}{3} \mathrm{Tr} \left( \overleftrightarrow{\mathsf{P}}(t)
1811 > \right).
1812   \end{equation}
1813  
1814   In eq.(\ref{eq:melchionna1}), $\tau_B$ is the time constant for
1815   relaxation of the pressure to the target value.  To set values for
1816   $\tau_B$ or $P_{\mathrm{target}}$ in a simulation, one would use the
1817 < {\tt tauBarostat} and {\tt targetPressure} keywords in the {\tt .bass}
1817 > {\tt tauBarostat} and {\tt targetPressure} keywords in the meta-data
1818   file.  The units for {\tt tauBarostat} are fs, and the units for the
1819   {\tt targetPressure} are atmospheres.  Like in the NVT integrator, the
1820   integration of the equations of motion is carried out in a
1821 < velocity-Verlet style 2 part algorithm:
1821 > velocity-Verlet style two part algorithm with only the following
1822 > differences:
1823  
1824   {\tt moveA:}
1825   \begin{align*}
1499 T(t) &\leftarrow \left\{{\bf v}(t)\right\}, \left\{{\bf j}(t)\right\} ,\\
1500 %
1826   P(t) &\leftarrow \left\{{\bf r}(t)\right\}, \left\{{\bf v}(t)\right\} ,\\
1827   %
1828   {\bf v}\left(t + h / 2\right)  &\leftarrow {\bf v}(t)
1829          + \frac{h}{2} \left( \frac{{\bf f}(t)}{m} - {\bf v}(t)
1830          \left(\chi(t) + \eta(t) \right) \right), \\
1831   %
1507 {\bf j}\left(t + h / 2 \right)  &\leftarrow {\bf j}(t)
1508        + \frac{h}{2} \left( {\bf \tau}^b(t) - {\bf j}(t)
1509        \chi(t) \right), \\
1510 %
1511 \mathsf{A}(t + h) &\leftarrow \mathrm{rotate}\left(h *
1512        {\bf j}(t + h / 2) \overleftrightarrow{\mathsf{I}}^{-1}
1513        \right) ,\\
1514 %
1515 \chi\left(t + h / 2 \right) &\leftarrow \chi(t) +
1516        \frac{h}{2 \tau_T^2} \left( \frac{T(t)}{T_{\mathrm{target}}} - 1
1517        \right) ,\\
1518 %
1832   \eta(t + h / 2) &\leftarrow \eta(t) + \frac{h
1833          \mathcal{V}(t)}{2 N k_B T(t) \tau_B^2} \left( P(t)
1834          - P_{\mathrm{target}} \right), \\
# Line 1529 | Line 1842 | Most of these equations are identical to their counter
1842          \mathsf{H}(t).
1843   \end{align*}
1844  
1845 < Most of these equations are identical to their counterparts in the NVT
1533 < integrator, but the propagation of positions to time $t + h$
1845 > The propagation of positions to time $t + h$
1846   depends on the positions at the same time.  {\sc oopse} carries out
1847   this step iteratively (with a limit of 5 passes through the iterative
1848   loop).  Also, the simulation box $\mathsf{H}$ is scaled uniformly for
# Line 1539 | Line 1851 | the box by
1851   h / 2$.  Reshaping the box uniformly also scales the volume of
1852   the box by
1853   \begin{equation}
1854 < \mathcal{V}(t + h) \leftarrow e^{ - 3 h \eta(t + h /2)}.
1855 < \mathcal{V}(t)
1854 > \mathcal{V}(t + h) \leftarrow e^{ - 3 h \eta(t + h /2)} \times
1855 > \mathcal{V}(t).
1856   \end{equation}
1857  
1858   The {\tt doForces} step for the NPTi integrator is exactly the same as
1859 < in both the DLM and NVT integrators.  Once the forces and torques have
1859 > in both the {\sc dlm} and NVT integrators.  Once the forces and torques have
1860   been obtained at the new time step, the velocities can be advanced to
1861   the same time value.
