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 1123 by mmeineke, Mon Apr 19 21:25:23 2004 UTC vs.
Revision 1168 by mmeineke, Wed May 12 19:32:47 2004 UTC

# Line 2 | Line 2
2   \usepackage{amsmath}
3   \usepackage{amssymb}
4   \usepackage{endfloat}
5 %\usepackage{berkeley}
5   \usepackage{listings}
6 + \usepackage{palatino}
7   \usepackage{graphicx}
8   \usepackage[ref]{overcite}
9   \usepackage{setspace}
# Line 25 | Line 25 | Engine for Molecular Dynamics}
25   \title{{\sc oopse}: An Open Source Object-Oriented Parallel Simulation
26   Engine for Molecular Dynamics}
27  
28 < \author{Matthew A. Meineke, Charles F. Vardeman II, Teng Lin, Christopher J. Fennell and J. Daniel Gezelter\\
28 > \author{Matthew A. Meineke, Charles F. Vardeman II, Teng Lin,\\
29 > Christopher J. Fennell and J. Daniel Gezelter\\
30   Department of Chemistry and Biochemistry\\
31   University of Notre Dame\\
32   Notre Dame, Indiana 46556}
# Line 35 | Line 36 | package ({\sc oopse}) that can perform molecular dynam
36  
37   \begin{abstract}
38   We detail the capabilities of a new open-source parallel simulation
39 < package ({\sc oopse}) that can perform molecular dynamics simulations
40 < on atom types that are missing from other popular packages.  In
41 < particular, {\sc oopse} is capable of performing orientational
41 < dynamics on dipolar systems, and it can handle simulations of metallic
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}).
43   \end{abstract}
44  
# Line 77 | Line 77 | We have structured this chapter to first discuss the e
77   those of the community of scientists who contribute back to the
78   project.
79  
80 < We have structured this chapter to first discuss the empirical energy
80 > We have structured this paper to first discuss the empirical energy
81   functions that {\sc oopse } implements in
82   Sec.~\ref{oopseSec:empiricalEnergy}. Following that is a discussion of
83   the various input and output files associated with the package
84   (Sec.~\ref{oopseSec:IOfiles}). Sec.~\ref{oopseSec:mechanics}
85   elucidates the various Molecular Dynamics algorithms {\sc oopse}
86   implements in the integration of the Newtonian equations of
87 < motion. Basic analysis of the trajectories obtained from the
88 < simulation is discussed in Sec.~\ref{oopseSec:props}. Program design
87 > motion.  Program design
88   considerations are presented in Sec.~\ref{oopseSec:design}. And
89   lastly, Sec.~\ref{oopseSec:conclusion} concludes the chapter.
90  
91 < \section{\label{oopseSec:empiricalEnergy}The Empirical Energy Functions}
91 > \section{\label{oopseSec:IOfiles}Concepts \& Files}
92  
93 < \subsection{\label{oopseSec:atomsMolecules}Atoms, Molecules and Rigid Bodies}
93 > \subsection{{\sc bass} and Model Files}
94  
95 < The basic unit of an {\sc oopse} simulation is the atom. The
96 < parameters describing the atom are generalized to make the atom as
97 < flexible a representation as possible. They may represent specific
98 < atoms of an element, or be used for collections of atoms such as
99 < methyl and carbonyl groups. The atoms are also capable of having
100 < directional components associated with them (\emph{e.g.}~permanent
101 < dipoles). Charges, permanent dipoles, and Lennard-Jones parameters for
102 < a given atom type are set in the force field parameter files.
95 > Every {\sc oopse} simulation begins with a Bizarre Atom Simulation
96 > Syntax ({\sc bass}) file. {\sc bass} is a script syntax that is parsed
97 > by {\sc oopse} at runtime. The {\sc bass} file allows for the user to
98 > 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}.
103  
104 < \begin{lstlisting}[float,caption={[Specifier for molecules and atoms] A sample specification of an Ar molecule},label=sch:AtmMole]
104 > \begin{lstlisting}[float,caption={[An example of a complete {\sc bass} file] An example showing a complete {\sc bass} file.},label={sch:bassExample}]
105 >
106   molecule{
107    name = "Ar";
108    nAtoms = 1;
# Line 111 | Line 111 | molecule{
111      position( 0.0, 0.0, 0.0 );
112    }
113   }
114 +
115 + nComponents = 1;
116 + component{
117 +  type = "Ar";
118 +  nMol = 108;
119 + }
120 +
121 + initialConfig = "./argon.init";
122 +
123 + forceField = "LJ";
124 + ensemble = "NVE"; // specify the simulation ensemble
125 + dt = 1.0;         // the time step for integration
126 + runTime = 1e3;    // the total simulation run time
127 + sampleTime = 100; // trajectory file frequency
128 + statusTime = 50;  // statistics file frequency
129 +
130   \end{lstlisting}
131  
132 + Within the \texttt{.bass} file it is necessary to provide a complete
133 + description of the molecule before it is actually placed in the
134 + simulation. The {\sc bass} syntax was originally developed with this
135 + goal in mind, and allows for the specification of all the atoms in a
136 + molecular prototype, as well as any bonds, bends, or torsions. These
137 + descriptions can become lengthy for complex molecules, and it would be
138 + inconvenient to duplicate the simulation at the beginning of each {\sc
139 + bass} script. Addressing this issue {\sc bass} allows for the
140 + inclusion of model files at the top of a \texttt{.bass} file. These
141 + model files, denoted with the \texttt{.mdl} extension, allow the user
142 + 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}.
147  
148 + \begin{lstlisting}[float,caption={An example \texttt{.mdl} file.},label={sch:mdlExample}]
149 +
150 + molecule{
151 +  name = "Ar";
152 +  nAtoms = 1;
153 +  atom[0]{
154 +    type="Ar";
155 +    position( 0.0, 0.0, 0.0 );
156 +  }
157 + }
158 +
159 + \end{lstlisting}
160 +
161 + \begin{lstlisting}[float,caption={Revised {\sc bass} example.},label={sch:bassExPrime}]
162 +
163 + #include "argon.mdl"
164 +
165 + nComponents = 1;
166 + component{
167 +  type = "Ar";
168 +  nMol = 108;
169 + }
170 +
171 + initialConfig = "./argon.init";
172 +
173 + forceField = "LJ";
174 + ensemble = "NVE";
175 + dt = 1.0;
176 + runTime = 1e3;
177 + sampleTime = 100;
178 + statusTime = 50;
179 +
180 + \end{lstlisting}
181 +
182 + \subsection{\label{oopseSec:atomsMolecules}Atoms, Molecules and Rigid Bodies}
183 +
184 + The basic unit of an {\sc oopse} simulation is the atom. The
185 + parameters describing the atom are generalized to make the atom as
186 + flexible a representation as possible. They may represent specific
187 + atoms of an element, or be used for collections of atoms such as
188 + methyl and carbonyl groups. The atoms are also capable of having
189 + directional components associated with them (\emph{e.g.}~permanent
190 + dipoles). Charges, permanent dipoles, and Lennard-Jones parameters for
191 + a given atom type are set in the force field parameter files.
