ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/oopsePaper/oopsePaper.tex
Revision: 1155
Committed: Tue May 11 19:36:28 2004 UTC (20 years, 1 month ago) by mmeineke
Content type: application/x-tex
File size: 84129 byte(s)
Log Message:
starting to rearrange the prose after the intro.

File Contents

# Content
1 \documentclass[11pt]{article}
2 \usepackage{amsmath}
3 \usepackage{amssymb}
4 \usepackage{endfloat}
5 \usepackage{listings}
6 \usepackage{palatino}
7 \usepackage{graphicx}
8 \usepackage[ref]{overcite}
9 \usepackage{setspace}
10 \usepackage{tabularx}
11 \pagestyle{plain}
12 \pagenumbering{arabic}
13 \oddsidemargin 0.0cm \evensidemargin 0.0cm
14 \topmargin -21pt \headsep 10pt
15 \textheight 9.0in \textwidth 6.5in
16 \brokenpenalty=10000
17 \renewcommand{\baselinestretch}{1.2}
18 \renewcommand\citemid{\ } % no comma in optional reference note
19
20 \begin{document}
21 \lstset{language=C,frame=TB,basicstyle=\small,basicstyle=\ttfamily, %
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
26 Engine for Molecular Dynamics}
27
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}
33
34 \date{\today}
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}).
43 \end{abstract}
44
45 \section{\label{sec:intro}Introduction}
46
47 When choosing to simulate a chemical system with molecular dynamics,
48 there are a variety of options available. For simple systems, one
49 might consider writing one's own programming code. However, as systems
50 grow larger and more complex, building and maintaining code for the
51 simulations becomes a time consuming task. In such cases it is usually
52 more convenient for a researcher to turn to pre-existing simulation
53 packages. These packages, such as {\sc amber}\cite{pearlman:1995} and
54 {\sc charmm}\cite{Brooks83}, provide powerful tools for researchers to
55 conduct simulations of their systems without spending their time
56 developing a code base to conduct their research. This then frees them
57 to perhaps explore experimental analogues to their models.
58
59 Despite their utility, problems with these packages arise when
60 researchers try to develop techniques or energetic models that the
61 code was not originally designed to simulate. Examples of techniques
62 and energetics not commonly implemented include; dipole-dipole
63 interactions, rigid body dynamics, and metallic potentials. When faced
64 with these obstacles, a researcher must either develop their own code
65 or license and extend one of the commercial packages. What we have
66 elected to do is develop a body of simulation code capable of
67 implementing the types of models upon which our research is based.
68
69 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.
79
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. 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:IOfiles}Concepts \& Files}
92
93 \subsection{{\sc bass} and Model Files}
94
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={[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;
109 atom[0]{
110 type="Ar";
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 \begin{lstlisting}[float,caption={[Specifier for molecules and atoms] A sample specification of an Ar molecule},label=sch:AtmMole]
194 molecule{
195 name = "Ar";
196 nAtoms = 1;
197 atom[0]{
198 type="Ar";
199 position( 0.0, 0.0, 0.0 );
200 }
201 }
202 \end{lstlisting}
203
204
205 Atoms can be collected into secondary structures such as rigid bodies
206 or molecules. The molecule is a way for {\sc oopse} to keep track of
207 the atoms in a simulation in logical manner. Molecular units store the
208 identities of all the atoms and rigid bodies associated with
209 themselves, and are responsible for the evaluation of their own
210 internal interactions (\emph{i.e.}~bonds, bends, and torsions). Scheme
211 \ref{sch:AtmMole} shows how one creates a molecule in a ``model'' or
212 \texttt{.mdl} file. The position of the atoms given in the
213 declaration are relative to the origin of the molecule, and is used
214 when creating a system containing the molecule.
215
216 As stated previously, one of the features that sets {\sc oopse} apart
217 from most of the current molecular simulation packages is the ability
218 to handle rigid body dynamics. Rigid bodies are non-spherical
219 particles or collections of particles that have a constant internal
220 potential and move collectively.\cite{Goldstein01} They are not
221 included in most simulation packages because of the algorithmic
222 complexity involved in propagating orientational degrees of
223 freedom. Until recently, integrators which propagate orientational
224 motion have been much worse than those available for translational
225 motion.
226
227 Moving a rigid body involves determination of both the force and
228 torque applied by the surroundings, which directly affect the
229 translational and rotational motion in turn. In order to accumulate
230 the total force on a rigid body, the external forces and torques must
231 first be calculated for all the internal particles. The total force on
232 the rigid body is simply the sum of these external forces.
233 Accumulation of the total torque on the rigid body is more complex
234 than the force because the torque is applied to the center of mass of
235 the rigid body. The torque on rigid body $i$ is
236 \begin{equation}
237 \boldsymbol{\tau}_i=
238 \sum_{a}\biggl[(\mathbf{r}_{ia}-\mathbf{r}_i)\times \mathbf{f}_{ia}
239 + \boldsymbol{\tau}_{ia}\biggr],
240 \label{eq:torqueAccumulate}
241 \end{equation}
242 where $\boldsymbol{\tau}_i$ and $\mathbf{r}_i$ are the torque on and
243 position of the center of mass respectively, while $\mathbf{f}_{ia}$,
244 $\mathbf{r}_{ia}$, and $\boldsymbol{\tau}_{ia}$ are the force on,
245 position of, and torque on the component particles of the rigid body.
246
247 The summation of the total torque is done in the body fixed axis of
248 each rigid body. In order to move between the space fixed and body
249 fixed coordinate axes, parameters describing the orientation must be
250 maintained for each rigid body. At a minimum, the rotation matrix
251 ($\mathsf{A}$) can be described by the three Euler angles ($\phi,
252 \theta,$ and $\psi$), where the elements of $\mathsf{A}$ are composed of
253 trigonometric operations involving $\phi, \theta,$ and
254 $\psi$.\cite{Goldstein01} In order to avoid numerical instabilities
255 inherent in using the Euler angles, the four parameter ``quaternion''
256 scheme is often used. The elements of $\mathsf{A}$ can be expressed as
257 arithmetic operations involving the four quaternions ($q_0, q_1, q_2,$
258 and $q_3$).\cite{allen87:csl} Use of quaternions also leads to
259 performance enhancements, particularly for very small
260 systems.\cite{Evans77}
261
262 {\sc oopse} utilizes a relatively new scheme that propagates the
263 entire nine parameter rotation matrix. Further discussion
264 on this choice can be found in Sec.~\ref{oopseSec:integrate}. An example
265 definition of a rigid body can be seen in Scheme
266 \ref{sch:rigidBody}. The positions in the atom definitions are the
267 placements of the atoms relative to the origin of the rigid body,
268 which itself has a position relative to the origin of the molecule.
269
270 \begin{lstlisting}[float,caption={[Defining rigid bodies]A sample definition of a rigid body},label={sch:rigidBody}]
271 molecule{
272 name = "TIP3P";
273 nAtoms = 3;
274 atom[0]{
275 type = "O_TIP3P";
276 position( 0.0, 0.0, -0.06556 );
277 }
278 atom[1]{
279 type = "H_TIP3P";
280 position( 0.0, 0.75695, 0.52032 );
281 }
282 atom[2]{
283 type = "H_TIP3P";
284 position( 0.0, -0.75695, 0.52032 );
285 }
286
287 nRigidBodies = 1;
288 rigidBody[0]{
289 nMembers = 3;
290 members(0, 1, 2);
291 }
292 }
293 \end{lstlisting}
294
295 \subsection{\label{sec:miscConcepts}Putting a Script Together}
296
297 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.
298
299 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}).
300
301 \subsection{\label{oopseSec:coordFiles}Coordinate Files}
302
303 The standard format for storage of a systems coordinates is a modified
304 xyz-file syntax, the exact details of which can be seen in
305 Scheme~\ref{sch:dumpFormat}. As all bonding and molecular information
306 is stored in the \texttt{.bass} and \texttt{.mdl} files, the
307 coordinate files are simply the complete set of coordinates for each
308 atom at a given simulation time. One important note, although the
309 simulation propagates the complete rotation matrix, directional
310 entities are written out using quanternions, to save space in the
311 output files.
312
313 \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]
314
315 nAtoms
316 time; Hxx Hyx Hzx; Hxy Hyy Hzy; Hxz Hyz Hzz;
317 Name1 x y z vx vy vz q0 q1 q2 q3 jx jy jz
318 Name2 x y z vx vy vz q0 q1 q2 q3 jx jy jz
319 etc...
320
321 \end{lstlisting}
322
323
324 There are three major files used by {\sc oopse} written in the
325 coordinate format, they are as follows: the initialization file
326 (\texttt{.init}), the simulation trajectory file (\texttt{.dump}), and
327 the final coordinates of the simulation (\texttt{.eor}). The initialization file is
328 necessary for {\sc oopse} to start the simulation with the proper
329 coordinates, and is generated before the simulation run. The
330 trajectory file is created at the beginning of the simulation, and is
331 used to store snapshots of the simulation at regular intervals. The
332 first frame is a duplication of the
333 \texttt{.init} file, and each subsequent frame is appended to the file
334 at an interval specified in the \texttt{.bass} file with the
335 \texttt{sampleTime} flag. The final coordinate file is the end of run file. The
336 \texttt{.eor} file stores the final configuration of the system for a
337 given simulation. The file is updated at the same time as the
338 \texttt{.dump} file, however, it only contains the most recent
339 frame. In this way, an \texttt{.eor} file may be used as the
340 initialization file to a second simulation in order to continue a
341 simulation or recover one from a processor that has crashed during the
342 course of the run.
343
344 \subsection{\label{oopseSec:initCoords}Generation of Initial Coordinates}
345
346 As was stated in Sec.~\ref{oopseSec:coordFiles}, an initialization
347 file is needed to provide the starting coordinates for a
348 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.
349
350 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:
351 \begin{enumerate}
352 \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.
353 \item The ordering of the atoms for each molecule follows the order declared in the molecule's declaration within the model file.