1862  
1863   {\tt moveB:}
1864   \begin{align*}
1553 T(t + h) &\leftarrow \left\{{\bf v}(t + h)\right\},
1554        \left\{{\bf j}(t + h)\right\} ,\\
1555 %
1865   P(t + h) &\leftarrow  \left\{{\bf r}(t + h)\right\},
1866          \left\{{\bf v}(t + h)\right\}, \\
1867   %
1559 \chi\left(t + h \right) &\leftarrow \chi\left(t + h /
1560        2 \right) + \frac{h}{2 \tau_T^2} \left( \frac{T(t+h)}
1561        {T_{\mathrm{target}}} - 1 \right), \\
1562 %
1868   \eta(t + h) &\leftarrow \eta(t + h / 2) +
1869          \frac{h \mathcal{V}(t + h)}{2 N k_B T(t + h)
1870          \tau_B^2} \left( P(t + h) - P_{\mathrm{target}} \right), \\
# Line 1576 | Line 1881 | to caclculate $T(t + h)$, $P(t + h)$, $\chi(t + h)$, a
1881   \end{align*}
1882  
1883   Once again, since ${\bf v}(t + h)$ and ${\bf j}(t + h)$ are required
1884 < to caclculate $T(t + h)$, $P(t + h)$, $\chi(t + h)$, and $\eta(t +
1884 > to calculate $T(t + h)$, $P(t + h)$, $\chi(t + h)$, and $\eta(t +
1885   h)$, they indirectly depend on their own values at time $t + h$.  {\tt
1886   moveB} is therefore done in an iterative fashion until $\chi(t + h)$
1887   and $\eta(t + h)$ become self-consistent.  The relative tolerance for
# Line 1616 | Line 1921 | the box shape.  The equations of motion for this metho
1921   {\it shape} as well as in the volume of the box.  This method utilizes
1922   the full $3 \times 3$ pressure tensor and introduces a tensor of
1923   extended variables ($\overleftrightarrow{\eta}$) to control changes to
1924 < the box shape.  The equations of motion for this method are
1924 > the box shape.  The equations of motion for this method differ from
1925 > those of NPTi as follows:
1926   \begin{eqnarray}
1927   \dot{{\bf r}} & = & {\bf v} + \overleftrightarrow{\eta} \cdot \left( {\bf r} - {\bf R}_0 \right), \\
1928   \dot{{\bf v}} & = & \frac{{\bf f}}{m} - (\overleftrightarrow{\eta} +
1929   \chi \cdot \mathsf{1}) {\bf v}, \\
1624 \dot{\mathsf{A}} & = & \mathsf{A} \cdot
1625 \mbox{ skew}\left(\overleftrightarrow{I}^{-1} \cdot {\bf j}\right) ,\\
1626 \dot{{\bf j}} & = & {\bf j} \times \left( \overleftrightarrow{I}^{-1}
1627 \cdot {\bf j} \right) - \mbox{ rot}\left(\mathsf{A}^{T} \cdot \frac{\partial
1628 V}{\partial \mathsf{A}} \right) - \chi {\bf j} ,\\
1629 \dot{\chi} & = & \frac{1}{\tau_{T}^2} \left(
1630 \frac{T}{T_{\mathrm{target}}} - 1 \right) ,\\
1930   \dot{\overleftrightarrow{\eta}} & = & \frac{1}{\tau_{B}^2 f k_B
1931   T_{\mathrm{target}}} V \left( \overleftrightarrow{\mathsf{P}} - P_{\mathrm{target}}\mathsf{1} \right) ,\\
1932   \dot{\mathsf{H}} & = &  \overleftrightarrow{\eta} \cdot \mathsf{H} .
# Line 1643 | Line 1942 | T(t) &\leftarrow \left\{{\bf v}(t)\right\}, \left\{{\b
1942  
1943   {\tt moveA:}
1944   \begin{align*}
1646 T(t) &\leftarrow \left\{{\bf v}(t)\right\}, \left\{{\bf j}(t)\right\} ,\\
1647 %
1945   \overleftrightarrow{\mathsf{P}}(t) &\leftarrow \left\{{\bf r}(t)\right\},
1946          \left\{{\bf v}(t)\right\} ,\\
1947   %
# Line 1653 | Line 1950 | T(t) &\leftarrow \left\{{\bf v}(t)\right\}, \left\{{\b
1950          \left(\chi(t)\mathsf{1} + \overleftrightarrow{\eta}(t) \right) \cdot
1951          {\bf v}(t) \right), \\
1952   %
1656 {\bf j}\left(t + h / 2 \right)  &\leftarrow {\bf j}(t)
1657        + \frac{h}{2} \left( {\bf \tau}^b(t) - {\bf j}(t)
1658        \chi(t) \right), \\
1659 %
1660 \mathsf{A}(t + h) &\leftarrow \mathrm{rotate}\left(h *
1661        {\bf j}(t + h / 2) \overleftrightarrow{\mathsf{I}}^{-1}
1662        \right), \\
1663 %
1664 \chi\left(t + h / 2 \right) &\leftarrow \chi(t) +
1665        \frac{h}{2 \tau_T^2} \left( \frac{T(t)}{T_{\mathrm{target}}}
1666        - 1 \right), \\
1667 %
1953   \overleftrightarrow{\eta}(t + h / 2) &\leftarrow
1954          \overleftrightarrow{\eta}(t) + \frac{h \mathcal{V}(t)}{2 N k_B
1955          T(t) \tau_B^2} \left( \overleftrightarrow{\mathsf{P}}(t)
# Line 1686 | Line 1971 | T(t + h) &\leftarrow \left\{{\bf v}(t + h)\right\},
1971  
1972   {\tt moveB:}
1973   \begin{align*}
1689 T(t + h) &\leftarrow \left\{{\bf v}(t + h)\right\},
1690        \left\{{\bf j}(t + h)\right\}, \\
1691 %
1974   \overleftrightarrow{\mathsf{P}}(t + h) &\leftarrow \left\{{\bf r}
1975          (t + h)\right\}, \left\{{\bf v}(t
1976          + h)\right\}, \left\{{\bf f}(t + h)\right\} ,\\
1977   %
1696 \chi\left(t + h \right) &\leftarrow \chi\left(t + h /
1697        2 \right) + \frac{h}{2 \tau_T^2} \left( \frac{T(t+
1698        h)}{T_{\mathrm{target}}} - 1 \right), \\
1699 %
1978   \overleftrightarrow{\eta}(t + h) &\leftarrow
1979          \overleftrightarrow{\eta}(t + h / 2) +
1980          \frac{h \mathcal{V}(t + h)}{2 N k_B T(t + h)
# Line 1708 | Line 1986 | T(t + h) &\leftarrow \left\{{\bf v}(t + h)\right\},
1986          \frac{{\bf f}(t + h)}{m} -
1987          (\chi(t + h)\mathsf{1} + \overleftrightarrow{\eta}(t
1988          + h)) \right) \cdot {\bf v}(t + h), \\
1711 %
1712 {\bf j}\left(t + h \right)  &\leftarrow {\bf j}\left(t
1713        + h / 2 \right) + \frac{h}{2} \left( {\bf \tau}^b(t
1714        + h) - {\bf j}(t + h) \chi(t + h) \right) .