192 +
193   Atoms can be collected into secondary structures such as rigid bodies
194   or molecules. The molecule is a way for {\sc oopse} to keep track of
195   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:AtmMole} shows how one creates a molecule in a ``model'' or
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.
# Line 133 | Line 209 | motion have been much worse than those available for t
209   included in most simulation packages because of the algorithmic
210   complexity involved in propagating orientational degrees of
211   freedom. Until recently, integrators which propagate orientational
212 < motion have been much worse than those available for translational
212 > motion have had energy conservation problems when compared to  those available for translational
213   motion.
214  
215   Moving a rigid body involves determination of both the force and
# Line 175 | Line 251 | definition of a rigid body can be seen in Scheme
251   entire nine parameter rotation matrix. Further discussion
252   on this choice can be found in Sec.~\ref{oopseSec:integrate}. An example
253   definition of a rigid body can be seen in Scheme
254 < \ref{sch:rigidBody}. The positions in the atom definitions are the
179 < placements of the atoms relative to the origin of the rigid body,
180 < which itself has a position relative to the origin of the molecule.
254 > \ref{sch:rigidBody}.
255  
256 < \begin{lstlisting}[float,caption={[Defining rigid bodies]A sample definition of a rigid body},label={sch:rigidBody}]
256 > \begin{lstlisting}[float,caption={[Defining rigid bodies]A sample definition of a molecule containing a rigid body},label={sch:rigidBody}]
257   molecule{
258    name = "TIP3P";
259    nAtoms = 3;
# Line 204 | Line 278 | molecule{
278   }
279   \end{lstlisting}
280  
281 + \subsection{\label{sec:miscConcepts}Putting a Script Together}
282 +
283 + 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.
284 +
285 + 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.
286 +
287 + 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}.
288 +
289 + \begin{table}
290 + \caption{The Global Keywords: Required Parameters}
291 + \label{table:reqParams}
292 + \begin{center}
293 + % Note when adding or removing columns, the \hsize numbers must add up to the total number
294 + % of columns.
295 + \begin{tabularx}{\linewidth}%
296 +  {>{\setlength{\hsize}{1.00\hsize}}X%
297 +  >{\setlength{\hsize}{0.4\hsize}}X%
298 +  >{\setlength{\hsize}{1.2\hsize}}X%
299 +  >{\setlength{\hsize}{1.4\hsize}}X}
300 +
301 + {\bf keyword} & {\bf units} & {\bf use} & {\bf remarks} \\ \hline
302 +
303 + {\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. \\
306 + {\tt nComponents} & integer & Sets the number of components. & Needs to appear before the first {\tt Component} block. \\
307 + {\tt initialConfig} & string & Sets the file containing the initial configuration. & Can point to any file containing the configuration in the correct order. \\
308 + {\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}. \\
309 +
310 +
311 + \end{tabularx}
312 + \end{center}
313 + \end{table}
314 +
315 + \begin{table}
316 + \caption{The Global Keywords: General Parameters}
317 + \label{table:genParams}
318 + \begin{center}
319 + % Note when adding or removing columns, the \hsize numbers must add up to the total number
320 + % of columns.
321 + \begin{tabularx}{\linewidth}%
322 +  {>{\setlength{\hsize}{1.00\hsize}}X%
323 +  >{\setlength{\hsize}{0.4\hsize}}X%
324 +  >{\setlength{\hsize}{1.2\hsize}}X%
325 +  >{\setlength{\hsize}{1.4\hsize}}X}
326 +
327 + {\bf keyword} & {\bf units} & {\bf use} & {\bf remarks} \\ \hline
328 +
329 + {\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. \\
330 + {\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. \\
331 + {\tt sampleTime} & fs & Sets the frequency at which the {\tt .dump} file is written. & Default sets the frequency to the {\tt runTime}. \\
332 + {\tt statusTime} & fs & Sets the frequency at which the {\tt .stat} file is written. & Defaults sets the frequency to the {\tt sampleTime}. \\
333 + {\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. \\
334 + {\tt electrostaticCutoffRadius}& & & \\
335 +      & $\mbox{\AA}$ & Manually sets the cutoff used by the electrostatic potentials. & Defaults to $15\mbox{\AA}$ \\
336 + {\tt electrostaticSkinThickness} & & & \\
337 +     & $\mbox{\AA}$  & Manually sets the skin thickness for the electrostatic switching function. & Defaults to 5~\% of the {\tt electrostaticSkinThickness}. \\
338 + {\tt useReactionField} & logical & Turns the reaction field correction on/off. & Default is "false". \\
339 + {\tt dielectric} & unitless & Sets the dielectric constant for reaction field. & If {\tt useReactionField} is true, then {\tt dielectric} must be set. \\
340 + {\tt usePeriodicBoundaryConditions} & & & \\
341 +        & logical & Turns periodic boundary conditions on/off. & Default is "true". \\
342 + {\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.
343 +
344 + \end{tabularx}
345 + \end{center}
346 + \end{table}
347 +
348 +
349 +
350 + \subsection{\label{oopseSec:coordFiles}Coordinate Files}
351 +
352 + The standard format for storage of a systems coordinates is a modified
353 + xyz-file syntax, the exact details of which can be seen in
354 + Scheme~\ref{sch:dumpFormat}. As all bonding and molecular information
355 + is stored in the \texttt{.bass} and \texttt{.mdl} files, the
356 + coordinate files are simply the complete set of coordinates for each
357 + atom at a given simulation time. One important note, although the
358 + simulation propagates the complete rotation matrix, directional
359 + entities are written out using quanternions, to save space in the
360 + output files.
361 +
362 + \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]
363 +
364 + nAtoms
365 + time; Hxx Hyx Hzx; Hxy Hyy Hzy; Hxz Hyz Hzz;
366 + Name1 x y z vx vy vz q0 q1 q2 q3 jx jy jz
367 + Name2 x y z vx vy vz q0 q1 q2 q3 jx jy jz
368 + etc...
369 +
370 + \end{lstlisting}
371 +
372 +
373 + There are three major files used by {\sc oopse} written in the
374 + coordinate format, they are as follows: the initialization file
375 + (\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.
392 +
393 + \subsection{\label{oopseSec:initCoords}Generation of Initial Coordinates}
394 +
395 + As was stated in Sec.~\ref{oopseSec:coordFiles}, an initialization
396 + file is needed to provide the starting coordinates for a
397 + 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.
398 +
399 + 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:
400 + \begin{enumerate}
401 + \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.
402 + \item The ordering of the atoms for each molecule follows the order declared in the molecule's declaration within the model file.