354 \end{enumerate}
355 An example is given in Scheme~\ref{sch:initEx1} resulting in the {\tt .init} file shown in Scheme~\ref{sch:initEx2}.
356
357 \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]
358
359 molecule{
360 name = "I2";
361 nAtoms = 2;
362 atom[0]{
363 type = "I";
364 }
365 atom[1]{
366 type = "I";
367 }
368 nBonds = 1;
369 bond[0]{
370 members( 0, 1);
371 }
372 }
373
374 molecule{
375 name = "HCl"
376 nAtoms = 2;
377 atom[0]{
378 type = "H";
379 }
380 atom[1]{
381 type = "Cl";
382 }
383 nBonds = 1;
384 bond[0]{
385 members( 0, 1);
386 }
387 }
388
389 nComponents = 2;
390 component{
391 type = "HCl";
392 nMol = 4;
393 }
394 component{
395 type = "I2";
396 nMol = 1;
397 }
398
399 initialConfig = "mixture.init";
400
401 \end{lstlisting}
402
403 \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]
404
405 10
406 0.0; 10.0 0.0 0.0; 0.0 10.0 0.0; 0.0 0.0 10.0;
407 H ...
408 Cl ...
409 H ...
410 Cl ...
411 H ...
412 Cl ...
413 H ...
414 Cl ...
415 I ...
416 I ...
417
418 \end{lstlisting}
419
420
421 \subsection{The Statistics File}
422
423 The last output file generated by {\sc oopse} is the statistics
424 file. This file records such statistical quantities as the
425 instantaneous temperature, volume, pressure, etc. It is written out
426 with the frequency specified in the \texttt{.bass} file with the
427 \texttt{statusTime} keyword. The file allows the user to observe the
428 system variables as a function of simulation time while the simulation
429 is in progress. One useful function the statistics file serves is to
430 monitor the conserved quantity of a given simulation ensemble, this
431 allows the user to observe the stability of the integrator. The
432 statistics file is denoted with the \texttt{.stat} file extension.
433
434
435 \section{\label{oopseSec:empiricalEnergy}The Empirical Energy Functions}
436
437 \
438 \subsection{\label{sec:LJPot}The Lennard Jones Force Field}
439
440 The most basic force field implemented in {\sc oopse} is the
441 Lennard-Jones force field, which mimics the van der Waals interaction at
442 long distances, and uses an empirical repulsion at short
443 distances. The Lennard-Jones potential is given by:
444 \begin{equation}
445 V_{\text{LJ}}(r_{ij}) =
446 4\epsilon_{ij} \biggl[
447 \biggl(\frac{\sigma_{ij}}{r_{ij}}\biggr)^{12}
448 - \biggl(\frac{\sigma_{ij}}{r_{ij}}\biggr)^{6}
449 \biggr],
450 \label{eq:lennardJonesPot}
451 \end{equation}
452 where $r_{ij}$ is the distance between particles $i$ and $j$,
453 $\sigma_{ij}$ scales the length of the interaction, and
454 $\epsilon_{ij}$ scales the well depth of the potential. Scheme
455 \ref{sch:LJFF} gives an example \texttt{.bass} file that
456 sets up a system of 108 Ar particles to be simulated using the
457 Lennard-Jones force field.
458
459 \begin{lstlisting}[float,caption={[Invocation of the Lennard-Jones force field] A sample system using the Lennard-Jones force field.},label={sch:LJFF}]
460
461 #include "argon.mdl"
462
463 nComponents = 1;
464 component{
465 type = "Ar";
466 nMol = 108;
467 }
468
469 initialConfig = "./argon.init";
470
471 forceField = "LJ";
472 \end{lstlisting}
473
474 Because this potential is calculated between all pairs, the force
475 evaluation can become computationally expensive for large systems. To
476 keep the pair evaluations to a manageable number, {\sc oopse} employs
477 a cut-off radius.\cite{allen87:csl} The cutoff radius can either be
478 specified in the \texttt{.bass} file, or left as its default value of
479 $2.5\sigma_{ii}$, where $\sigma_{ii}$ is the largest Lennard-Jones
480 length parameter present in the simulation. Truncating the calculation
481 at $r_{\text{cut}}$ introduces a discontinuity into the potential
482 energy and the force. To offset this discontinuity in the potential,
483 the energy value at $r_{\text{cut}}$ is subtracted from the
484 potential. This causes the potential to go to zero smoothly at the
485 cut-off radius, and preserves conservation of energy in integrating
486 the equations of motion. There still remains a discontinuity in the derivative (the forces), however, this does not significantly affect the dynamics.
487
488 Interactions between dissimilar particles requires the generation of
489 cross term parameters for $\sigma$ and $\epsilon$. These are
490 calculated through the Lorentz-Berthelot mixing
491 rules:\cite{allen87:csl}
492 \begin{equation}
493 \sigma_{ij} = \frac{1}{2}[\sigma_{ii} + \sigma_{jj}],
494 \label{eq:sigmaMix}
495 \end{equation}
496 and
497 \begin{equation}
498 \epsilon_{ij} = \sqrt{\epsilon_{ii} \epsilon_{jj}}.
499 \label{eq:epsilonMix}
500 \end{equation}
501
502 \subsection{\label{oopseSec:DUFF}Dipolar Unified-Atom Force Field}
503
504 The dipolar unified-atom force field ({\sc duff}) was developed to
505 simulate lipid bilayers. The simulations require a model capable of
506 forming bilayers, while still being sufficiently computationally
507 efficient to allow large systems ($\sim$100's of phospholipids,
508 $\sim$1000's of waters) to be simulated for long times
509 ($\sim$10's of nanoseconds).
510
511 With this goal in mind, {\sc duff} has no point
512 charges. Charge-neutral distributions were replaced with dipoles,
513 while most atoms and groups of atoms were reduced to Lennard-Jones
514 interaction sites. This simplification cuts the length scale of long
515 range interactions from $\frac{1}{r}$ to $\frac{1}{r^3}$, and allows
516 us to avoid the computationally expensive Ewald sum. Instead, we can
517 use neighbor-lists and cutoff radii for the dipolar interactions, or
518 include a reaction field to mimic larger range interactions.
519
520 As an example, lipid head-groups in {\sc duff} are represented as
521 point dipole interaction sites. By placing a dipole at the head
522 group's center of mass, our model mimics the charge separation found
523 in common phospholipid head groups such as
524 phosphatidylcholine.\cite{Cevc87} Additionally, a large Lennard-Jones
525 site is located at the pseudoatom's center of mass. The model is
526 illustrated by the red atom in Fig.~\ref{oopseFig:lipidModel}. The
527 water model we use to complement the dipoles of the lipids is our
528 reparameterization of the soft sticky dipole (SSD) model of Ichiye
529 \emph{et al.}\cite{liu96:new_model}
530
531 \begin{figure}
532 \centering
533 \includegraphics[width=\linewidth]{twoChainFig.pdf}
534 \caption[A representation of a lipid model in {\sc duff}]{A representation of the lipid model. $\phi$ is the torsion angle, $\theta$ %
535 is the bend angle, and $\mu$ is the dipole moment of the head group.}
536 \label{oopseFig:lipidModel}
537 \end{figure}
538
539 We have used a set of scalable parameters to model the alkyl groups
540 with Lennard-Jones sites. For this, we have borrowed parameters from
541 the TraPPE force field of Siepmann
542 \emph{et al}.\cite{Siepmann1998} TraPPE is a unified-atom
543 representation of n-alkanes, which is parametrized against phase
544 equilibria using Gibbs ensemble Monte Carlo simulation
545 techniques.\cite{Siepmann1998} One of the advantages of TraPPE is that
546 it generalizes the types of atoms in an alkyl chain to keep the number
547 of pseudoatoms to a minimum; the parameters for a unified atom such as
548 $\text{CH}_2$ do not change depending on what species are bonded to
549 it.
550
551 TraPPE also constrains all bonds to be of fixed length. Typically,
552 bond vibrations are the fastest motions in a molecular dynamic
553 simulation. Small time steps between force evaluations must be used to
554 ensure adequate energy conservation in the bond degrees of freedom. By
555 constraining the bond lengths, larger time steps may be used when
556 integrating the equations of motion. A simulation using {\sc duff} is
557 illustrated in Scheme \ref{sch:DUFF}.
558
559 \begin{lstlisting}[float,caption={[Invocation of {\sc duff}]A portion of a \texttt{.bass} file showing a simulation utilizing {\sc duff}},label={sch:DUFF}]
560
561 #include "water.mdl"
562 #include "lipid.mdl"
563
564 nComponents = 2;
565 component{
566 type = "simpleLipid_16";
567 nMol = 60;
568 }
569
570 component{
571 type = "SSD_water";
572 nMol = 1936;
573 }
574
575 initialConfig = "bilayer.init";
576
577 forceField = "DUFF";
578
579 \end{lstlisting}
580
581 \subsection{\label{oopseSec:energyFunctions}{\sc duff} Energy Functions}
582
583 The total potential energy function in {\sc duff} is
584 \begin{equation}
585 V = \sum^{N}_{I=1} V^{I}_{\text{Internal}}
586 + \sum^{N-1}_{I=1} \sum_{J>I} V^{IJ}_{\text{Cross}},
587 \label{eq:totalPotential}
588 \end{equation}
589 where $V^{I}_{\text{Internal}}$ is the internal potential of molecule $I$:
590 \begin{equation}
591 V^{I}_{\text{Internal}} =
592 \sum_{\theta_{ijk} \in I} V_{\text{bend}}(\theta_{ijk})
593 + \sum_{\phi_{ijkl} \in I} V_{\text{torsion}}(\phi_{ijkl})
594 + \sum_{i \in I} \sum_{(j>i+4) \in I}
595 \biggl[ V_{\text{LJ}}(r_{ij}) + V_{\text{dipole}}
596 (\mathbf{r}_{ij},\boldsymbol{\Omega}_{i},\boldsymbol{\Omega}_{j})
597 \biggr].
598 \label{eq:internalPotential}
599 \end{equation}
600 Here $V_{\text{bend}}$ is the bend potential for all 1, 3 bonded pairs
601 within the molecule $I$, and $V_{\text{torsion}}$ is the torsion potential
602 for all 1, 4 bonded pairs. The pairwise portions of the internal
603 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.
604
605
606 The bend potential of a molecule is represented by the following function:
607 \begin{equation}
608 V_{\text{bend}}(\theta_{ijk}) = k_{\theta}( \theta_{ijk} - \theta_0 )^2, \label{eq:bendPot}
609 \end{equation}
610 where $\theta_{ijk}$ is the angle defined by atoms $i$, $j$, and $k$
611 (see Fig.~\ref{oopseFig:lipidModel}), $\theta_0$ is the equilibrium
612 bond angle, and $k_{\theta}$ is the force constant which determines the
613 strength of the harmonic bend. The parameters for $k_{\theta}$ and
614 $\theta_0$ are borrowed from those in TraPPE.\cite{Siepmann1998}
615
616 The torsion potential and parameters are also borrowed from TraPPE. It is
617 of the form:
618 \begin{equation}
619 V_{\text{torsion}}(\phi) = c_1[1 + \cos \phi]
620 + c_2[1 + \cos(2\phi)]
621 + c_3[1 + \cos(3\phi)],
622 \label{eq:origTorsionPot}
623 \end{equation}
624 where:
625 \begin{equation}
626 \cos\phi = (\hat{\mathbf{r}}_{ij} \times \hat{\mathbf{r}}_{jk}) \cdot
627 (\hat{\mathbf{r}}_{jk} \times \hat{\mathbf{r}}_{kl}).