1989   \end{align*}
1990  
1991   The iterative schemes for both {\tt moveA} and {\tt moveB} are
# Line 1729 | Line 2003 | elongated and sheared geometries which become smaller
2003   This integrator must be used with care, particularly in liquid
2004   simulations.  Liquids have very small restoring forces in the
2005   off-diagonal directions, and the simulation box can very quickly form
2006 < elongated and sheared geometries which become smaller than the
2007 < electrostatic or Lennard-Jones cutoff radii.  The NPTf integrator
2008 < finds most use in simulating crystals or liquid crystals which assume
1735 < non-orthorhombic geometries.
2006 > elongated and sheared geometries which become smaller than the cutoff
2007 > radius.  The NPTf integrator finds most use in simulating crystals or
2008 > liquid crystals which assume non-orthorhombic geometries.
2009  
2010   \subsection{\label{nptxyz}Constant pressure in 3 axes (NPTxyz)}
2011  
# Line 1762 | Line 2035 | Andersen.\cite{andersen83} The algorithm is a velocity
2035  
2036   In order to satisfy the constraints of fixed bond lengths within {\sc
2037   oopse}, we have implemented the {\sc rattle} algorithm of
2038 < Andersen.\cite{andersen83} The algorithm is a velocity verlet
2039 < formulation of the {\sc shake} method\cite{ryckaert77} of iteratively
2040 < solving the Lagrange multipliers of constraint.
2038 > Andersen.\cite{andersen83} {\sc rattle} is a velocity-Verlet
2039 > formulation of the {\sc shake} method\cite{ryckaert77} for iteratively
2040 > solving the Lagrange multipliers which maintain the holonomic
2041 > constraints.  Both methods are covered in depth in the
2042 > literature,\cite{leach01:mm,allen87:csl} and a detailed description of
2043 > this method would be redundant.
2044  
2045 < \subsubsection{\label{oopseSec:zcons}Z-Constraint Method}
2045 > \subsubsection{\label{oopseSec:zcons}The Z-Constraint Method}
2046  
2047 < Based on the fluctuation-dissipation theorem, a force auto-correlation
2048 < method was developed by Roux and Karplus to investigate the dynamics
2047 > A force auto-correlation method based on the fluctuation-dissipation
2048 > theorem was developed by Roux and Karplus to investigate the dynamics
2049   of ions inside ion channels.\cite{Roux91} The time-dependent friction
2050   coefficient can be calculated from the deviation of the instantaneous
2051 < force from its mean force.
2051 > force from its mean value:
2052   \begin{equation}
2053   \xi(z,t)=\langle\delta F(z,t)\delta F(z,0)\rangle/k_{B}T,
2054   \end{equation}
# Line 1781 | Line 2057 | where%
2057   \delta F(z,t)=F(z,t)-\langle F(z,t)\rangle.
2058   \end{equation}
2059  
1784
2060   If the time-dependent friction decays rapidly, the static friction
2061   coefficient can be approximated by
2062   \begin{equation}
2063   \xi_{\text{static}}(z)=\int_{0}^{\infty}\langle\delta F(z,t)\delta F(z,0)\rangle dt.
2064   \end{equation}
2065 < Allowing diffusion constant to then be calculated through the
2065 >
2066 > This allows the diffusion constant to then be calculated through the
2067   Einstein relation:\cite{Marrink94}
2068   \begin{equation}
2069   D(z)=\frac{k_{B}T}{\xi_{\text{static}}(z)}=\frac{(k_{B}T)^{2}}{\int_{0}^{\infty
2070   }\langle\delta F(z,t)\delta F(z,0)\rangle dt}.%
2071   \end{equation}
2072  
2073 < The Z-Constraint method, which fixes the z coordinates of the
2074 < molecules with respect to the center of the mass of the system, has
2075 < been a method suggested to obtain the forces required for the force
2076 < auto-correlation calculation.\cite{Marrink94} However, simply resetting the
2077 < coordinate will move the center of the mass of the whole system. To
2078 < avoid this problem, a new method was used in {\sc oopse}. Instead of
2079 < resetting the coordinate, we reset the forces of z-constrained
2080 < molecules as well as subtract the total constraint forces from the
2081 < rest of the system after the force calculation at each time step.