403 + \end{enumerate}
404 + An example is given in Scheme~\ref{sch:initEx1} resulting in the {\tt .init} file shown in Scheme~\ref{sch:initEx2}.
405 +
406 + \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]
407 +
408 + molecule{
409 +  name = "I2";
410 +  nAtoms = 2;
411 +  atom[0]{
412 +    type = "I";
413 +  }
414 +  atom[1]{
415 +    type = "I";
416 +  }
417 +  nBonds = 1;
418 +  bond[0]{
419 +    members( 0, 1);
420 +  }
421 + }
422 +
423 + molecule{
424 +  name = "HCl"
425 +  nAtoms = 2;
426 +  atom[0]{
427 +    type = "H";
428 +  }
429 +  atom[1]{
430 +    type = "Cl";
431 +  }
432 +  nBonds = 1;
433 +  bond[0]{
434 +    members( 0, 1);
435 +  }
436 + }
437 +
438 + nComponents = 2;
439 + component{
440 +  type = "HCl";
441 +  nMol = 4;
442 + }
443 + component{
444 +  type = "I2";
445 +  nMol = 1;
446 + }
447 +
448 + initialConfig = "mixture.init";
449 +
450 + \end{lstlisting}
451 +
452 + \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]
453 +
454 + 10
455 + 0.0;  10.0  0.0  0.0;  0.0  10.0  0.0;  0.0  0.0  10.0;
456 + H  ...
457 + Cl ...
458 + H  ...
459 + Cl ...
460 + H  ...
461 + Cl ...
462 + H  ...
463 + Cl ...
464 + I  ...
465 + I  ...
466 +
467 + \end{lstlisting}
468 +
469 +
470 + \subsection{The Statistics File}
471 +
472 + The last output file generated by {\sc oopse} is the statistics
473 + file. This file records such statistical quantities as the
474 + instantaneous temperature, volume, pressure, etc. It is written out
475 + with the frequency specified in the \texttt{.bass} file with the
476 + \texttt{statusTime} keyword. The file allows the user to observe the
477 + system variables as a function of simulation time while the simulation
478 + is in progress. One useful function the statistics file serves is to
479 + monitor the conserved quantity of a given simulation ensemble, this
480 + allows the user to observe the stability of the integrator. The
481 + statistics file is denoted with the \texttt{.stat} file extension.
482 +
483 +
484 + \section{\label{oopseSec:empiricalEnergy}The Empirical Energy Functions}
485 +
486 + \
487   \subsection{\label{sec:LJPot}The Lennard Jones Force Field}
488  
489   The most basic force field implemented in {\sc oopse} is the
# Line 281 | Line 561 | range interactions from $\frac{1}{r}$ to $\frac{1}{r^3
561   charges. Charge-neutral distributions were replaced with dipoles,
562   while most atoms and groups of atoms were reduced to Lennard-Jones
563   interaction sites. This simplification cuts the length scale of long
564 < range interactions from $\frac{1}{r}$ to $\frac{1}{r^3}$, and allows
285 < us to avoid the computationally expensive Ewald sum. Instead, we can
286 < use neighbor-lists and cutoff radii for the dipolar interactions, or
287 < include a reaction field to mimic larger range interactions.
564 > 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.
565  
566   As an example, lipid head-groups in {\sc duff} are represented as
567   point dipole interaction sites. By placing a dipole at the head
# Line 294 | Line 571 | reparameterization of the soft sticky dipole (SSD) mod
571   site is located at the pseudoatom's center of mass. The model is
572   illustrated by the red atom in Fig.~\ref{oopseFig:lipidModel}. The
573   water model we use to complement the dipoles of the lipids is our
574 < reparameterization of the soft sticky dipole (SSD) model of Ichiye
574 > reparameterization\cite{fennell04} of the soft sticky dipole (SSD) model of Ichiye
575   \emph{et al.}\cite{liu96:new_model}
576  
577   \begin{figure}
# Line 317 | Line 594 | TraPPE also constrains all bonds to be of fixed length
594   $\text{CH}_2$ do not change depending on what species are bonded to
595   it.
596  
597 < TraPPE also constrains all bonds to be of fixed length. Typically,
597 > TraPPE and {\sc duff} also constrain all bonds to be of fixed length. Typically,
598   bond vibrations are the fastest motions in a molecular dynamic
599   simulation. Small time steps between force evaluations must be used to
600   ensure adequate energy conservation in the bond degrees of freedom. By
# Line 347 | Line 624 | forceField = "DUFF";
624  
625   \end{lstlisting}
626  
627 < \subsection{\label{oopseSec:energyFunctions}{\sc duff} Energy Functions}
627 > \subsubsection{\label{oopseSec:energyFunctions}{\sc duff} Energy Functions}
628  
629   The total potential energy function in {\sc duff} is
630   \begin{equation}
# Line 476 | Line 753 | cutoff. The switching thickness can be set in the \tex
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.
755  
756 < \subsection{\label{oopseSec:SSD}The {\sc duff} Water Models: SSD/E and SSD/RF}
756 > \subsubsection{\label{oopseSec:SSD}The {\sc duff} Water Models: SSD/E and SSD/RF}
757  
758   In the interest of computational efficiency, the default solvent used
759   by {\sc oopse} is the extended Soft Sticky Dipole (SSD/E) water
# Line 594 | Line 871 | There are Molecular Dynamics packages which have the
871  
872   \subsection{\label{oopseSec:eam}Embedded Atom Method}
873  
874 < There are Molecular Dynamics packages which have the
875 < capacity to simulate metallic systems, including some that have
876 < parallel computational abilities\cite{plimpton93}. Potentials that
600 < describe bonding transition metal
601 < systems\cite{Finnis84,Ercolessi88,Chen90,Qi99,Ercolessi02} have an
602 < attractive interaction which models  ``Embedding''
874 > {\sc oopse} implements a potential that
875 > describes bonding transition metal
876 > systems\cite{Finnis84,Ercolessi88,Chen90,Qi99,Ercolessi02} and has attractive interaction which models  ``Embedding''
877   a positively charged metal ion in the electron density due to the
878   free valance ``sea'' of electrons created by the surrounding atoms in
879   the system. A mostly-repulsive pairwise part of the potential
# Line 693 | Line 967 | the inter-atomic forces.
967   the inter-atomic forces.