628 \label{eq:torsPhi}
629 \end{equation}
630 Here, $\hat{\mathbf{r}}_{\alpha\beta}$ are the set of unit bond
631 vectors between atoms $i$, $j$, $k$, and $l$. For computational
632 efficiency, the torsion potential has been recast after the method of
633 {\sc charmm},\cite{Brooks83} in which the angle series is converted to
634 a power series of the form:
635 \begin{equation}
636 V_{\text{torsion}}(\phi) =
637 k_3 \cos^3 \phi + k_2 \cos^2 \phi + k_1 \cos \phi + k_0,
638 \label{eq:torsionPot}
639 \end{equation}
640 where:
641 \begin{align*}
642 k_0 &= c_1 + c_3, \\
643 k_1 &= c_1 - 3c_3, \\
644 k_2 &= 2 c_2, \\
645 k_3 &= 4c_3.
646 \end{align*}
647 By recasting the potential as a power series, repeated trigonometric
648 evaluations are avoided during the calculation of the potential energy.
649
650
651 The cross potential between molecules $I$ and $J$, $V^{IJ}_{\text{Cross}}$, is
652 as follows:
653 \begin{equation}
654 V^{IJ}_{\text{Cross}} =
655 \sum_{i \in I} \sum_{j \in J}
656 \biggl[ V_{\text{LJ}}(r_{ij}) + V_{\text{dipole}}
657 (\mathbf{r}_{ij},\boldsymbol{\Omega}_{i},\boldsymbol{\Omega}_{j})
658 + V_{\text{sticky}}
659 (\mathbf{r}_{ij},\boldsymbol{\Omega}_{i},\boldsymbol{\Omega}_{j})
660 \biggr],
661 \label{eq:crossPotentail}
662 \end{equation}
663 where $V_{\text{LJ}}$ is the Lennard Jones potential,
664 $V_{\text{dipole}}$ is the dipole dipole potential, and
665 $V_{\text{sticky}}$ is the sticky potential defined by the SSD model
666 (Sec.~\ref{oopseSec:SSD}). Note that not all atom types include all
667 interactions.
668
669 The dipole-dipole potential has the following form:
670 \begin{equation}
671 V_{\text{dipole}}(\mathbf{r}_{ij},\boldsymbol{\Omega}_{i},
672 \boldsymbol{\Omega}_{j}) = \frac{|\mu_i||\mu_j|}{4\pi\epsilon_{0}r_{ij}^{3}} \biggl[
673 \boldsymbol{\hat{u}}_{i} \cdot \boldsymbol{\hat{u}}_{j}
674 -
675 3(\boldsymbol{\hat{u}}_i \cdot \hat{\mathbf{r}}_{ij}) %
676 (\boldsymbol{\hat{u}}_j \cdot \hat{\mathbf{r}}_{ij}) \biggr].
677 \label{eq:dipolePot}
678 \end{equation}
679 Here $\mathbf{r}_{ij}$ is the vector starting at atom $i$ pointing
680 towards $j$, and $\boldsymbol{\Omega}_i$ and $\boldsymbol{\Omega}_j$
681 are the orientational degrees of freedom for atoms $i$ and $j$
682 respectively. $|\mu_i|$ is the magnitude of the dipole moment of atom
683 $i$, $\boldsymbol{\hat{u}}_i$ is the standard unit orientation vector
684 of $\boldsymbol{\Omega}_i$, and $\boldsymbol{\hat{r}}_{ij}$ is the
685 unit vector pointing along $\mathbf{r}_{ij}$
686 ($\boldsymbol{\hat{r}}_{ij}=\mathbf{r}_{ij}/|\mathbf{r}_{ij}|$).
687
688 To improve computational efficiency of the dipole-dipole interactions,
689 {\sc oopse} employs an electrostatic cutoff radius. This parameter can
690 be set in the \texttt{.bass} file, and controls the length scale over
691 which dipole interactions are felt. To compensate for the
692 discontinuity in the potential and the forces at the cutoff radius, we
693 have implemented a switching function to smoothly scale the
694 dipole-dipole interaction at the cutoff.
695 \begin{equation}
696 S(r_{ij}) =
697 \begin{cases}
698 1 & \text{if $r_{ij} \le r_t$},\\
699 \frac{(r_{\text{cut}} + 2r_{ij} - 3r_t)(r_{\text{cut}} - r_{ij})^2}
700 {(r_{\text{cut}} - r_t)^2}
701 & \text{if $r_t < r_{ij} \le r_{\text{cut}}$}, \\
702 0 & \text{if $r_{ij} > r_{\text{cut}}$.}
703 \end{cases}
704 \label{eq:dipoleSwitching}
705 \end{equation}
706 Here $S(r_{ij})$ scales the potential at a given $r_{ij}$, and $r_t$
707 is the taper radius some given thickness less than the electrostatic
708 cutoff. The switching thickness can be set in the \texttt{.bass} file.
709
710 \subsection{\label{oopseSec:SSD}The {\sc duff} Water Models: SSD/E and SSD/RF}
711
712 In the interest of computational efficiency, the default solvent used
713 by {\sc oopse} is the extended Soft Sticky Dipole (SSD/E) water
714 model.\cite{fennell04} The original SSD was developed by Ichiye
715 \emph{et al.}\cite{liu96:new_model} as a modified form of the hard-sphere
716 water model proposed by Bratko, Blum, and
717 Luzar.\cite{Bratko85,Bratko95} It consists of a single point dipole
718 with a Lennard-Jones core and a sticky potential that directs the
719 particles to assume the proper hydrogen bond orientation in the first
720 solvation shell. Thus, the interaction between two SSD water molecules
721 \emph{i} and \emph{j} is given by the potential
722 \begin{equation}
723 V_{ij} =
724 V_{ij}^{LJ} (r_{ij})\ + V_{ij}^{dp}
725 (\mathbf{r}_{ij},\boldsymbol{\Omega}_i,\boldsymbol{\Omega}_j)\ +
726 V_{ij}^{sp}
727 (\mathbf{r}_{ij},\boldsymbol{\Omega}_i,\boldsymbol{\Omega}_j),
728 \label{eq:ssdPot}
729 \end{equation}
730 where the $\mathbf{r}_{ij}$ is the position vector between molecules
731 \emph{i} and \emph{j} with magnitude equal to the distance $r_{ij}$, and
732 $\boldsymbol{\Omega}_i$ and $\boldsymbol{\Omega}_j$ represent the
733 orientations of the respective molecules. The Lennard-Jones and dipole
734 parts of the potential are given by equations \ref{eq:lennardJonesPot}
735 and \ref{eq:dipolePot} respectively. The sticky part is described by
736 the following,
737 \begin{equation}
738 u_{ij}^{sp}(\mathbf{r}_{ij},\boldsymbol{\Omega}_i,\boldsymbol{\Omega}_j)=
739 \frac{\nu_0}{2}[s(r_{ij})w(\mathbf{r}_{ij},
740 \boldsymbol{\Omega}_i,\boldsymbol{\Omega}_j) +
741 s^\prime(r_{ij})w^\prime(\mathbf{r}_{ij},
742 \boldsymbol{\Omega}_i,\boldsymbol{\Omega}_j)]\ ,
743 \label{eq:stickyPot}
744 \end{equation}
745 where $\nu_0$ is a strength parameter for the sticky potential, and
746 $s$ and $s^\prime$ are cubic switching functions which turn off the
747 sticky interaction beyond the first solvation shell. The $w$ function
748 can be thought of as an attractive potential with tetrahedral
749 geometry:
750 \begin{equation}
751 w({\bf r}_{ij},{\bf \Omega}_i,{\bf \Omega}_j)=
752 \sin\theta_{ij}\sin2\theta_{ij}\cos2\phi_{ij},
753 \label{eq:stickyW}
754 \end{equation}
755 while the $w^\prime$ function counters the normal aligned and
756 anti-aligned structures favored by point dipoles:
757 \begin{equation}
758 w^\prime({\bf r}_{ij},{\bf \Omega}_i,{\bf \Omega}_j)=
759 (\cos\theta_{ij}-0.6)^2(\cos\theta_{ij}+0.8)^2-w^0,
760 \label{eq:stickyWprime}
761 \end{equation}
762 It should be noted that $w$ is proportional to the sum of the $Y_3^2$
763 and $Y_3^{-2}$ spherical harmonics (a linear combination which
764 enhances the tetrahedral geometry for hydrogen bonded structures),
765 while $w^\prime$ is a purely empirical function. A more detailed
766 description of the functional parts and variables in this potential
767 can be found in the original SSD
768 articles.\cite{liu96:new_model,liu96:monte_carlo,chandra99:ssd_md,Ichiye03}
769
770 Since SSD/E is a single-point {\it dipolar} model, the force
771 calculations are simplified significantly relative to the standard
772 {\it charged} multi-point models. In the original Monte Carlo
773 simulations using this model, Ichiye {\it et al.} reported that using
774 SSD decreased computer time by a factor of 6-7 compared to other
775 models.\cite{liu96:new_model} What is most impressive is that these savings
776 did not come at the expense of accurate depiction of the liquid state
777 properties. Indeed, SSD/E maintains reasonable agreement with the Head-Gordon
778 diffraction data for the structural features of liquid
779 water.\cite{hura00,liu96:new_model} Additionally, the dynamical properties
780 exhibited by SSD/E agree with experiment better than those of more
781 computationally expensive models (like TIP3P and
782 SPC/E).\cite{chandra99:ssd_md} The combination of speed and accurate depiction
783 of solvent properties makes SSD/E a very attractive model for the
784 simulation of large scale biochemical simulations.
785
786 Recent constant pressure simulations revealed issues in the original
787 SSD model that led to lower than expected densities at all target
788 pressures.\cite{Ichiye03,fennell04} The default model in {\sc oopse}
789 is therefore SSD/E, a density corrected derivative of SSD that
790 exhibits improved liquid structure and transport behavior. If the use
791 of a reaction field long-range interaction correction is desired, it
792 is recommended that the parameters be modified to those of the SSD/RF
793 model (an SSD variant parameterized for reaction field). Solvent parameters can be easily modified in an accompanying
794 \texttt{.bass} file as illustrated in the scheme below. A table of the
795 parameter values and the drawbacks and benefits of the different
796 density corrected SSD models can be found in
797 reference~\cite{fennell04}.
798
799 \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}]
800
801 #include "water.mdl"
802
803 nComponents = 1;
804 component{
805 type = "SSD_water";
806 nMol = 864;
807 }
808
809 initialConfig = "liquidWater.init";
810
811 forceField = "DUFF";
812
813 /*
814 * The following two flags set the cutoff
815 * radius for the electrostatic forces
816 * as well as the skin thickness of the switching
817 * function.