2073 > The Z-Constraint method, which fixes the $z$ coordinates of a few
2074 > ``tagged'' molecules with respect to the center of the mass of the
2075 > system is a technique that was proposed to obtain the forces required
2076 > for the force auto-correlation calculation.\cite{Marrink94} However,
2077 > simply resetting the coordinate will move the center of the mass of
2078 > the whole system. To avoid this problem, we have developed a new
2079 > method that is utilized in {\sc oopse}. Instead of resetting the
2080 > coordinates, we reset the forces of $z$-constrained molecules and
2081 > subtract the total constraint forces from the rest of the system after
2082 > the force calculation at each time step.
2083  
2084 < After the force calculation, define $G_\alpha$ as
2084 > After the force calculation, the total force on molecule $\alpha$ is:
2085   \begin{equation}
2086   G_{\alpha} = \sum_i F_{\alpha i},
2087   \label{oopseEq:zc1}
2088   \end{equation}
2089 < where $F_{\alpha i}$ is the force in the z direction of atom $i$ in
2090 < z-constrained molecule $\alpha$. The forces of the z constrained
2091 < molecule are then set to:
2089 > where $F_{\alpha i}$ is the force in the $z$ direction on atom $i$ in
2090 > $z$-constrained molecule $\alpha$. The forces on the atoms in the
2091 > $z$-constrained molecule are then adjusted to remove the total force
2092 > on molecule $\alpha$:
2093   \begin{equation}
2094   F_{\alpha i} = F_{\alpha i} -
2095          \frac{m_{\alpha i} G_{\alpha}}{\sum_i m_{\alpha i}}.
2096   \end{equation}
2097 < Here, $m_{\alpha i}$ is the mass of atom $i$ in the z-constrained
2098 < molecule. Having rescaled the forces, the velocities must also be
2099 < rescaled to subtract out any center of mass velocity in the z
2100 < direction.
2097 > Here, $m_{\alpha i}$ is the mass of atom $i$ in the $z$-constrained
2098 > molecule.  After the forces have been adjusted, the velocities must
2099 > also be modified to subtract out molecule $\alpha$'s center-of-mass
2100 > velocity in the $z$ direction.
2101   \begin{equation}
2102   v_{\alpha i} = v_{\alpha i} -
2103          \frac{\sum_i m_{\alpha i} v_{\alpha i}}{\sum_i m_{\alpha i}},
2104   \end{equation}
2105 < where $v_{\alpha i}$ is the velocity of atom $i$ in the z direction.
2106 < Lastly, all of the accumulated z constrained forces must be subtracted
2107 < from the system to keep the system center of mass from drifting.
2105 > where $v_{\alpha i}$ is the velocity of atom $i$ in the $z$ direction.
2106 > Lastly, all of the accumulated constraint forces must be subtracted
2107 > from the rest of the unconstrained system to keep the system center of
2108 > mass of the entire system from drifting.
2109   \begin{equation}
2110   F_{\beta i} = F_{\beta i} - \frac{m_{\beta i} \sum_{\alpha} G_{\alpha}}
2111          {\sum_{\beta}\sum_i m_{\beta i}},
2112   \end{equation}
2113 < where $\beta$ are all of the unconstrained molecules in the
2113 > where $\beta$ denotes all {\it unconstrained} molecules in the
2114   system. Similarly, the velocities of the unconstrained molecules must
2115 < also be scaled.
2115 > also be scaled:
2116   \begin{equation}
2117 < v_{\beta i} = v_{\beta i} + \sum_{\alpha}
2118 <        \frac{\sum_i m_{\alpha i} v_{\alpha i}}{\sum_i m_{\alpha i}}.
2117 > v_{\beta i} = v_{\beta i} + \sum_{\alpha} \frac{\sum_i m_{\alpha i}
2118 > v_{\alpha i}}{\sum_i m_{\alpha i}}.
2119   \end{equation}
2120  
2121 < At the very beginning of the simulation, the molecules may not be at their
2122 < constrained positions. To move a z-constrained molecule to its specified
2123 < position, a simple harmonic potential is used
2121 > This method will pin down the centers-of-mass of all of the
2122 > $z$-constrained molecules, and will also keep the entire system fixed
2123 > at the original system center-of-mass location.