968  
969  
696 \section{\label{oopseSec:IOfiles}Input and Output Files}
970  
698 \subsection{{\sc bass} and Model Files}
699
700 Every {\sc oopse} simulation begins with a Bizarre Atom Simulation
701 Syntax ({\sc bass}) file. {\sc bass} is a script syntax that is parsed
702 by {\sc oopse} at runtime. The {\sc bass} file allows for the user to
703 completely describe the system they wish to simulate, as well as tailor
704 {\sc oopse}'s behavior during the simulation. {\sc bass} files are
705 denoted with the extension
706 \texttt{.bass}, an example file is shown in
707 Scheme~\ref{sch:bassExample}.
708
709 \begin{lstlisting}[float,caption={[An example of a complete {\sc bass} file] An example showing a complete {\sc bass} file.},label={sch:bassExample}]
710
711 molecule{
712  name = "Ar";
713  nAtoms = 1;
714  atom[0]{
715    type="Ar";
716    position( 0.0, 0.0, 0.0 );
717  }
718 }
719
720 nComponents = 1;
721 component{
722  type = "Ar";
723  nMol = 108;
724 }
725
726 initialConfig = "./argon.init";
727
728 forceField = "LJ";
729 ensemble = "NVE"; // specify the simulation ensemble
730 dt = 1.0;         // the time step for integration
731 runTime = 1e3;    // the total simulation run time
732 sampleTime = 100; // trajectory file frequency
733 statusTime = 50;  // statistics file frequency
734
735 \end{lstlisting}
736
737 Within the \texttt{.bass} file it is necessary to provide a complete
738 description of the molecule before it is actually placed in the
739 simulation. The {\sc bass} syntax was originally developed with this
740 goal in mind, and allows for the specification of all the atoms in a
741 molecular prototype, as well as any bonds, bends, or torsions. These
742 descriptions can become lengthy for complex molecules, and it would be
743 inconvenient to duplicate the simulation at the beginning of each {\sc
744 bass} script. Addressing this issue {\sc bass} allows for the
745 inclusion of model files at the top of a \texttt{.bass} file. These
746 model files, denoted with the \texttt{.mdl} extension, allow the user
747 to describe a molecular prototype once, then simply include it into
748 each simulation containing that molecule. Returning to the example in
749 Scheme~\ref{sch:bassExample}, the \texttt{.mdl} file's contents would
750 be Scheme~\ref{sch:mdlExample}, and the new \texttt{.bass} file would
751 become Scheme~\ref{sch:bassExPrime}.
752
753 \begin{lstlisting}[float,caption={An example \texttt{.mdl} file.},label={sch:mdlExample}]
754
755 molecule{
756  name = "Ar";
757  nAtoms = 1;
758  atom[0]{
759    type="Ar";
760    position( 0.0, 0.0, 0.0 );
761  }
762 }
763
764 \end{lstlisting}
765
766 \begin{lstlisting}[float,caption={Revised {\sc bass} example.},label={sch:bassExPrime}]
767
768 #include "argon.mdl"
769
770 nComponents = 1;
771 component{
772  type = "Ar";
773  nMol = 108;
774 }
775
776 initialConfig = "./argon.init";
777
778 forceField = "LJ";
779 ensemble = "NVE";
780 dt = 1.0;
781 runTime = 1e3;
782 sampleTime = 100;
783 statusTime = 50;
784
785 \end{lstlisting}
786
787 \subsection{\label{oopseSec:coordFiles}Coordinate Files}
788
789 The standard format for storage of a systems coordinates is a modified
790 xyz-file syntax, the exact details of which can be seen in
791 Scheme~\ref{sch:dumpFormat}. As all bonding and molecular information
792 is stored in the \texttt{.bass} and \texttt{.mdl} files, the
793 coordinate files are simply the complete set of coordinates for each
794 atom at a given simulation time. One important note, although the
795 simulation propagates the complete rotation matrix, directional
796 entities are written out using quanternions, to save space in the
797 output files.
798
799 \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 angular velocities.},label=sch:dumpFormat]
800
801 nAtoms
802 time; Hxx Hyx Hzx; Hxy Hyy Hzy; Hxz Hyz Hzz;
803 Name1 x y z vx vy vz q0 q1 q2 q3 jx jy jz
804 Name2 x y z vx vy vz q0 q1 q2 q3 jx jy jz
805 etc...
806
807 \end{lstlisting}
808
809
810 There are three major files used by {\sc oopse} written in the
811 coordinate format, they are as follows: the initialization file
812 (\texttt{.init}), the simulation trajectory file (\texttt{.dump}), and
813 the final coordinates of the simulation. The initialization file is
814 necessary for {\sc oopse} to start the simulation with the proper
815 coordinates, and is generated before the simulation run. The
816 trajectory file is created at the beginning of the simulation, and is
817 used to store snapshots of the simulation at regular intervals. The
818 first frame is a duplication of the
819 \texttt{.init} file, and each subsequent frame is appended to the file
820 at an interval specified in the \texttt{.bass} file with the
821 \texttt{sampleTime} flag. The final coordinate file is the end of run file. The
822 \texttt{.eor} file stores the final configuration of the system for a
823 given simulation. The file is updated at the same time as the
824 \texttt{.dump} file, however, it only contains the most recent
825 frame. In this way, an \texttt{.eor} file may be used as the
826 initialization file to a second simulation in order to continue a
827 simulation or recover one from a processor that has crashed during the
828 course of the run.
829
830 \subsection{\label{oopseSec:initCoords}Generation of Initial Coordinates}
831
832 As was stated in Sec.~\ref{oopseSec:coordFiles}, an initialization
833 file is needed to provide the starting coordinates for a
834 simulation. The {\sc oopse} package provides several system building
835 programs to aid in the creation of the \texttt{.init}
836 file. The programs use {\sc bass}, and will recognize
837 arguments and parameters in the \texttt{.bass} file that would
838 otherwise be ignored by the simulation.
839
840 \subsection{The Statistics File}
841
842 The last output file generated by {\sc oopse} is the statistics
843 file. This file records such statistical quantities as the
844 instantaneous temperature, volume, pressure, etc. It is written out
845 with the frequency specified in the \texttt{.bass} file with the
846 \texttt{statusTime} keyword. The file allows the user to observe the
847 system variables as a function of simulation time while the simulation
848 is in progress. One useful function the statistics file serves is to
849 monitor the conserved quantity of a given simulation ensemble, this
850 allows the user to observe the stability of the integrator. The
851 statistics file is denoted with the \texttt{.stat} file extension.
852
971   \section{\label{oopseSec:mechanics}Mechanics}
972  
973   \subsection{\label{oopseSec:integrate}Integrating the Equations of Motion: the
# Line 1637 | Line 1755 | simulations).
1755   orientational anisotropy in the system (i.e. in lipid bilayer
1756   simulations).