818 */
819
820 electrostaticCutoffRadius = 9.2;
821 electrostaticSkinThickness = 1.38;
822
823 \end{lstlisting}
824
825
826 \subsection{\label{oopseSec:eam}Embedded Atom Method}
827
828 There are Molecular Dynamics packages which have the
829 capacity to simulate metallic systems, including some that have
830 parallel computational abilities\cite{plimpton93}. Potentials that
831 describe bonding transition metal
832 systems\cite{Finnis84,Ercolessi88,Chen90,Qi99,Ercolessi02} have an
833 attractive interaction which models ``Embedding''
834 a positively charged metal ion in the electron density due to the
835 free valance ``sea'' of electrons created by the surrounding atoms in
836 the system. A mostly-repulsive pairwise part of the potential
837 describes the interaction of the positively charged metal core ions
838 with one another. A particular potential description called the
839 Embedded Atom Method\cite{Daw84,FBD86,johnson89,Lu97}({\sc eam}) that has
840 particularly wide adoption has been selected for inclusion in {\sc oopse}. A
841 good review of {\sc eam} and other metallic potential formulations was written
842 by Voter.\cite{voter}
843
844 The {\sc eam} potential has the form:
845 \begin{eqnarray}
846 V & = & \sum_{i} F_{i}\left[\rho_{i}\right] + \sum_{i} \sum_{j \neq i}
847 \phi_{ij}({\bf r}_{ij}), \\
848 \rho_{i} & = & \sum_{j \neq i} f_{j}({\bf r}_{ij}),
849 \end{eqnarray}
850 where $F_{i} $ is the embedding function that equates the energy
851 required to embed a positively-charged core ion $i$ into a linear
852 superposition of spherically averaged atomic electron densities given
853 by $\rho_{i}$. $\phi_{ij}$ is a primarily repulsive pairwise
854 interaction between atoms $i$ and $j$. In the original formulation of
855 {\sc eam}\cite{Daw84}, $\phi_{ij}$ was an entirely repulsive term,
856 however in later refinements to {\sc eam} have shown that non-uniqueness
857 between $F$ and $\phi$ allow for more general forms for
858 $\phi$.\cite{Daw89} There is a cutoff distance, $r_{cut}$, which
859 limits the summations in the {\sc eam} equation to the few dozen atoms
860 surrounding atom $i$ for both the density $\rho$ and pairwise $\phi$
861 interactions. Foiles \emph{et al}.~fit {\sc eam} potentials for the fcc
862 metals Cu, Ag, Au, Ni, Pd, Pt and alloys of these metals.\cite{FBD86}
863 These fits are included in {\sc oopse}.
864
865 \subsection{\label{oopseSec:pbc}Periodic Boundary Conditions}
866
867 \newcommand{\roundme}{\operatorname{round}}
868
869 \textit{Periodic boundary conditions} are widely used to simulate bulk properties with a relatively small number of particles. The
870 simulation box is replicated throughout space to form an infinite
871 lattice. During the simulation, when a particle moves in the primary
872 cell, its image in other cells move in exactly the same direction with
873 exactly the same orientation. Thus, as a particle leaves the primary
874 cell, one of its images will enter through the opposite face. If the
875 simulation box is large enough to avoid ``feeling'' the symmetries of
876 the periodic lattice, surface effects can be ignored. The available
877 periodic cells in OOPSE are cubic, orthorhombic and parallelepiped. We
878 use a $3 \times 3$ matrix, $\mathsf{H}$, to describe the shape and
879 size of the simulation box. $\mathsf{H}$ is defined:
880 \begin{equation}
881 \mathsf{H} = ( \mathbf{h}_x, \mathbf{h}_y, \mathbf{h}_z ),
882 \end{equation}
883 where $\mathbf{h}_{\alpha}$ is the column vector of the $\alpha$ axis of the
884 box. During the course of the simulation both the size and shape of
885 the box can be changed to allow volume fluctuations when constraining
886 the pressure.
887
888 A real space vector, $\mathbf{r}$ can be transformed in to a box space
889 vector, $\mathbf{s}$, and back through the following transformations:
890 \begin{align}
891 \mathbf{s} &= \mathsf{H}^{-1} \mathbf{r}, \\
892 \mathbf{r} &= \mathsf{H} \mathbf{s}.
893 \end{align}
894 The vector $\mathbf{s}$ is now a vector expressed as the number of box
895 lengths in the $\mathbf{h}_x$, $\mathbf{h}_y$, and $\mathbf{h}_z$
896 directions. To find the minimum image of a vector $\mathbf{r}$, we
897 first convert it to its corresponding vector in box space, and then,
898 cast each element to lie in the range $[-0.5,0.5]$:
899 \begin{equation}
900 s_{i}^{\prime}=s_{i}-\roundme(s_{i}),
901 \end{equation}
902 where $s_i$ is the $i$th element of $\mathbf{s}$, and
903 $\roundme(s_i)$ is given by
904 \begin{equation}
905 \roundme(x) =
906 \begin{cases}
907 \lfloor x+0.5 \rfloor & \text{if $x \ge 0$,} \\
908 \lceil x-0.5 \rceil & \text{if $x < 0$.}
909 \end{cases}
910 \end{equation}
911 Here $\lfloor x \rfloor$ is the floor operator, and gives the largest
912 integer value that is not greater than $x$, and $\lceil x \rceil$ is
913 the ceiling operator, and gives the smallest integer that is not less
914 than $x$. For example, $\roundme(3.6)=4$, $\roundme(3.1)=3$,
915 $\roundme(-3.6)=-4$, $\roundme(-3.1)=-3$.
916
917 Finally, we obtain the minimum image coordinates $\mathbf{r}^{\prime}$ by
918 transforming back to real space,
919 \begin{equation}
920 \mathbf{r}^{\prime}=\mathsf{H}^{-1}\mathbf{s}^{\prime}.%
921 \end{equation}
922 In this way, particles are allowed to diffuse freely in $\mathbf{r}$,
923 but their minimum images, $\mathbf{r}^{\prime}$ are used to compute
924 the inter-atomic forces.
925
926
927
928 \section{\label{oopseSec:mechanics}Mechanics}
929
930 \subsection{\label{oopseSec:integrate}Integrating the Equations of Motion: the
931 DLM method}
932
933 The default method for integrating the equations of motion in {\sc
934 oopse} is a velocity-Verlet version of the symplectic splitting method
935 proposed by Dullweber, Leimkuhler and McLachlan
936 (DLM).\cite{Dullweber1997} When there are no directional atoms or
937 rigid bodies present in the simulation, this integrator becomes the
938 standard velocity-Verlet integrator which is known to sample the
939 microcanonical (NVE) ensemble.\cite{Frenkel1996}
940
941 Previous integration methods for orientational motion have problems
942 that are avoided in the DLM method. Direct propagation of the Euler
943 angles has a known $1/\sin\theta$ divergence in the equations of
944 motion for $\phi$ and $\psi$,\cite{allen87:csl} leading to
945 numerical instabilities any time one of the directional atoms or rigid
946 bodies has an orientation near $\theta=0$ or $\theta=\pi$. More
947 modern quaternion-based integration methods have relatively poor
948 energy conservation. While quaternions work well for orientational
949 motion in other ensembles, the microcanonical ensemble has a
950 constant energy requirement that is quite sensitive to errors in the
951 equations of motion. An earlier implementation of {\sc oopse}
952 utilized quaternions for propagation of rotational motion; however, a
953 detailed investigation showed that they resulted in a steady drift in
954 the total energy, something that has been observed by
955 Laird {\it et al.}\cite{Laird97}
956
957 The key difference in the integration method proposed by Dullweber
958 \emph{et al.} is that the entire $3 \times 3$ rotation matrix is
959 propagated from one time step to the next. In the past, this would not
960 have been feasible, since the rotation matrix for a single body has
961 nine elements compared with the more memory-efficient methods (using
962 three Euler angles or 4 quaternions). Computer memory has become much
963 less costly in recent years, and this can be translated into
964 substantial benefits in energy conservation.
965
966 The basic equations of motion being integrated are derived from the
967 Hamiltonian for conservative systems containing rigid bodies,
968 \begin{equation}
969 H = \sum_{i} \left( \frac{1}{2} m_i {\bf v}_i^T \cdot {\bf v}_i +
970 \frac{1}{2} {\bf j}_i^T \cdot \overleftrightarrow{\mathsf{I}}_i^{-1} \cdot
971 {\bf j}_i \right) +
972 V\left(\left\{{\bf r}\right\}, \left\{\mathsf{A}\right\}\right),
973 \end{equation}
974 where ${\bf r}_i$ and ${\bf v}_i$ are the cartesian position vector
975 and velocity of the center of mass of particle $i$, and ${\bf j}_i$,
976 $\overleftrightarrow{\mathsf{I}}_i$ are the body-fixed angular
977 momentum and moment of inertia tensor respectively, and the
978 superscript $T$ denotes the transpose of the vector. $\mathsf{A}_i$
979 is the $3 \times 3$ rotation matrix describing the instantaneous
980 orientation of the particle. $V$ is the potential energy function
981 which may depend on both the positions $\left\{{\bf r}\right\}$ and
982 orientations $\left\{\mathsf{A}\right\}$ of all particles. The
983 equations of motion for the particle centers of mass are derived from
984 Hamilton's equations and are quite simple,
985 \begin{eqnarray}
986 \dot{{\bf r}} & = & {\bf v}, \\
987 \dot{{\bf v}} & = & \frac{{\bf f}}{m},
988 \end{eqnarray}
989 where ${\bf f}$ is the instantaneous force on the center of mass
990 of the particle,
991 \begin{equation}
992 {\bf f} = - \frac{\partial}{\partial
993 {\bf r}} V(\left\{{\bf r}(t)\right\}, \left\{\mathsf{A}(t)\right\}).
994 \end{equation}
995
996 The equations of motion for the orientational degrees of freedom are
997 \begin{eqnarray}
998 \dot{\mathsf{A}} & = & \mathsf{A} \cdot
999 \mbox{ skew}\left(\overleftrightarrow{\mathsf{I}}^{-1} \cdot {\bf j}\right),\\
1000 \dot{{\bf j}} & = & {\bf j} \times \left( \overleftrightarrow{\mathsf{I}}^{-1}
1001 \cdot {\bf j} \right) - \mbox{ rot}\left(\mathsf{A}^{T} \cdot \frac{\partial
1002 V}{\partial \mathsf{A}} \right).
1003 \end{eqnarray}
1004 In these equations of motion, the $\mbox{skew}$ matrix of a vector
1005 ${\bf v} = \left( v_1, v_2, v_3 \right)$ is defined:
1006 \begin{equation}
1007 \mbox{skew}\left( {\bf v} \right) := \left(
1008 \begin{array}{ccc}
1009 0 & v_3 & - v_2 \\
1010 -v_3 & 0 & v_1 \\
1011 v_2 & -v_1 & 0
1012 \end{array}
1013 \right).
1014 \end{equation}
1015 The $\mbox{rot}$ notation refers to the mapping of the $3 \times 3$
1016 rotation matrix to a vector of orientations by first computing the
1017 skew-symmetric part $\left(\mathsf{A} - \mathsf{A}^{T}\right)$ and
1018 then associating this with a length 3 vector by inverting the
1019 $\mbox{skew}$ function above:
1020 \begin{equation}
1021 \mbox{rot}\left(\mathsf{A}\right) := \mbox{ skew}^{-1}\left(\mathsf{A}
1022 - \mathsf{A}^{T} \right).