2124 >
2125 > At the very beginning of the simulation, the molecules may not be at
2126 > their desired positions. To steer a $z$-constrained molecule to its
2127 > specified position, a simple harmonic potential is used:
2128   \begin{equation}
2129   U(t)=\frac{1}{2}k_{\text{Harmonic}}(z(t)-z_{\text{cons}})^{2},%
2130   \end{equation}
2131 < where $k_{\text{Harmonic}}$ is the harmonic force constant, $z(t)$ is the
2132 < current $z$ coordinate of the center of mass of the constrained molecule, and
2133 < $z_{\text{cons}}$ is the constrained position. The harmonic force operating
2134 < on the z-constrained molecule at time $t$ can be calculated by
2131 > where $k_{\text{Harmonic}}$ is an harmonic force constant, $z(t)$ is
2132 > the current $z$ coordinate of the center of mass of the constrained
2133 > molecule, and $z_{\text{cons}}$ is the desired constraint
2134 > position. The harmonic force operating on the $z$-constrained molecule
2135 > at time $t$ can be calculated by
2136   \begin{equation}
2137   F_{z_{\text{Harmonic}}}(t)=-\frac{\partial U(t)}{\partial z(t)}=
2138          -k_{\text{Harmonic}}(z(t)-z_{\text{cons}}).
2139   \end{equation}
2140  
2141 < \section{\label{oopseSec:design}Program Design}
2141 > The user may also specify the use of a constant velocity method
2142 > (steered molecular dynamics) to move the molecules to their desired
2143 > initial positions. Based on concepts from atomic force microscopy,
2144 > {\sc smd} has been used to study many processes which occur via rare
2145 > events on the time scale of a few hundreds of picoseconds.  For
2146 > example,{\sc smd} has been used to observe the dissociation of
2147 > Streptavidin-biotin Complex.\cite{smd}  
2148  
2149 < \subsection{\label{sec:architecture} {\sc oopse} Architecture}
2149 > To use of the $z$-constraint method in an {\sc oopse} simulation, the
2150 > molecules must be specified using the {\tt nZconstraints} keyword in
2151 > the meta-data file.  The other parameters for modifying the behavior
2152 > of the $z$-constraint method are listed in table~\ref{table:zconParams}.
2153  
2154 < The core of OOPSE is divided into two main object libraries:
2155 < \texttt{libBASS} and \texttt{libmdtools}. \texttt{libBASS} is the
2156 < library developed around the parsing engine and \texttt{libmdtools}
2157 < is the software library developed around the simulation engine. These
2158 < two libraries are designed to encompass all the basic functions and
2159 < tools that {\sc oopse} provides. Utility programs, such as the
2160 < property analyzers, need only link against the software libraries to
2161 < gain access to parsing, force evaluation, and input / output
2162 < routines.
2154 > \begin{table}
2155 > \caption{Meta-data Keywords: Z-Constraint Parameters}
2156 > \label{table:zconParams}
2157 > \begin{center}
2158 > % Note when adding or removing columns, the \hsize numbers must add up to the total number
2159 > % of columns.
2160 > \begin{tabularx}{\linewidth}%
2161 >  {>{\setlength{\hsize}{1.00\hsize}}X%
2162 >  >{\setlength{\hsize}{0.4\hsize}}X%
2163 >  >{\setlength{\hsize}{1.2\hsize}}X%
2164 >  >{\setlength{\hsize}{1.4\hsize}}X}
2165  
2166 < Contained in \texttt{libBASS} are all the routines associated with
1872 < reading and parsing the \texttt{.bass} input files. Given a
1873 < \texttt{.bass} file, \texttt{libBASS} will open it and any associated
1874 < \texttt{.mdl} files; then create structures in memory that are
1875 < templates of all the molecules specified in the input files. In
1876 < addition, any simulation parameters set in the \texttt{.bass} file
1877 < will be placed in a structure for later query by the controlling
1878 < program.
2166 > {\bf keyword} & {\bf units} & {\bf use} & {\bf remarks} \\ \hline
2167  
2168 < Located in \texttt{libmdtools} are all other routines necessary to a
2169 < Molecular Dynamics simulation. The library uses the main data
2170 < structures returned by \texttt{libBASS} to initialize the various
2171 < parts of the simulation: the atom structures and positions, the force
2172 < field, the integrator, \emph{et cetera}. After initialization, the
2173 < library can be used to perform a variety of tasks: integrate a
2174 < Molecular Dynamics trajectory, query phase space information from a
2175 < specific frame of a completed trajectory, or even recalculate force or
2176 < energetic information about specific frames from a completed
2177 < trajectory.