1757  
1758 < \subsection{\label{oopseSec:rattle}The {\sc rattle} Method for Bond
1758 > \subsection{\label{sec:constraints}Constraint Methods}
1759 >
1760 > \subsubsection{\label{oopseSec:rattle}The {\sc rattle} Method for Bond
1761          Constraints}
1762  
1763   In order to satisfy the constraints of fixed bond lengths within {\sc
1764   oopse}, we have implemented the {\sc rattle} algorithm of
1765   Andersen.\cite{andersen83} The algorithm is a velocity verlet
1766   formulation of the {\sc shake} method\cite{ryckaert77} of iteratively
1767 < solving the Lagrange multipliers of constraint. The system of Lagrange
1648 < multipliers allows one to reformulate the equations of motion with
1649 < explicit constraint forces.\cite{fowles99:lagrange}
1650 <
1651 < Consider a system described by coordinates $q_1$ and $q_2$ subject to an
1652 < equation of constraint:
1653 < \begin{equation}
1654 < \sigma(q_1, q_2,t) = 0
1655 < \label{oopseEq:lm1}
1656 < \end{equation}
1657 < The Lagrange formulation of the equations of motion can be written:
1658 < \begin{equation}
1659 < \delta\int_{t_1}^{t_2}L\, dt =
1660 <        \int_{t_1}^{t_2} \sum_i \biggl [ \frac{\partial L}{\partial q_i}
1661 <        - \frac{d}{dt}\biggl(\frac{\partial L}{\partial \dot{q}_i}
1662 <        \biggr ) \biggr] \delta q_i \, dt = 0.
1663 < \label{oopseEq:lm2}
1664 < \end{equation}
1665 < Here, $\delta q_i$ is not independent for each $q$, as $q_1$ and $q_2$
1666 < are linked by $\sigma$. However, $\sigma$ is fixed at any given
1667 < instant of time, giving:
1668 < \begin{align}
1669 < \delta\sigma &= \biggl( \frac{\partial\sigma}{\partial q_1} \delta q_1 %
1670 <        + \frac{\partial\sigma}{\partial q_2} \delta q_2 \biggr) = 0 ,\\
1671 < %
1672 < \frac{\partial\sigma}{\partial q_1} \delta q_1 &= %
1673 <        - \frac{\partial\sigma}{\partial q_2} \delta q_2, \\
1674 < %
1675 < \delta q_2 &= - \biggl(\frac{\partial\sigma}{\partial q_1} \bigg / %
1676 <        \frac{\partial\sigma}{\partial q_2} \biggr) \delta q_1.
1677 < \end{align}
1678 < Substituted back into Eq.~\ref{oopseEq:lm2},
1679 < \begin{equation}
1680 < \int_{t_1}^{t_2}\biggl [ \biggl(\frac{\partial L}{\partial q_1}
1681 <        - \frac{d}{dt}\,\frac{\partial L}{\partial \dot{q}_1}
1682 <        \biggr)
1683 <        - \biggl( \frac{\partial L}{\partial q_1}
1684 <        - \frac{d}{dt}\,\frac{\partial L}{\partial \dot{q}_1}
1685 <        \biggr) \biggl(\frac{\partial\sigma}{\partial q_1} \bigg / %
1686 <        \frac{\partial\sigma}{\partial q_2} \biggr)\biggr] \delta q_1 \, dt = 0.
1687 < \label{oopseEq:lm3}
1688 < \end{equation}
1689 < Leading to,
1690 < \begin{equation}
1691 < \frac{\biggl(\frac{\partial L}{\partial q_1}
1692 <        - \frac{d}{dt}\,\frac{\partial L}{\partial \dot{q}_1}
1693 <        \biggr)}{\frac{\partial\sigma}{\partial q_1}} =
1694 < \frac{\biggl(\frac{\partial L}{\partial q_2}
1695 <        - \frac{d}{dt}\,\frac{\partial L}{\partial \dot{q}_2}
1696 <        \biggr)}{\frac{\partial\sigma}{\partial q_2}}.
1697 < \label{oopseEq:lm4}
1698 < \end{equation}
1699 < This relation can only be statisfied, if both are equal to a single
1700 < function $-\lambda(t)$,
1701 < \begin{align}
1702 < \frac{\biggl(\frac{\partial L}{\partial q_1}
1703 <        - \frac{d}{dt}\,\frac{\partial L}{\partial \dot{q}_1}
1704 <        \biggr)}{\frac{\partial\sigma}{\partial q_1}} &= -\lambda(t), \\
1705 < %
1706 < \frac{\partial L}{\partial q_1}
1707 <        - \frac{d}{dt}\,\frac{\partial L}{\partial \dot{q}_1} &=
1708 <         -\lambda(t)\,\frac{\partial\sigma}{\partial q_1} ,\\
1709 < %
1710 < \frac{\partial L}{\partial q_1}
1711 <        - \frac{d}{dt}\,\frac{\partial L}{\partial \dot{q}_1}
1712 <         + \mathcal{G}_i &= 0,
1713 < \end{align}
1714 < where $\mathcal{G}_i$, the force of constraint on $i$, is:
1715 < \begin{equation}
1716 < \mathcal{G}_i = \lambda(t)\,\frac{\partial\sigma}{\partial q_1}.
1717 < \label{oopseEq:lm5}
1718 < \end{equation}
1767 > solving the Lagrange multipliers of constraint.
1768  
1769 < In a simulation, this would involve the solution of a set of $(m + n)$
1721 < number of equations. Where $m$ is the number of constraints, and $n$
1722 < is the number of constrained coordinates. In practice, this is not
1723 < done, as the matrix inversion necessary to solve the system of
1724 < equations would be very time consuming to solve. Additionally, the
1725 < numerical error in the solution of the set of $\lambda$'s would be
1726 < compounded by the error inherent in propagating by the Velocity Verlet
1727 < algorithm ($\Delta t^4$). The Verlet propagation error is negligible
1728 < in an unconstrained system, as one is interested in the statistics of
1729 < the run, and not that the run be numerically exact to the ``true''
1730 < integration. This relates back to the ergodic hypothesis that a time
1731 < integral of a valid trajectory will still give the correct ensemble
1732 < average. However, in the case of constraints, if the equations of
1733 < motion leave the ``true'' trajectory, they are departing from the
1734 < constrained surface. The method that is used, is to iteratively solve
1735 < for $\lambda(t)$ at each time step.
1736 <
1737 < In {\sc rattle} the equations of motion are modified subject to the
1738 < following two constraints:
1739 < \begin{align}
1740 < \sigma_{ij}[\mathbf{r}(t)] \equiv
1741 <        [ \mathbf{r}_i(t) - \mathbf{r}_j(t)]^2  - d_{ij}^2 &= 0 %
1742 <        \label{oopseEq:c1}, \\
1743 < %
1744 < [\mathbf{\dot{r}}_i(t) - \mathbf{\dot{r}}_j(t)] \cdot
1745 <        [\mathbf{r}_i(t) - \mathbf{r}_j(t)] &= 0 .\label{oopseEq:c2}
1746 < \end{align}
1747 < Eq.~\ref{oopseEq:c1} is the set of bond constraints, where $d_{ij}$ is
1748 < the constrained distance between atom $i$ and
1749 < $j$. Eq.~\ref{oopseEq:c2} constrains the velocities of $i$ and $j$ to
1750 < be perpendicular to the bond vector, so that the bond can neither grow
1751 < nor shrink. The constrained dynamics equations become:
1752 < \begin{equation}
1753 < m_i \mathbf{\ddot{r}}_i = \mathbf{F}_i + \mathbf{\mathcal{G}}_i,
1754 < \label{oopseEq:r1}
1755 < \end{equation}
1756 < where,$\mathbf{\mathcal{G}}_i$ are the forces of constraint on $i$,
1757 < and are defined:
1758 < \begin{equation}
1759 < \mathbf{\mathcal{G}}_i = - \sum_j \lambda_{ij}(t)\,\nabla \sigma_{ij}.