1023 \end{equation}
1024 Written this way, the $\mbox{rot}$ operation creates a set of
1025 conjugate angle coordinates to the body-fixed angular momenta
1026 represented by ${\bf j}$. This equation of motion for angular momenta
1027 is equivalent to the more familiar body-fixed forms,
1028 \begin{eqnarray}
1029 \dot{j_{x}} & = & \tau^b_x(t) +
1030 \left(\overleftrightarrow{\mathsf{I}}_{yy} - \overleftrightarrow{\mathsf{I}}_{zz} \right) j_y j_z, \\
1031 \dot{j_{y}} & = & \tau^b_y(t) +
1032 \left(\overleftrightarrow{\mathsf{I}}_{zz} - \overleftrightarrow{\mathsf{I}}_{xx} \right) j_z j_x,\\
1033 \dot{j_{z}} & = & \tau^b_z(t) +
1034 \left(\overleftrightarrow{\mathsf{I}}_{xx} - \overleftrightarrow{\mathsf{I}}_{yy} \right) j_x j_y,
1035 \end{eqnarray}
1036 which utilize the body-fixed torques, ${\bf \tau}^b$. Torques are
1037 most easily derived in the space-fixed frame,
1038 \begin{equation}
1039 {\bf \tau}^b(t) = \mathsf{A}(t) \cdot {\bf \tau}^s(t),
1040 \end{equation}
1041 where the torques are either derived from the forces on the
1042 constituent atoms of the rigid body, or for directional atoms,
1043 directly from derivatives of the potential energy,
1044 \begin{equation}
1045 {\bf \tau}^s(t) = - \hat{\bf u}(t) \times \left( \frac{\partial}
1046 {\partial \hat{\bf u}} V\left(\left\{ {\bf r}(t) \right\}, \left\{
1047 \mathsf{A}(t) \right\}\right) \right).
1048 \end{equation}
1049 Here $\hat{\bf u}$ is a unit vector pointing along the principal axis
1050 of the particle in the space-fixed frame.
1051
1052 The DLM method uses a Trotter factorization of the orientational
1053 propagator. This has three effects:
1054 \begin{enumerate}
1055 \item the integrator is area-preserving in phase space (i.e. it is
1056 {\it symplectic}),
1057 \item the integrator is time-{\it reversible}, making it suitable for Hybrid
1058 Monte Carlo applications, and
1059 \item the error for a single time step is of order $\mathcal{O}\left(h^4\right)$
1060 for timesteps of length $h$.
1061 \end{enumerate}
1062
1063 The integration of the equations of motion is carried out in a
1064 velocity-Verlet style 2-part algorithm, where $h= \delta t$:
1065
1066 {\tt moveA:}
1067 \begin{align*}
1068 {\bf v}\left(t + h / 2\right) &\leftarrow {\bf v}(t)
1069 + \frac{h}{2} \left( {\bf f}(t) / m \right), \\
1070 %
1071 {\bf r}(t + h) &\leftarrow {\bf r}(t)
1072 + h {\bf v}\left(t + h / 2 \right), \\
1073 %
1074 {\bf j}\left(t + h / 2 \right) &\leftarrow {\bf j}(t)
1075 + \frac{h}{2} {\bf \tau}^b(t), \\
1076 %
1077 \mathsf{A}(t + h) &\leftarrow \mathrm{rotate}\left( h {\bf j}
1078 (t + h / 2) \cdot \overleftrightarrow{\mathsf{I}}^{-1} \right).
1079 \end{align*}
1080
1081 In this context, the $\mathrm{rotate}$ function is the reversible product
1082 of the three body-fixed rotations,
1083 \begin{equation}
1084 \mathrm{rotate}({\bf a}) = \mathsf{G}_x(a_x / 2) \cdot
1085 \mathsf{G}_y(a_y / 2) \cdot \mathsf{G}_z(a_z) \cdot \mathsf{G}_y(a_y /
1086 2) \cdot \mathsf{G}_x(a_x /2),
1087 \end{equation}
1088 where each rotational propagator, $\mathsf{G}_\alpha(\theta)$, rotates
1089 both the rotation matrix ($\mathsf{A}$) and the body-fixed angular
1090 momentum (${\bf j}$) by an angle $\theta$ around body-fixed axis
1091 $\alpha$,
1092 \begin{equation}
1093 \mathsf{G}_\alpha( \theta ) = \left\{
1094 \begin{array}{lcl}
1095 \mathsf{A}(t) & \leftarrow & \mathsf{A}(0) \cdot \mathsf{R}_\alpha(\theta)^T, \\
1096 {\bf j}(t) & \leftarrow & \mathsf{R}_\alpha(\theta) \cdot {\bf j}(0).
1097 \end{array}
1098 \right.
1099 \end{equation}
1100 $\mathsf{R}_\alpha$ is a quadratic approximation to
1101 the single-axis rotation matrix. For example, in the small-angle
1102 limit, the rotation matrix around the body-fixed x-axis can be
1103 approximated as
1104 \begin{equation}
1105 \mathsf{R}_x(\theta) \approx \left(
1106 \begin{array}{ccc}
1107 1 & 0 & 0 \\
1108 0 & \frac{1-\theta^2 / 4}{1 + \theta^2 / 4} & -\frac{\theta}{1+
1109 \theta^2 / 4} \\
1110 0 & \frac{\theta}{1+
1111 \theta^2 / 4} & \frac{1-\theta^2 / 4}{1 + \theta^2 / 4}
1112 \end{array}
1113 \right).
1114 \end{equation}
1115 All other rotations follow in a straightforward manner.
1116
1117 After the first part of the propagation, the forces and body-fixed
1118 torques are calculated at the new positions and orientations
1119
1120 {\tt doForces:}
1121 \begin{align*}
1122 {\bf f}(t + h) &\leftarrow
1123 - \left(\frac{\partial V}{\partial {\bf r}}\right)_{{\bf r}(t + h)}, \\
1124 %
1125 {\bf \tau}^{s}(t + h) &\leftarrow {\bf u}(t + h)
1126 \times \frac{\partial V}{\partial {\bf u}}, \\
1127 %
1128 {\bf \tau}^{b}(t + h) &\leftarrow \mathsf{A}(t + h)
1129 \cdot {\bf \tau}^s(t + h).
1130 \end{align*}
1131
1132 {\sc oopse} automatically updates ${\bf u}$ when the rotation matrix
1133 $\mathsf{A}$ is calculated in {\tt moveA}. Once the forces and
1134 torques have been obtained at the new time step, the velocities can be
1135 advanced to the same time value.
1136
1137 {\tt moveB:}
1138 \begin{align*}
1139 {\bf v}\left(t + h \right) &\leftarrow {\bf v}\left(t + h / 2 \right)
1140 + \frac{h}{2} \left( {\bf f}(t + h) / m \right), \\
1141 %
1142 {\bf j}\left(t + h \right) &\leftarrow {\bf j}\left(t + h / 2 \right)
1143 + \frac{h}{2} {\bf \tau}^b(t + h) .
1144 \end{align*}
1145
1146 The matrix rotations used in the DLM method end up being more costly
1147 computationally than the simpler arithmetic quaternion
1148 propagation. With the same time step, a 1000-molecule water simulation
1149 shows an average 7\% increase in computation time using the DLM method
1150 in place of quaternions. This cost is more than justified when
1151 comparing the energy conservation of the two methods as illustrated in
1152 Fig.~\ref{timestep}.
1153
1154 \begin{figure}
1155 \centering
1156 \includegraphics[width=\linewidth]{timeStep.pdf}
1157 \caption[Energy conservation for quaternion versus DLM dynamics]{Energy conservation using quaternion based integration versus
1158 the method proposed by Dullweber \emph{et al.} with increasing time
1159 step. For each time step, the dotted line is total energy using the
1160 DLM integrator, and the solid line comes from the quaternion
1161 integrator. The larger time step plots are shifted up from the true
1162 energy baseline for clarity.}
1163 \label{timestep}
1164 \end{figure}
1165
1166 In Fig.~\ref{timestep}, the resulting energy drift at various time
1167 steps for both the DLM and quaternion integration schemes is
1168 compared. All of the 1000 molecule water simulations started with the
1169 same configuration, and the only difference was the method for
1170 handling rotational motion. At time steps of 0.1 and 0.5 fs, both
1171 methods for propagating molecule rotation conserve energy fairly well,
1172 with the quaternion method showing a slight energy drift over time in
1173 the 0.5 fs time step simulation. At time steps of 1 and 2 fs, the
1174 energy conservation benefits of the DLM method are clearly
1175 demonstrated. Thus, while maintaining the same degree of energy
1176 conservation, one can take considerably longer time steps, leading to
1177 an overall reduction in computation time.
1178
1179 There is only one specific keyword relevant to the default integrator,
1180 and that is the time step for integrating the equations of motion.
1181
1182 \begin{center}
1183 \begin{tabular}{llll}
1184 {\bf variable} & {\bf {\tt .bass} keyword} & {\bf units} & {\bf
1185 default value} \\
1186 $h$ & {\tt dt = 2.0;} & fs & none
1187 \end{tabular}
1188 \end{center}
1189
1190 \subsection{\label{sec:extended}Extended Systems for other Ensembles}
1191
1192 {\sc oopse} implements a number of extended system integrators for
1193 sampling from other ensembles relevant to chemical physics. The
1194 integrator can selected with the {\tt ensemble} keyword in the
1195 {\tt .bass} file:
1196
1197 \begin{center}
1198 \begin{tabular}{lll}
1199 {\bf Integrator} & {\bf Ensemble} & {\bf {\tt .bass} line} \\
1200 NVE & microcanonical & {\tt ensemble = NVE; } \\
1201 NVT & canonical & {\tt ensemble = NVT; } \\
1202 NPTi & isobaric-isothermal & {\tt ensemble = NPTi;} \\
1203 & (with isotropic volume changes) & \\
1204 NPTf & isobaric-isothermal & {\tt ensemble = NPTf;} \\
1205 & (with changes to box shape) & \\
1206 NPTxyz & approximate isobaric-isothermal & {\tt ensemble = NPTxyz;} \\
1207 & (with separate barostats on each box dimension) & \\
1208 \end{tabular}
1209 \end{center}
1210
1211 The relatively well-known Nos\'e-Hoover thermostat\cite{Hoover85} is
1212 implemented in {\sc oopse}'s NVT integrator. This method couples an
1213 extra degree of freedom (the thermostat) to the kinetic energy of the
1214 system, and has been shown to sample the canonical distribution in the
1215 system degrees of freedom while conserving a quantity that is, to
1216 within a constant, the Helmholtz free energy.\cite{melchionna93}
1217
1218 NPT algorithms attempt to maintain constant pressure in the system by
1219 coupling the volume of the system to a barostat. {\sc oopse} contains
1220 three different constant pressure algorithms. The first two, NPTi and
1221 NPTf have been shown to conserve a quantity that is, to within a
1222 constant, the Gibbs free energy.\cite{melchionna93} The Melchionna
1223 modification to the Hoover barostat is implemented in both NPTi and
1224 NPTf. NPTi allows only isotropic changes in the simulation box, while
1225 box {\it shape} variations are allowed in NPTf. The NPTxyz integrator
1226 has {\it not} been shown to sample from the isobaric-isothermal
1227 ensemble. It is useful, however, in that it maintains orthogonality
1228 for the axes of the simulation box while attempting to equalize
1229 pressure along the three perpendicular directions in the box.