2168 > {\tt nZconstraints} & integer &  The number of $z$-constrained
2169 > molecules & If using the $z$-constraint method, {\tt nZconstraints}
2170 > must be set \\
2171 > {\tt zconsTime} & fs & Sets the frequency at which the {\tt .fz} file
2172 > is written &  \\
2173 > {\tt zconsForcePolicy} & string & The strategy for subtracting
2174 > the $z$-constraint force from the {\it unconstrained} molecules & Possible
2175 > strategies are {\tt BYMASS} and {\tt BYNUMBER}. The default
2176 > strategy is {\tt BYMASS}\\
2177 > {\tt zconsGap} & $\mbox{\AA}$ & Sets the distance between two adjacent
2178 > constraint positions&Used mainly to move molecules through a
2179 > simulation to estimate potentials of mean force. \\
2180 > {\tt zconsFixtime} & fs & Sets the length of time the $z$-constraint
2181 > molecule is held fixed & {\tt zconsFixtime} must be set if {\tt
2182 > zconsGap} is set\\
2183 > {\tt zconsUsingSMD} & logical & Flag for using Steered Molecular
2184 > Dynamics to move the molecules to the correct constrained positions  &
2185 > Harmonic Forces are used by default\\
2186  
2187 < With these core libraries in place, several programs have been
2188 < developed to utilize the routines provided by \texttt{libBASS} and
2189 < \texttt{libmdtools}. The main program of the package is \texttt{oopse}
1894 < and the corresponding parallel version \texttt{oopse\_MPI}. These two
1895 < programs will take the \texttt{.bass} file, and create (and integrate)
1896 < the simulation specified in the script. The two analysis programs
1897 < \texttt{staticProps} and \texttt{dynamicProps} utilize the core
1898 < libraries to initialize and read in trajectories from previously
1899 < completed simulations, in addition to the ability to use functionality
1900 < from \texttt{libmdtools} to recalculate forces and energies at key
1901 < frames in the trajectories. Lastly, the family of system building
1902 < programs (Sec.~\ref{oopseSec:initCoords}) also use the libraries to
1903 < store and output the system configurations they create.
2187 > \end{tabularx}
2188 > \end{center}
2189 > \end{table}
2190  
1905 \subsection{\label{oopseSec:parallelization} Parallelization of {\sc oopse}}
2191  
2192 < Although processor power is continually growing roughly following
1908 < Moore's Law, it is still unreasonable to simulate systems of more then
1909 < a 1000 atoms on a single processor. To facilitate study of larger
1910 < system sizes or smaller systems on long time scales in a reasonable
1911 < period of time, parallel methods were developed allowing multiple
1912 < CPU's to share the simulation workload. Three general categories of
1913 < parallel decomposition methods have been developed including atomic,
1914 < spatial and force decomposition methods.
2192 > \section{\label{oopseSec:minimizer}Energy Minimization}
2193  
2194 < Algorithmically simplest of the three methods is atomic decomposition
2195 < where N particles in a simulation are split among P processors for the
2196 < duration of the simulation. Computational cost scales as an optimal
2197 < $\mathcal{O}(N/P)$ for atomic decomposition. Unfortunately all
2198 < processors must communicate positions and forces with all other
2199 < processors at every force evaluation, leading communication costs to
2200 < scale as an unfavorable $\mathcal{O}(N)$, \emph{independent of the
2194 > As one of the basic procedures of molecular modeling, energy
2195 > minimization is used to identify local configurations that are stable
2196 > points on the potential energy surface. There is a vast literature on
2197 > energy minimization algorithms have been developed to search for the
2198 > global energy minimum as well as to find local structures which are
2199 > stable fixed points on the surface.  We have included two simple
2200 > minimization algorithms: steepest descent, ({\sc sd}) and conjugate
2201 > gradient ({\sc cg}) to help users find reasonable local minima from
2202 > their initial configurations. Since {\sc oopse} handles atoms and
2203 > rigid bodies which have orientational coordinates as well as
2204 > translational coordinates, there is some subtlety to the choice of
2205 > parameters for minimization algorithms.
2206 >
2207 > Given a coordinate set $x_{k}$ and a search direction $d_{k}$, a line
2208 > search algorithm is performed along $d_{k}$ to produce
2209 > $x_{k+1}=x_{k}+$ $\lambda _{k}d_{k}$. In the steepest descent ({\sc
2210 > sd}) algorithm,%
2211 > \begin{equation}
2212 > d_{k}=-\nabla V(x_{k}).
2213 > \end{equation}
2214 > The gradient and the direction of next step are always orthogonal.
2215 > This may cause oscillatory behavior in narrow valleys.  To overcome
2216 > this problem, the Fletcher-Reeves variant~\cite{FletcherReeves} of the
2217 > conjugate gradient ({\sc cg}) algorithm is used to generate $d_{k+1}$
2218 > via simple recursion:
2219 > \begin{equation}
2220 > d_{k+1}  =-\nabla V(x_{k+1})+\gamma_{k}d_{k}
2221 > \end{equation}
2222 > where
2223 > \begin{equation}
2224 > \gamma_{k}  =\frac{\nabla V(x_{k+1})^{T}\nabla V(x_{k+1})}{\nabla
2225 > V(x_{k})^{T}\nabla V(x_{k})}.
2226 > \end{equation}
2227 >
2228 > The Polak-Ribiere variant~\cite{PolakRibiere} of the conjugate
2229 > gradient ($\gamma_{k}$) is defined as%
2230 > \begin{equation}
2231 > \gamma_{k}=\frac{[\nabla V(x_{k+1})-\nabla V(x)]^{T}\nabla V(x_{k+1})}{\nabla
2232 > V(x_{k})^{T}\nabla V(x_{k})}%
2233 > \end{equation}
2234 > It is widely agreed that the Polak-Ribiere variant gives better
2235 > convergence than the Fletcher-Reeves variant, so the conjugate
2236 > gradient approach implemented in {\sc oopse} is the Polak-Ribiere
2237 > variant.