1760 < \label{oopseEq:r2}
1761 < \end{equation}
1762 <
1763 < In Velocity Verlet, if $\Delta t = h$, the propagation can be written:
1764 < \begin{align}
1765 < \mathbf{r}_i(t+h) &=
1766 <        \mathbf{r}_i(t) + h\mathbf{\dot{r}}(t) +
1767 <        \frac{h^2}{2m_i}\,\Bigl[ \mathbf{F}_i(t) +
1768 <        \mathbf{\mathcal{G}}_{Ri}(t) \Bigr] \label{oopseEq:vv1}, \\
1769 < %
1770 < \mathbf{\dot{r}}_i(t+h) &=
1771 <        \mathbf{\dot{r}}_i(t) + \frac{h}{2m_i}
1772 <        \Bigl[ \mathbf{F}_i(t) + \mathbf{\mathcal{G}}_{Ri}(t) +
1773 <        \mathbf{F}_i(t+h) + \mathbf{\mathcal{G}}_{Vi}(t+h) \Bigr] ,%
1774 <        \label{oopseEq:vv2}
1775 < \end{align}
1776 < where:
1777 < \begin{align}
1778 < \mathbf{\mathcal{G}}_{Ri}(t) &=
1779 <        -2 \sum_j \lambda_{Rij}(t) \mathbf{r}_{ij}(t) ,\\
1780 < %
1781 < \mathbf{\mathcal{G}}_{Vi}(t+h) &=
1782 <        -2 \sum_j \lambda_{Vij}(t+h) \mathbf{r}(t+h).
1783 < \end{align}
1784 < Next, define:
1785 < \begin{align}
1786 < g_{ij} &= h \lambda_{Rij}(t) ,\\
1787 < k_{ij} &= h \lambda_{Vij}(t+h), \\
1788 < \mathbf{q}_i &= \mathbf{\dot{r}}_i(t) + \frac{h}{2m_i} \mathbf{F}_i(t)
1789 <        - \frac{1}{m_i}\sum_j g_{ij}\mathbf{r}_{ij}(t).
1790 < \end{align}
1791 < Using these definitions, Eq.~\ref{oopseEq:vv1} and \ref{oopseEq:vv2}
1792 < can be rewritten as,
1793 < \begin{align}
1794 < \mathbf{r}_i(t+h) &= \mathbf{r}_i(t) + h \mathbf{q}_i ,\\
1795 < %
1796 < \mathbf{\dot{r}}(t+h) &= \mathbf{q}_i + \frac{h}{2m_i}\mathbf{F}_i(t+h)
1797 <        -\frac{1}{m_i}\sum_j k_{ij} \mathbf{r}_{ij}(t+h).
1798 < \end{align}
1799 <
1800 < To integrate the equations of motion, the {\sc rattle} algorithm first
1801 < solves for $\mathbf{r}(t+h)$. Let,
1802 < \begin{equation}
1803 < \mathbf{q}_i = \mathbf{\dot{r}}(t) + \frac{h}{2m_i}\mathbf{F}_i(t).
1804 < \end{equation}
1805 < Here $\mathbf{q}_i$ corresponds to an initial unconstrained move. Next
1806 < pick a constraint $j$, and let,
1807 < \begin{equation}
1808 < \mathbf{s} = \mathbf{r}_i(t) + h\mathbf{q}_i(t)
1809 <        - \mathbf{r}_j(t) + h\mathbf{q}_j(t).
1810 < \label{oopseEq:ra1}
1811 < \end{equation}
1812 < If
1813 < \begin{equation}
1814 < \Big| |\mathbf{s}|^2 - d_{ij}^2 \Big| > \text{tolerance},
1815 < \end{equation}
1816 < then the constraint is unsatisfied, and corrections are made to the
1817 < positions. First we define a test corrected configuration as,
1818 < \begin{align}
1819 < \mathbf{r}_i^T(t+h) = \mathbf{r}_i(t) + h\biggl[\mathbf{q}_i -
1820 <        g_{ij}\,\frac{\mathbf{r}_{ij}(t)}{m_i} \biggr] ,\\
1821 < %
1822 < \mathbf{r}_j^T(t+h) = \mathbf{r}_j(t) + h\biggl[\mathbf{q}_j +
1823 <        g_{ij}\,\frac{\mathbf{r}_{ij}(t)}{m_j} \biggr].
1824 < \end{align}
1825 < And we chose $g_{ij}$ such that, $|\mathbf{r}_i^T - \mathbf{r}_j^T|^2
1826 < = d_{ij}^2$. Solving the quadratic for $g_{ij}$ we obtain the
1827 < approximation,
1828 < \begin{equation}
1829 < g_{ij} = \frac{(s^2 - d^2)}{2h[\mathbf{s}\cdot\mathbf{r}_{ij}(t)]
1830 <        (\frac{1}{m_i} + \frac{1}{m_j})}.
1831 < \end{equation}
1832 < Although not an exact solution for $g_{ij}$, as this is an iterative
1833 < scheme overall, the eventual solution will converge. With a trial
1834 < $g_{ij}$, the new $\mathbf{q}$'s become,
1835 < \begin{align}
1836 < \mathbf{q}_i &= \mathbf{q}^{\text{old}}_i - g_{ij}\,
1837 <        \frac{\mathbf{r}_{ij}(t)}{m_i} ,\\
1838 < %
1839 < \mathbf{q}_j &= \mathbf{q}^{\text{old}}_j + g_{ij}\,
1840 <        \frac{\mathbf{r}_{ij}(t)}{m_j} .
1841 < \end{align}
1842 < The whole algorithm is then repeated from Eq.~\ref{oopseEq:ra1} until
1843 < all constraints are satisfied.
1769 > \subsubsection{\label{oopseSec:zcons}Z-Constraint Method}
1770  
1845 The second step of {\sc rattle}, is to then update the velocities. The
1846 step starts with,
1847 \begin{equation}
1848 \mathbf{\dot{r}}_i(t+h) = \mathbf{q}_i + \frac{h}{2m_i}\mathbf{F}_i(t+h).
1849 \end{equation}
1850 Next we pick a constraint $j$, and calculate the dot product $\ell$.
1851 \begin{equation}
1852 \ell = \mathbf{r}_{ij}(t+h) \cdot \mathbf{\dot{r}}_{ij}(t+h).