1230
1231 Each of the extended system integrators requires additional keywords
1232 to set target values for the thermodynamic state variables that are
1233 being held constant. Keywords are also required to set the
1234 characteristic decay times for the dynamics of the extended
1235 variables.
1236
1237 \begin{center}
1238 \begin{tabular}{llll}
1239 {\bf variable} & {\bf {\tt .bass} keyword} & {\bf units} & {\bf
1240 default value} \\
1241 $T_{\mathrm{target}}$ & {\tt targetTemperature = 300;} & K & none \\
1242 $P_{\mathrm{target}}$ & {\tt targetPressure = 1;} & atm & none \\
1243 $\tau_T$ & {\tt tauThermostat = 1e3;} & fs & none \\
1244 $\tau_B$ & {\tt tauBarostat = 5e3;} & fs & none \\
1245 & {\tt resetTime = 200;} & fs & none \\
1246 & {\tt useInitialExtendedSystemState = true;} & logical &
1247 true
1248 \end{tabular}
1249 \end{center}
1250
1251 Two additional keywords can be used to either clear the extended
1252 system variables periodically ({\tt resetTime}), or to maintain the
1253 state of the extended system variables between simulations ({\tt
1254 useInitialExtendedSystemState}). More details on these variables
1255 and their use in the integrators follows below.
1256
1257 \subsection{\label{oopseSec:noseHooverThermo}Nos\'{e}-Hoover Thermostatting}
1258
1259 The Nos\'e-Hoover equations of motion are given by\cite{Hoover85}
1260 \begin{eqnarray}
1261 \dot{{\bf r}} & = & {\bf v}, \\
1262 \dot{{\bf v}} & = & \frac{{\bf f}}{m} - \chi {\bf v} ,\\
1263 \dot{\mathsf{A}} & = & \mathsf{A} \cdot
1264 \mbox{ skew}\left(\overleftrightarrow{\mathsf{I}}^{-1} \cdot {\bf j}\right), \\
1265 \dot{{\bf j}} & = & {\bf j} \times \left( \overleftrightarrow{\mathsf{I}}^{-1}
1266 \cdot {\bf j} \right) - \mbox{ rot}\left(\mathsf{A}^{T} \cdot \frac{\partial
1267 V}{\partial \mathsf{A}} \right) - \chi {\bf j}.
1268 \label{eq:nosehoovereom}
1269 \end{eqnarray}
1270
1271 $\chi$ is an ``extra'' variable included in the extended system, and
1272 it is propagated using the first order equation of motion
1273 \begin{equation}
1274 \dot{\chi} = \frac{1}{\tau_{T}^2} \left( \frac{T}{T_{\mathrm{target}}} - 1 \right).
1275 \label{eq:nosehooverext}
1276 \end{equation}
1277
1278 The instantaneous temperature $T$ is proportional to the total kinetic
1279 energy (both translational and orientational) and is given by
1280 \begin{equation}
1281 T = \frac{2 K}{f k_B}
1282 \end{equation}
1283 Here, $f$ is the total number of degrees of freedom in the system,
1284 \begin{equation}
1285 f = 3 N + 3 N_{\mathrm{orient}} - N_{\mathrm{constraints}},
1286 \end{equation}
1287 and $K$ is the total kinetic energy,
1288 \begin{equation}
1289 K = \sum_{i=1}^{N} \frac{1}{2} m_i {\bf v}_i^T \cdot {\bf v}_i +
1290 \sum_{i=1}^{N_{\mathrm{orient}}} \frac{1}{2} {\bf j}_i^T \cdot
1291 \overleftrightarrow{\mathsf{I}}_i^{-1} \cdot {\bf j}_i.
1292 \end{equation}
1293
1294 In eq.(\ref{eq:nosehooverext}), $\tau_T$ is the time constant for
1295 relaxation of the temperature to the target value. To set values for
1296 $\tau_T$ or $T_{\mathrm{target}}$ in a simulation, one would use the
1297 {\tt tauThermostat} and {\tt targetTemperature} keywords in the {\tt
1298 .bass} file. The units for {\tt tauThermostat} are fs, and the units
1299 for the {\tt targetTemperature} are degrees K. The integration of
1300 the equations of motion is carried out in a velocity-Verlet style 2
1301 part algorithm:
1302
1303 {\tt moveA:}
1304 \begin{align*}
1305 T(t) &\leftarrow \left\{{\bf v}(t)\right\}, \left\{{\bf j}(t)\right\} ,\\
1306 %
1307 {\bf v}\left(t + h / 2\right) &\leftarrow {\bf v}(t)
1308 + \frac{h}{2} \left( \frac{{\bf f}(t)}{m} - {\bf v}(t)
1309 \chi(t)\right), \\
1310 %
1311 {\bf r}(t + h) &\leftarrow {\bf r}(t)
1312 + h {\bf v}\left(t + h / 2 \right) ,\\
1313 %
1314 {\bf j}\left(t + h / 2 \right) &\leftarrow {\bf j}(t)
1315 + \frac{h}{2} \left( {\bf \tau}^b(t) - {\bf j}(t)
1316 \chi(t) \right) ,\\
1317 %
1318 \mathsf{A}(t + h) &\leftarrow \mathrm{rotate}
1319 \left(h * {\bf j}(t + h / 2)
1320 \overleftrightarrow{\mathsf{I}}^{-1} \right) ,\\
1321 %
1322 \chi\left(t + h / 2 \right) &\leftarrow \chi(t)
1323 + \frac{h}{2 \tau_T^2} \left( \frac{T(t)}
1324 {T_{\mathrm{target}}} - 1 \right) .
1325 \end{align*}
1326
1327 Here $\mathrm{rotate}(h * {\bf j}
1328 \overleftrightarrow{\mathsf{I}}^{-1})$ is the same symplectic Trotter
1329 factorization of the three rotation operations that was discussed in
1330 the section on the DLM integrator. Note that this operation modifies
1331 both the rotation matrix $\mathsf{A}$ and the angular momentum ${\bf
1332 j}$. {\tt moveA} propagates velocities by a half time step, and
1333 positional degrees of freedom by a full time step. The new positions
1334 (and orientations) are then used to calculate a new set of forces and
1335 torques in exactly the same way they are calculated in the {\tt
1336 doForces} portion of the DLM integrator.
1337
1338 Once the forces and torques have been obtained at the new time step,
1339 the temperature, velocities, and the extended system variable can be
1340 advanced to the same time value.
1341
1342 {\tt moveB:}
1343 \begin{align*}
1344 T(t + h) &\leftarrow \left\{{\bf v}(t + h)\right\},
1345 \left\{{\bf j}(t + h)\right\}, \\
1346 %
1347 \chi\left(t + h \right) &\leftarrow \chi\left(t + h /
1348 2 \right) + \frac{h}{2 \tau_T^2} \left( \frac{T(t+h)}
1349 {T_{\mathrm{target}}} - 1 \right), \\
1350 %
1351 {\bf v}\left(t + h \right) &\leftarrow {\bf v}\left(t
1352 + h / 2 \right) + \frac{h}{2} \left(
1353 \frac{{\bf f}(t + h)}{m} - {\bf v}(t + h)
1354 \chi(t h)\right) ,\\
1355 %
1356 {\bf j}\left(t + h \right) &\leftarrow {\bf j}\left(t
1357 + h / 2 \right) + \frac{h}{2}
1358 \left( {\bf \tau}^b(t + h) - {\bf j}(t + h)
1359 \chi(t + h) \right) .
1360 \end{align*}
1361
1362 Since ${\bf v}(t + h)$ and ${\bf j}(t + h)$ are required to caclculate
1363 $T(t + h)$ as well as $\chi(t + h)$, they indirectly depend on their
1364 own values at time $t + h$. {\tt moveB} is therefore done in an
1365 iterative fashion until $\chi(t + h)$ becomes self-consistent. The
1366 relative tolerance for the self-consistency check defaults to a value
1367 of $\mbox{10}^{-6}$, but {\sc oopse} will terminate the iteration
1368 after 4 loops even if the consistency check has not been satisfied.
1369
1370 The Nos\'e-Hoover algorithm is known to conserve a Hamiltonian for the
1371 extended system that is, to within a constant, identical to the
1372 Helmholtz free energy,\cite{melchionna93}
1373 \begin{equation}
1374 H_{\mathrm{NVT}} = V + K + f k_B T_{\mathrm{target}} \left(
1375 \frac{\tau_{T}^2 \chi^2(t)}{2} + \int_{0}^{t} \chi(t^\prime) dt^\prime
1376 \right).
1377 \end{equation}
1378 Poor choices of $h$ or $\tau_T$ can result in non-conservation
1379 of $H_{\mathrm{NVT}}$, so the conserved quantity is maintained in the
1380 last column of the {\tt .stat} file to allow checks on the quality of
1381 the integration.
1382
1383 Bond constraints are applied at the end of both the {\tt moveA} and
1384 {\tt moveB} portions of the algorithm. Details on the constraint
1385 algorithms are given in section \ref{oopseSec:rattle}.
1386
1387 \subsection{\label{sec:NPTi}Constant-pressure integration with
1388 isotropic box deformations (NPTi)}
1389
1390 To carry out isobaric-isothermal ensemble calculations {\sc oopse}
1391 implements the Melchionna modifications to the Nos\'e-Hoover-Andersen
1392 equations of motion,\cite{melchionna93}
1393
1394 \begin{eqnarray}
1395 \dot{{\bf r}} & = & {\bf v} + \eta \left( {\bf r} - {\bf R}_0 \right), \\
1396 \dot{{\bf v}} & = & \frac{{\bf f}}{m} - (\eta + \chi) {\bf v}, \\
1397 \dot{\mathsf{A}} & = & \mathsf{A} \cdot
1398 \mbox{ skew}\left(\overleftrightarrow{I}^{-1} \cdot {\bf j}\right),\\
1399 \dot{{\bf j}} & = & {\bf j} \times \left( \overleftrightarrow{I}^{-1}
1400 \cdot {\bf j} \right) - \mbox{ rot}\left(\mathsf{A}^{T} \cdot \frac{\partial
1401 V}{\partial \mathsf{A}} \right) - \chi {\bf j}, \\
1402 \dot{\chi} & = & \frac{1}{\tau_{T}^2} \left(
1403 \frac{T}{T_{\mathrm{target}}} - 1 \right) ,\\
1404 \dot{\eta} & = & \frac{1}{\tau_{B}^2 f k_B T_{\mathrm{target}}} V \left( P -
1405 P_{\mathrm{target}} \right), \\
1406 \dot{\mathcal{V}} & = & 3 \mathcal{V} \eta .