2238 >
2239 > The conjugate gradient method assumes that the conformation is close
2240 > enough to a local minimum that the potential energy surface is very
2241 > nearly quadratic.  When the initial structure is far from the minimum,
2242 > the steepest descent method can be superior to the conjugate gradient
2243 > method. Hence, the steepest descent method is often used for the first
2244 > 10-100 steps of minimization. Another useful feature of minimization
2245 > methods in {\sc oopse} is that a modified {\sc shake} algorithm can be
2246 > applied during the minimization to constraint the bond lengths if this
2247 > is required by the force field. Meta-data parameters concerning the
2248 > minimizer are given in Table~\ref{table:minimizeParams}
2249 >
2250 > \begin{table}
2251 > \caption{Meta-data Keywords: Energy Minimizer Parameters}
2252 > \label{table:minimizeParams}
2253 > \begin{center}
2254 > % Note when adding or removing columns, the \hsize numbers must add up to the total number
2255 > % of columns.
2256 > \begin{tabularx}{\linewidth}%
2257 >  {>{\setlength{\hsize}{1.2\hsize}}X%
2258 >  >{\setlength{\hsize}{0.6\hsize}}X%
2259 >  >{\setlength{\hsize}{1.1\hsize}}X%
2260 >  >{\setlength{\hsize}{1.1\hsize}}X}
2261 >
2262 > {\bf keyword} & {\bf units} & {\bf use} & {\bf remarks} \\ \hline
2263 >
2264 > {\tt minimizer} & string &  selects the minimization method to be used
2265 > & either {\tt CG} (conjugate gradient) or {\tt SD} (steepest
2266 > descent) \\
2267 > {\tt minimizerMaxIter} & steps & Sets the maximum number of iterations
2268 > for the energy minimization & The default value is 200\\
2269 > {\tt minimizerWriteFreq} & steps & Sets the frequency with which the {\tt .dump} and {\tt .stat} files are writtern during energy minimization & \\
2270 > {\tt minimizerStepSize} & $\mbox{\AA}$ &  Sets the step size for the
2271 > line search & The default value is 0.01\\
2272 > {\tt minimizerFTol} & $\mbox{kcal mol}^{-1}$  & Sets the energy tolerance
2273 > for stopping the minimziation. & The default value is $10^{-8}$\\
2274 > {\tt minimizerGTol} & $\mbox{kcal mol}^{-1}\mbox{\AA}^{-1}$ & Sets the
2275 > gradient tolerance for stopping the minimization. & The default value
2276 > is  $10^{-8}$\\
2277 > {\tt minimizerLSTol} &  $\mbox{kcal mol}^{-1}$ & Sets line search
2278 > tolerance for terminating each step of the minimization. & The default
2279 > value is $10^{-8}$\\
2280 > {\tt minimizerLSMaxIter} & steps &  Sets the maximum number of
2281 > iterations for each line search & The default value is 50\\
2282 >
2283 > \end{tabularx}
2284 > \end{center}
2285 > \end{table}
2286 >
2287 > \section{\label{oopseSec:parallelization} Parallel Simulation Implementation}
2288 >
2289 > Although processor power is continually improving, it is still
2290 > unreasonable to simulate systems of more than 10,000 atoms on a single
2291 > processor. To facilitate study of larger system sizes or smaller
2292 > systems for longer time scales, parallel methods were developed to
2293 > allow multiple CPU's to share the simulation workload. Three general
2294 > categories of parallel decomposition methods have been developed:
2295 > these are the atomic,\cite{Fox88} spatial~\cite{plimpton95} and
2296 > force~\cite{Paradyn} decomposition methods.
2297 >
2298 > Algorithmically simplest of the three methods is atomic decomposition,
2299 > where $N$ particles in a simulation are split among $P$ processors for
2300 > the duration of the simulation. Computational cost scales as an
2301 > optimal $\mathcal{O}(N/P)$ for atomic decomposition. Unfortunately, all
2302 > processors must communicate positions and forces with all other
2303 > processors at every force evaluation, leading the communication costs
2304 > to scale as an unfavorable $\mathcal{O}(N)$, \emph{independent of the
2305   number of processors}. This communication bottleneck led to the
2306 < development of spatial and force decomposition methods in which
2306 > development of spatial and force decomposition methods, in which
2307   communication among processors scales much more favorably. Spatial or
2308   domain decomposition divides the physical spatial domain into 3D boxes
2309   in which each processor is responsible for calculation of forces and
2310   positions of particles located in its box. Particles are reassigned to
2311   different processors as they move through simulation space. To
2312 < calculate forces on a given particle, a processor must know the
2312 > calculate forces on a given particle, a processor must simply know the
2313   positions of particles within some cutoff radius located on nearby
2314 < processors instead of the positions of particles on all
2314 > processors rather than the positions of particles on all
2315   processors. Both communication between processors and computation
2316   scale as $\mathcal{O}(N/P)$ in the spatial method. However, spatial
2317   decomposition adds algorithmic complexity to the simulation code and
2318 < is not very efficient for small N since the overall communication
2318 > is not very efficient for small $N$, since the overall communication
2319   scales as the surface to volume ratio $\mathcal{O}(N/P)^{2/3}$ in
2320   three dimensions.