1853 \label{oopseEq:rv1}
1854 \end{equation}
1855 Here if constraint Eq.~\ref{oopseEq:c2} holds, $\ell$ should be
1856 zero. Therefore if $\ell$ is greater than some tolerance, then
1857 corrections are made to the $i$ and $j$ velocities.
1858 \begin{align}
1859 \mathbf{\dot{r}}_i^T &= \mathbf{\dot{r}}_i(t+h) - k_{ij}
1860        \frac{\mathbf{\dot{r}}_{ij}(t+h)}{m_i}, \\
1861 %
1862 \mathbf{\dot{r}}_j^T &= \mathbf{\dot{r}}_j(t+h) + k_{ij}
1863        \frac{\mathbf{\dot{r}}_{ij}(t+h)}{m_j}.
1864 \end{align}
1865 Like in the previous step, we select a value for $k_{ij}$ such that
1866 $\ell$ is zero.
1867 \begin{equation}
1868 k_{ij} = \frac{\ell}{d^2_{ij}(\frac{1}{m_i} + \frac{1}{m_j})}.
1869 \end{equation}
1870 The test velocities, $\mathbf{\dot{r}}^T_i$ and
1871 $\mathbf{\dot{r}}^T_j$, then replace their respective velocities, and
1872 the algorithm is iterated from Eq.~\ref{oopseEq:rv1} until all
1873 constraints are satisfied.
1874
1875
1876 \subsection{\label{oopseSec:zcons}Z-Constraint Method}
1877
1771   Based on the fluctuation-dissipation theorem, a force auto-correlation
1772   method was developed by Roux and Karplus to investigate the dynamics
1773   of ions inside ion channels.\cite{Roux91} The time-dependent friction
# Line 1959 | Line 1852 | F_{z_{\text{Harmonic}}}(t)=-\frac{\partial U(t)}{\part
1852   \begin{equation}
1853   F_{z_{\text{Harmonic}}}(t)=-\frac{\partial U(t)}{\partial z(t)}=
1854          -k_{\text{Harmonic}}(z(t)-z_{\text{cons}}).
1962 \end{equation}
1963
1964 \section{\label{oopseSec:props}Trajectory Analysis}
1965
1966 \subsection{\label{oopseSec:staticProps}Static Property Analysis}
1967
1968 The static properties of the trajectories are analyzed with the
1969 program \texttt{staticProps}. The code is capable of calculating a
1970 number of pair correlations between species A and B. Some of which
1971 only apply to directional entities. The summary of pair correlations
1972 can be found in Table~\ref{oopseTb:gofrs}
1973
1974 \begin{table}
1975 \caption{THE DIFFERENT PAIR CORRELATIONS IN \texttt{staticProps}}
1976 \label{oopseTb:gofrs}
1977 \begin{center}
1978 \begin{tabular}{|l|c|c|}
1979 \hline
1980 Name      & Equation & Directional Atom \\ \hline
1981 $g_{\text{AB}}(r)$              & Eq.~\ref{eq:gofr}         & neither \\ \hline
1982 $g_{\text{AB}}(r, \cos \theta)$ & Eq.~\ref{eq:gofrCosTheta} & A \\ \hline
1983 $g_{\text{AB}}(r, \cos \omega)$ & Eq.~\ref{eq:gofrCosOmega} & both \\ \hline
1984 $g_{\text{AB}}(x, y, z)$        & Eq.~\ref{eq:gofrXYZ}      & neither \\ \hline
1985 $\langle \cos \omega \rangle_{\text{AB}}(r)$ & Eq.~\ref{eq:cosOmegaOfR} &%
1986        both \\ \hline
1987 \end{tabular}
1988 \begin{minipage}{\linewidth}
1989 \centering
1990 \vspace{2mm}
1991 The third column specifies which atom, if any, need be a directional entity.
1992 \end{minipage}
1993 \end{center}
1994 \end{table}
1995
1996 The first pair correlation, $g_{\text{AB}}(r)$, is defined as follows:
1997 \begin{equation}
1998 g_{\text{AB}}(r) = \frac{V}{N_{\text{A}}N_{\text{B}}}\langle %%
1999        \sum_{i \in \text{A}} \sum_{j \in \text{B}} %%
2000        \delta( r - |\mathbf{r}_{ij}|) \rangle, \label{eq:gofr}
2001 \end{equation}
2002 where $\mathbf{r}_{ij}$ is the vector
2003 \begin{equation*}
2004 \mathbf{r}_{ij} = \mathbf{r}_j - \mathbf{r}_i, \notag
2005 \end{equation*}
2006 and $\frac{V}{N_{\text{A}}N_{\text{B}}}$ normalizes the average over
2007 the expected pair density at a given $r$.
2008
2009 The next two pair correlations, $g_{\text{AB}}(r, \cos \theta)$ and
2010 $g_{\text{AB}}(r, \cos \omega)$, are similar in that they are both two
2011 dimensional histograms. Both use $r$ for the primary axis then a
2012 $\cos$ for the secondary axis ($\cos \theta$ for
2013 Eq.~\ref{eq:gofrCosTheta} and $\cos \omega$ for
2014 Eq.~\ref{eq:gofrCosOmega}). This allows for the investigator to
2015 correlate alignment on directional entities. $g_{\text{AB}}(r, \cos
2016 \theta)$ is defined as follows:
2017 \begin{equation}
2018 g_{\text{AB}}(r, \cos \theta) = \frac{V}{N_{\text{A}}N_{\text{B}}}\langle  
2019 \sum_{i \in \text{A}} \sum_{j \in \text{B}}  
2020 \delta( \cos \theta - \cos \theta_{ij})
2021 \delta( r - |\mathbf{r}_{ij}|) \rangle.
2022 \label{eq:gofrCosTheta}
2023 \end{equation}
2024 Here
2025 \begin{equation*}
2026 \cos \theta_{ij} = \mathbf{\hat{i}} \cdot \mathbf{\hat{r}}_{ij},
2027 \end{equation*}
2028 where $\mathbf{\hat{i}}$ is the unit directional vector of species $i$
2029 and $\mathbf{\hat{r}}_{ij}$ is the unit vector associated with vector
2030 $\mathbf{r}_{ij}$.
2031
2032 The second two dimensional histogram is of the form:
2033 \begin{equation}
2034 g_{\text{AB}}(r, \cos \omega) =
2035        \frac{V}{N_{\text{A}}N_{\text{B}}}\langle
2036        \sum_{i \in \text{A}} \sum_{j \in \text{B}}
2037        \delta( \cos \omega - \cos \omega_{ij})
2038        \delta( r - |\mathbf{r}_{ij}|) \rangle. \label{eq:gofrCosOmega}
1855   \end{equation}
2040 Here
2041 \begin{equation*}
2042 \cos \omega_{ij} = \mathbf{\hat{i}} \cdot \mathbf{\hat{j}}.