1407 \label{eq:melchionna1}
1408 \end{eqnarray}
1409
1410 $\chi$ and $\eta$ are the ``extra'' degrees of freedom in the extended
1411 system. $\chi$ is a thermostat, and it has the same function as it
1412 does in the Nos\'e-Hoover NVT integrator. $\eta$ is a barostat which
1413 controls changes to the volume of the simulation box. ${\bf R}_0$ is
1414 the location of the center of mass for the entire system, and
1415 $\mathcal{V}$ is the volume of the simulation box. At any time, the
1416 volume can be calculated from the determinant of the matrix which
1417 describes the box shape:
1418 \begin{equation}
1419 \mathcal{V} = \det(\mathsf{H}).
1420 \end{equation}
1421
1422 The NPTi integrator requires an instantaneous pressure. This quantity
1423 is calculated via the pressure tensor,
1424 \begin{equation}
1425 \overleftrightarrow{\mathsf{P}}(t) = \frac{1}{\mathcal{V}(t)} \left(
1426 \sum_{i=1}^{N} m_i {\bf v}_i(t) \otimes {\bf v}_i(t) \right) +
1427 \overleftrightarrow{\mathsf{W}}(t).
1428 \end{equation}
1429 The kinetic contribution to the pressure tensor utilizes the {\it
1430 outer} product of the velocities denoted by the $\otimes$ symbol. The
1431 stress tensor is calculated from another outer product of the
1432 inter-atomic separation vectors (${\bf r}_{ij} = {\bf r}_j - {\bf
1433 r}_i$) with the forces between the same two atoms,
1434 \begin{equation}
1435 \overleftrightarrow{\mathsf{W}}(t) = \sum_{i} \sum_{j>i} {\bf r}_{ij}(t)
1436 \otimes {\bf f}_{ij}(t).
1437 \end{equation}
1438 The instantaneous pressure is then simply obtained from the trace of
1439 the Pressure tensor,
1440 \begin{equation}
1441 P(t) = \frac{1}{3} \mathrm{Tr} \left( \overleftrightarrow{\mathsf{P}}(t).
1442 \right)
1443 \end{equation}
1444
1445 In eq.(\ref{eq:melchionna1}), $\tau_B$ is the time constant for
1446 relaxation of the pressure to the target value. To set values for
1447 $\tau_B$ or $P_{\mathrm{target}}$ in a simulation, one would use the
1448 {\tt tauBarostat} and {\tt targetPressure} keywords in the {\tt .bass}
1449 file. The units for {\tt tauBarostat} are fs, and the units for the
1450 {\tt targetPressure} are atmospheres. Like in the NVT integrator, the
1451 integration of the equations of motion is carried out in a
1452 velocity-Verlet style 2 part algorithm:
1453
1454 {\tt moveA:}
1455 \begin{align*}
1456 T(t) &\leftarrow \left\{{\bf v}(t)\right\}, \left\{{\bf j}(t)\right\} ,\\
1457 %
1458 P(t) &\leftarrow \left\{{\bf r}(t)\right\}, \left\{{\bf v}(t)\right\} ,\\
1459 %
1460 {\bf v}\left(t + h / 2\right) &\leftarrow {\bf v}(t)
1461 + \frac{h}{2} \left( \frac{{\bf f}(t)}{m} - {\bf v}(t)
1462 \left(\chi(t) + \eta(t) \right) \right), \\
1463 %
1464 {\bf j}\left(t + h / 2 \right) &\leftarrow {\bf j}(t)
1465 + \frac{h}{2} \left( {\bf \tau}^b(t) - {\bf j}(t)
1466 \chi(t) \right), \\
1467 %
1468 \mathsf{A}(t + h) &\leftarrow \mathrm{rotate}\left(h *
1469 {\bf j}(t + h / 2) \overleftrightarrow{\mathsf{I}}^{-1}
1470 \right) ,\\
1471 %
1472 \chi\left(t + h / 2 \right) &\leftarrow \chi(t) +
1473 \frac{h}{2 \tau_T^2} \left( \frac{T(t)}{T_{\mathrm{target}}} - 1
1474 \right) ,\\
1475 %
1476 \eta(t + h / 2) &\leftarrow \eta(t) + \frac{h
1477 \mathcal{V}(t)}{2 N k_B T(t) \tau_B^2} \left( P(t)
1478 - P_{\mathrm{target}} \right), \\
1479 %
1480 {\bf r}(t + h) &\leftarrow {\bf r}(t) + h
1481 \left\{ {\bf v}\left(t + h / 2 \right)
1482 + \eta(t + h / 2)\left[ {\bf r}(t + h)
1483 - {\bf R}_0 \right] \right\} ,\\
1484 %
1485 \mathsf{H}(t + h) &\leftarrow e^{-h \eta(t + h / 2)}
1486 \mathsf{H}(t).
1487 \end{align*}
1488
1489 Most of these equations are identical to their counterparts in the NVT
1490 integrator, but the propagation of positions to time $t + h$
1491 depends on the positions at the same time. {\sc oopse} carries out
1492 this step iteratively (with a limit of 5 passes through the iterative
1493 loop). Also, the simulation box $\mathsf{H}$ is scaled uniformly for
1494 one full time step by an exponential factor that depends on the value
1495 of $\eta$ at time $t +
1496 h / 2$. Reshaping the box uniformly also scales the volume of
1497 the box by
1498 \begin{equation}
1499 \mathcal{V}(t + h) \leftarrow e^{ - 3 h \eta(t + h /2)}.
1500 \mathcal{V}(t)
1501 \end{equation}
1502
1503 The {\tt doForces} step for the NPTi integrator is exactly the same as
1504 in both the DLM and NVT integrators. Once the forces and torques have
1505 been obtained at the new time step, the velocities can be advanced to
1506 the same time value.
1507
1508 {\tt moveB:}
1509 \begin{align*}
1510 T(t + h) &\leftarrow \left\{{\bf v}(t + h)\right\},
1511 \left\{{\bf j}(t + h)\right\} ,\\
1512 %
1513 P(t + h) &\leftarrow \left\{{\bf r}(t + h)\right\},
1514 \left\{{\bf v}(t + h)\right\}, \\
1515 %
1516 \chi\left(t + h \right) &\leftarrow \chi\left(t + h /
1517 2 \right) + \frac{h}{2 \tau_T^2} \left( \frac{T(t+h)}
1518 {T_{\mathrm{target}}} - 1 \right), \\
1519 %
1520 \eta(t + h) &\leftarrow \eta(t + h / 2) +
1521 \frac{h \mathcal{V}(t + h)}{2 N k_B T(t + h)
1522 \tau_B^2} \left( P(t + h) - P_{\mathrm{target}} \right), \\
1523 %
1524 {\bf v}\left(t + h \right) &\leftarrow {\bf v}\left(t
1525 + h / 2 \right) + \frac{h}{2} \left(
1526 \frac{{\bf f}(t + h)}{m} - {\bf v}(t + h)
1527 (\chi(t + h) + \eta(t + h)) \right) ,\\
1528 %
1529 {\bf j}\left(t + h \right) &\leftarrow {\bf j}\left(t
1530 + h / 2 \right) + \frac{h}{2} \left( {\bf
1531 \tau}^b(t + h) - {\bf j}(t + h)
1532 \chi(t + h) \right) .
1533 \end{align*}
1534
1535 Once again, since ${\bf v}(t + h)$ and ${\bf j}(t + h)$ are required
1536 to caclculate $T(t + h)$, $P(t + h)$, $\chi(t + h)$, and $\eta(t +
1537 h)$, they indirectly depend on their own values at time $t + h$. {\tt
1538 moveB} is therefore done in an iterative fashion until $\chi(t + h)$
1539 and $\eta(t + h)$ become self-consistent. The relative tolerance for
1540 the self-consistency check defaults to a value of $\mbox{10}^{-6}$,
1541 but {\sc oopse} will terminate the iteration after 4 loops even if the
1542 consistency check has not been satisfied.
1543
1544 The Melchionna modification of the Nos\'e-Hoover-Andersen algorithm is
1545 known to conserve a Hamiltonian for the extended system that is, to
1546 within a constant, identical to the Gibbs free energy,
1547 \begin{equation}
1548 H_{\mathrm{NPTi}} = V + K + f k_B T_{\mathrm{target}} \left(
1549 \frac{\tau_{T}^2 \chi^2(t)}{2} + \int_{0}^{t} \chi(t^\prime) dt^\prime
1550 \right) + P_{\mathrm{target}} \mathcal{V}(t).
1551 \end{equation}
1552 Poor choices of $\delta t$, $\tau_T$, or $\tau_B$ can result in
1553 non-conservation of $H_{\mathrm{NPTi}}$, so the conserved quantity is
1554 maintained in the last column of the {\tt .stat} file to allow checks
1555 on the quality of the integration. It is also known that this
1556 algorithm samples the equilibrium distribution for the enthalpy
1557 (including contributions for the thermostat and barostat),
1558 \begin{equation}
1559 H_{\mathrm{NPTi}} = V + K + \frac{f k_B T_{\mathrm{target}}}{2} \left(
1560 \chi^2 \tau_T^2 + \eta^2 \tau_B^2 \right) + P_{\mathrm{target}}
1561 \mathcal{V}(t).
1562 \end{equation}
1563
1564 Bond constraints are applied at the end of both the {\tt moveA} and
1565 {\tt moveB} portions of the algorithm. Details on the constraint
1566 algorithms are given in section \ref{oopseSec:rattle}.
1567
1568 \subsection{\label{sec:NPTf}Constant-pressure integration with a
1569 flexible box (NPTf)}
1570
1571 There is a relatively simple generalization of the
1572 Nos\'e-Hoover-Andersen method to include changes in the simulation box
1573 {\it shape} as well as in the volume of the box. This method utilizes
1574 the full $3 \times 3$ pressure tensor and introduces a tensor of
1575 extended variables ($\overleftrightarrow{\eta}$) to control changes to
1576 the box shape. The equations of motion for this method are
1577 \begin{eqnarray}
1578 \dot{{\bf r}} & = & {\bf v} + \overleftrightarrow{\eta} \cdot \left( {\bf r} - {\bf R}_0 \right), \\
1579 \dot{{\bf v}} & = & \frac{{\bf f}}{m} - (\overleftrightarrow{\eta} +
1580 \chi \cdot \mathsf{1}) {\bf v}, \\
1581 \dot{\mathsf{A}} & = & \mathsf{A} \cdot
1582 \mbox{ skew}\left(\overleftrightarrow{I}^{-1} \cdot {\bf j}\right) ,\\
1583 \dot{{\bf j}} & = & {\bf j} \times \left( \overleftrightarrow{I}^{-1}
1584 \cdot {\bf j} \right) - \mbox{ rot}\left(\mathsf{A}^{T} \cdot \frac{\partial
1585 V}{\partial \mathsf{A}} \right) - \chi {\bf j} ,\\
1586 \dot{\chi} & = & \frac{1}{\tau_{T}^2} \left(
1587 \frac{T}{T_{\mathrm{target}}} - 1 \right) ,\\
1588 \dot{\overleftrightarrow{\eta}} & = & \frac{1}{\tau_{B}^2 f k_B
1589 T_{\mathrm{target}}} V \left( \overleftrightarrow{\mathsf{P}} - P_{\mathrm{target}}\mathsf{1} \right) ,\\
1590 \dot{\mathsf{H}} & = & \overleftrightarrow{\eta} \cdot \mathsf{H} .