2321  
2322   The parallelization method used in {\sc oopse} is the force
2323 < decomposition method.  Force decomposition assigns particles to
2324 < processors based on a block decomposition of the force
2323 > decomposition method.\cite{hendrickson:95} Force decomposition assigns
2324 > particles to processors based on a block decomposition of the force
2325   matrix. Processors are split into an optimally square grid forming row
2326   and column processor groups. Forces are calculated on particles in a
2327 < given row by particles located in that processors column
2328 < assignment. Force decomposition is less complex to implement than the
2329 < spatial method but still scales computationally as $\mathcal{O}(N/P)$
2330 < and scales as $\mathcal{O}(N/\sqrt{P})$ in communication
2331 < cost. Plimpton has also found that force decompositions scale more
2332 < favorably than spatial decompositions for systems up to 10,000 atoms
2333 < and favorably compete with spatial methods up to 100,000
2334 < atoms.\cite{plimpton95}
2335 <
2327 > given row by particles located in that processor's column
2328 > assignment. One deviation from the algorithm described by Hendrickson
2329 > {\it et al.} is the use of column ordering based on the row indexes
2330 > preventing the need for a transpose operation necessitating a second
2331 > communication step when gathering the final force components.  Force
2332 > decomposition is less complex to implement than the spatial method but
2333 > still scales computationally as $\mathcal{O}(N/P)$ and scales as
2334 > $\mathcal{O}(N/\sqrt{P})$ in communication cost. Plimpton has also
2335 > found that force decompositions scale more favorably than spatial
2336 > decompositions for systems up to 10,000 atoms and favorably compete
2337 > with spatial methods up to 100,000 atoms.\cite{plimpton95}
2338 >
2339   \section{\label{oopseSec:conclusion}Conclusion}
2340  
2341 < We have presented the design and implementation of our open source
2342 < simulation package {\sc oopse}. The package offers novel capabilities
2343 < to the field of Molecular Dynamics simulation packages in the form of
2344 < dipolar force fields, and symplectic integration of rigid body
2345 < dynamics. It is capable of scaling across multiple processors through
2346 < the use of force based decomposition using MPI. It also implements
2347 < several advanced integrators allowing the end user control over
2348 < temperature and pressure. In addition, it is capable of integrating
2349 < constrained dynamics through both the {\sc rattle} algorithm and the
2350 < z-constraint method.
2341 > We have presented a new parallel simulation program called {\sc
2342 > oopse}. This program offers some novel capabilities, but mostly makes
2343 > available a library of modern object-oriented code for the scientific
2344 > community to use freely.  Notably, {\sc oopse} can handle symplectic
2345 > integration of objects (atoms and rigid bodies) which have
2346 > orientational degrees of freedom.  It can also work with transition
2347 > metal force fields and point-dipoles. It is capable of scaling across
2348 > multiple processors through the use of force based decomposition. It
2349 > also implements several advanced integrators allowing the end user
2350 > control over temperature and pressure. In addition, it is capable of
2351 > integrating constrained dynamics through both the {\sc rattle}
2352 > algorithm and the $z$-constraint method.
2353  
2354 < These features are all brought together in a single open-source
2355 < program. This allows researchers to not only benefit from
2356 < {\sc oopse}, but also contribute to {\sc oopse}'s development as
2357 < well.
2354 > We encourage other researchers to download and apply this program to
2355 > their own research problems.  By making the code available, we hope to
2356 > encourage other researchers to contribute their own code and make it a
2357 > more powerful package for everyone in the molecular dynamics community
2358 > to use.  All source code for {\sc oopse} is available for download at
2359 > {\tt http://oopse.org}.
2360  
1972
2361   \newpage
2362   \section{Acknowledgments}
1975 The authors would like to thank the Notre Dame BoB computer cluster where much of this project was tested. Additionally, the authors would like to acknowledge their funding from {\LARGE FIX ME}.
2363  
2364 < \bibliographystyle{achemso}
2364 > Development of {\sc oopse} was funded by a New Faculty Award from the
2365 > Camille and Henry Dreyfus Foundation and by the National Science
2366 > Foundation under grant CHE-0134881. Computation time was provided by
2367 > the Notre Dame Bunch-of-Boxes (B.o.B) computer cluster under NSF grant
2368 > DMR-0079647.
2369 >
2370 > \bibliographystyle{jcc}
2371   \bibliography{oopsePaper}
2372  
2373   \end{document}

Diff Legend

Removed lines
+ Added lines
< Changed lines
> Changed lines