2043 \end{equation*}
2044 Again, $\mathbf{\hat{i}}$ and $\mathbf{\hat{j}}$ are the unit
2045 directional vectors of species $i$ and $j$.
1856  
2047 The static analysis code is also cable of calculating a three
2048 dimensional pair correlation of the form:
2049 \begin{equation}\label{eq:gofrXYZ}
2050 g_{\text{AB}}(x, y, z) =
2051        \frac{V}{N_{\text{A}}N_{\text{B}}}\langle
2052        \sum_{i \in \text{A}} \sum_{j \in \text{B}}
2053        \delta( x - x_{ij})
2054        \delta( y - y_{ij})
2055        \delta( z - z_{ij}) \rangle,
2056 \end{equation}
2057 where $x_{ij}$, $y_{ij}$, and $z_{ij}$ are the $x$, $y$, and $z$
2058 components respectively of vector $\mathbf{r}_{ij}$.
2059
2060 The final pair correlation is similar to
2061 Eq.~\ref{eq:gofrCosOmega}. $\langle \cos \omega
2062 \rangle_{\text{AB}}(r)$ is calculated in the following way:
2063 \begin{equation}\label{eq:cosOmegaOfR}
2064 \langle \cos \omega \rangle_{\text{AB}}(r)  =
2065        \langle \sum_{i \in \text{A}} \sum_{j \in \text{B}}
2066        (\cos \omega_{ij}) \delta( r - |\mathbf{r}_{ij}|) \rangle.
2067 \end{equation}
2068 Here $\cos \omega_{ij}$ is defined in the same way as in
2069 Eq.~\ref{eq:gofrCosOmega}. This equation is a single dimensional pair
2070 correlation that gives the average correlation of two directional
2071 entities as a function of their distance from each other.
2072
2073 \subsection{\label{dynamicProps}Dynamic Property Analysis}
2074
2075 The dynamic properties of a trajectory are calculated with the program
2076 \texttt{dynamicProps}. The program calculates the following properties:
2077 \begin{gather}
2078 \langle | \mathbf{r}(t) - \mathbf{r}(0) |^2 \rangle, \label{eq:rms}\\
2079 \langle \mathbf{v}(t) \cdot \mathbf{v}(0) \rangle, \label{eq:velCorr} \\
2080 \langle \mathbf{j}(t) \cdot \mathbf{j}(0) \rangle. \label{eq:angularVelCorr}
2081 \end{gather}
2082
2083 Eq.~\ref{eq:rms} is the root mean square displacement function. Which
2084 allows one to observe the average displacement of an atom as a
2085 function of time. The quantity is useful when calculating diffusion
2086 coefficients because of the Einstein Relation, which is valid at long
2087 times.\cite{allen87:csl}
2088 \begin{equation}
2089 2tD = \langle | \mathbf{r}(t) - \mathbf{r}(0) |^2 \rangle.
2090 \label{oopseEq:einstein}
2091 \end{equation}
2092
2093 Eq.~\ref{eq:velCorr} and \ref{eq:angularVelCorr} are the translational
2094 velocity and angular velocity correlation functions respectively. The
2095 latter is only applicable to directional species in the
2096 simulation. The velocity autocorrelation functions are useful when
2097 determining vibrational information about the system of interest.
2098
1857   \section{\label{oopseSec:design}Program Design}
1858  
1859   \subsection{\label{sec:architecture} {\sc oopse} Architecture}
# Line 2193 | Line 1951 | atoms.\cite{plimpton95}
1951   and favorably compete with spatial methods up to 100,000
1952   atoms.\cite{plimpton95}
1953  
2196 \subsection{\label{oopseSec:memAlloc}Memory Issues in Trajectory Analysis}
2197
2198 For large simulations, the trajectory files can sometimes reach sizes
2199 in excess of several gigabytes. In order to effectively analyze that
2200 amount of data, two memory management schemes have been devised for
2201 \texttt{staticProps} and for \texttt{dynamicProps}. The first scheme,
2202 developed for \texttt{staticProps}, is the simplest. As each frame's
2203 statistics are calculated independent of each other, memory is
2204 allocated for each frame, then freed once correlation calculations are
2205 complete for the snapshot. To prevent multiple passes through a
2206 potentially large file, \texttt{staticProps} is capable of calculating
2207 all requested correlations per frame with only a single pair loop in
2208 each frame and a single read of the file.
2209
2210 The second, more advanced memory scheme, is used by
2211 \texttt{dynamicProps}. Here, the program must have multiple frames in
2212 memory to calculate time dependent correlations. In order to prevent a
2213 situation where the program runs out of memory due to large
2214 trajectories, the user is able to specify that the trajectory be read
2215 in blocks. The number of frames in each block is specified by the
2216 user, and upon reading a block of the trajectory,
2217 \texttt{dynamicProps} will calculate all of the time correlation frame
2218 pairs within the block. After in-block correlations are complete, a
2219 second block of the trajectory is read, and the cross correlations are
2220 calculated between the two blocks. This second block is then freed and
2221 then incremented and the process repeated until the end of the
2222 trajectory. Once the end is reached, the first block is freed then
2223 incremented, and the again the internal time correlations are
2224 calculated. The algorithm with the second block is then repeated with
2225 the new origin block, until all frame pairs have been correlated in
2226 time. This process is illustrated in
2227 Fig.~\ref{oopseFig:dynamicPropsMemory}.
2228
2229 %\begin{figure}
2230 %\centering
2231 %\includegraphics[width=\linewidth]{dynamicPropsMem.eps}
2232 %\caption[A representation of the block correlations in \texttt{dynamicProps}]{This diagram illustrates the memory management used by \texttt{dynamicProps}, which follows the scheme: $\sum^{N_{\text{memory blocks}}}_{i=1}[ \operatorname{self}(i) + \sum^{N_{\text{memory blocks}}}_{j>i} \operatorname{cross}(i,j)]$. The shaded region represents the self correlation of the memory block, and the open blocks are read one at a time and the cross correlations between blocks are calculated.}
2233 %\label{oopseFig:dynamicPropsMemory}
2234 %\end{figure}
2235
1954   \section{\label{oopseSec:conclusion}Conclusion}
1955  
1956   We have presented the design and implementation of our open source
# Line 2254 | Line 1972 | The authors would like to thank espresso for fueling t
1972  
1973   \newpage
1974   \section{Acknowledgments}
1975 < The authors would like to thank espresso for fueling this work, and
2258 < would also like to send a special acknowledgement to single malt
2259 < scotch for its wonderful calming effects and its ability to make the
2260 < troubles of the world float away.
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}.
1976  
1977   \bibliographystyle{achemso}
1978   \bibliography{oopsePaper}

Diff Legend

Removed lines
+ Added lines
< Changed lines
> Changed lines