1591 \label{eq:melchionna2}
1592 \end{eqnarray}
1593
1594 Here, $\mathsf{1}$ is the unit matrix and $\overleftrightarrow{\mathsf{P}}$
1595 is the pressure tensor. Again, the volume, $\mathcal{V} = \det
1596 \mathsf{H}$.
1597
1598 The propagation of the equations of motion is nearly identical to the
1599 NPTi integration:
1600
1601 {\tt moveA:}
1602 \begin{align*}
1603 T(t) &\leftarrow \left\{{\bf v}(t)\right\}, \left\{{\bf j}(t)\right\} ,\\
1604 %
1605 \overleftrightarrow{\mathsf{P}}(t) &\leftarrow \left\{{\bf r}(t)\right\},
1606 \left\{{\bf v}(t)\right\} ,\\
1607 %
1608 {\bf v}\left(t + h / 2\right) &\leftarrow {\bf v}(t)
1609 + \frac{h}{2} \left( \frac{{\bf f}(t)}{m} -
1610 \left(\chi(t)\mathsf{1} + \overleftrightarrow{\eta}(t) \right) \cdot
1611 {\bf v}(t) \right), \\
1612 %
1613 {\bf j}\left(t + h / 2 \right) &\leftarrow {\bf j}(t)
1614 + \frac{h}{2} \left( {\bf \tau}^b(t) - {\bf j}(t)
1615 \chi(t) \right), \\
1616 %
1617 \mathsf{A}(t + h) &\leftarrow \mathrm{rotate}\left(h *
1618 {\bf j}(t + h / 2) \overleftrightarrow{\mathsf{I}}^{-1}
1619 \right), \\
1620 %
1621 \chi\left(t + h / 2 \right) &\leftarrow \chi(t) +
1622 \frac{h}{2 \tau_T^2} \left( \frac{T(t)}{T_{\mathrm{target}}}
1623 - 1 \right), \\
1624 %
1625 \overleftrightarrow{\eta}(t + h / 2) &\leftarrow
1626 \overleftrightarrow{\eta}(t) + \frac{h \mathcal{V}(t)}{2 N k_B
1627 T(t) \tau_B^2} \left( \overleftrightarrow{\mathsf{P}}(t)
1628 - P_{\mathrm{target}}\mathsf{1} \right), \\
1629 %
1630 {\bf r}(t + h) &\leftarrow {\bf r}(t) + h \left\{ {\bf v}
1631 \left(t + h / 2 \right) + \overleftrightarrow{\eta}(t +
1632 h / 2) \cdot \left[ {\bf r}(t + h)
1633 - {\bf R}_0 \right] \right\}, \\
1634 %
1635 \mathsf{H}(t + h) &\leftarrow \mathsf{H}(t) \cdot e^{-h
1636 \overleftrightarrow{\eta}(t + h / 2)} .
1637 \end{align*}
1638 {\sc oopse} uses a power series expansion truncated at second order
1639 for the exponential operation which scales the simulation box.
1640
1641 The {\tt moveB} portion of the algorithm is largely unchanged from the
1642 NPTi integrator:
1643
1644 {\tt moveB:}
1645 \begin{align*}
1646 T(t + h) &\leftarrow \left\{{\bf v}(t + h)\right\},
1647 \left\{{\bf j}(t + h)\right\}, \\
1648 %
1649 \overleftrightarrow{\mathsf{P}}(t + h) &\leftarrow \left\{{\bf r}
1650 (t + h)\right\}, \left\{{\bf v}(t
1651 + h)\right\}, \left\{{\bf f}(t + h)\right\} ,\\
1652 %
1653 \chi\left(t + h \right) &\leftarrow \chi\left(t + h /
1654 2 \right) + \frac{h}{2 \tau_T^2} \left( \frac{T(t+
1655 h)}{T_{\mathrm{target}}} - 1 \right), \\
1656 %
1657 \overleftrightarrow{\eta}(t + h) &\leftarrow
1658 \overleftrightarrow{\eta}(t + h / 2) +
1659 \frac{h \mathcal{V}(t + h)}{2 N k_B T(t + h)
1660 \tau_B^2} \left( \overleftrightarrow{P}(t + h)
1661 - P_{\mathrm{target}}\mathsf{1} \right) ,\\
1662 %
1663 {\bf v}\left(t + h \right) &\leftarrow {\bf v}\left(t
1664 + h / 2 \right) + \frac{h}{2} \left(
1665 \frac{{\bf f}(t + h)}{m} -
1666 (\chi(t + h)\mathsf{1} + \overleftrightarrow{\eta}(t
1667 + h)) \right) \cdot {\bf v}(t + h), \\
1668 %
1669 {\bf j}\left(t + h \right) &\leftarrow {\bf j}\left(t
1670 + h / 2 \right) + \frac{h}{2} \left( {\bf \tau}^b(t
1671 + h) - {\bf j}(t + h) \chi(t + h) \right) .
1672 \end{align*}
1673
1674 The iterative schemes for both {\tt moveA} and {\tt moveB} are
1675 identical to those described for the NPTi integrator.
1676
1677 The NPTf integrator is known to conserve the following Hamiltonian:
1678 \begin{equation}
1679 H_{\mathrm{NPTf}} = V + K + f k_B T_{\mathrm{target}} \left(
1680 \frac{\tau_{T}^2 \chi^2(t)}{2} + \int_{0}^{t} \chi(t^\prime) dt^\prime
1681 \right) + P_{\mathrm{target}} \mathcal{V}(t) + \frac{f k_B
1682 T_{\mathrm{target}}}{2}
1683 \mathrm{Tr}\left[\overleftrightarrow{\eta}(t)\right]^2 \tau_B^2.
1684 \end{equation}
1685
1686 This integrator must be used with care, particularly in liquid
1687 simulations. Liquids have very small restoring forces in the
1688 off-diagonal directions, and the simulation box can very quickly form
1689 elongated and sheared geometries which become smaller than the
1690 electrostatic or Lennard-Jones cutoff radii. The NPTf integrator
1691 finds most use in simulating crystals or liquid crystals which assume
1692 non-orthorhombic geometries.
1693
1694 \subsection{\label{nptxyz}Constant pressure in 3 axes (NPTxyz)}
1695
1696 There is one additional extended system integrator which is somewhat
1697 simpler than the NPTf method described above. In this case, the three
1698 axes have independent barostats which each attempt to preserve the
1699 target pressure along the box walls perpendicular to that particular
1700 axis. The lengths of the box axes are allowed to fluctuate
1701 independently, but the angle between the box axes does not change.
1702 The equations of motion are identical to those described above, but
1703 only the {\it diagonal} elements of $\overleftrightarrow{\eta}$ are
1704 computed. The off-diagonal elements are set to zero (even when the
1705 pressure tensor has non-zero off-diagonal elements).
1706
1707 It should be noted that the NPTxyz integrator is {\it not} known to
1708 preserve any Hamiltonian of interest to the chemical physics
1709 community. The integrator is extremely useful, however, in generating
1710 initial conditions for other integration methods. It {\it is} suitable
1711 for use with liquid simulations, or in cases where there is
1712 orientational anisotropy in the system (i.e. in lipid bilayer
1713 simulations).
1714
1715 \subsection{\label{sec:constraints}Constraint Methods}
1716
1717 \subsubsection{\label{oopseSec:rattle}The {\sc rattle} Method for Bond
1718 Constraints}
1719
1720 In order to satisfy the constraints of fixed bond lengths within {\sc
1721 oopse}, we have implemented the {\sc rattle} algorithm of
1722 Andersen.\cite{andersen83} The algorithm is a velocity verlet
1723 formulation of the {\sc shake} method\cite{ryckaert77} of iteratively
1724 solving the Lagrange multipliers of constraint.
1725
1726 \subsubsection{\label{oopseSec:zcons}Z-Constraint Method}
1727
1728 Based on the fluctuation-dissipation theorem, a force auto-correlation
1729 method was developed by Roux and Karplus to investigate the dynamics
1730 of ions inside ion channels.\cite{Roux91} The time-dependent friction
1731 coefficient can be calculated from the deviation of the instantaneous
1732 force from its mean force.
1733 \begin{equation}
1734 \xi(z,t)=\langle\delta F(z,t)\delta F(z,0)\rangle/k_{B}T,
1735 \end{equation}
1736 where%
1737 \begin{equation}
1738 \delta F(z,t)=F(z,t)-\langle F(z,t)\rangle.
1739 \end{equation}
1740
1741
1742 If the time-dependent friction decays rapidly, the static friction
1743 coefficient can be approximated by
1744 \begin{equation}
1745 \xi_{\text{static}}(z)=\int_{0}^{\infty}\langle\delta F(z,t)\delta F(z,0)\rangle dt.
1746 \end{equation}
1747 Allowing diffusion constant to then be calculated through the
1748 Einstein relation:\cite{Marrink94}
1749 \begin{equation}
1750 D(z)=\frac{k_{B}T}{\xi_{\text{static}}(z)}=\frac{(k_{B}T)^{2}}{\int_{0}^{\infty
1751 }\langle\delta F(z,t)\delta F(z,0)\rangle dt}.%
1752 \end{equation}
1753
1754 The Z-Constraint method, which fixes the z coordinates of the
1755 molecules with respect to the center of the mass of the system, has
1756 been a method suggested to obtain the forces required for the force
1757 auto-correlation calculation.\cite{Marrink94} However, simply resetting the
1758 coordinate will move the center of the mass of the whole system. To
1759 avoid this problem, a new method was used in {\sc oopse}. Instead of
1760 resetting the coordinate, we reset the forces of z-constrained
1761 molecules as well as subtract the total constraint forces from the
1762 rest of the system after the force calculation at each time step.
1763
1764 After the force calculation, define $G_\alpha$ as
1765 \begin{equation}
1766 G_{\alpha} = \sum_i F_{\alpha i},
1767 \label{oopseEq:zc1}
1768 \end{equation}
1769 where $F_{\alpha i}$ is the force in the z direction of atom $i$ in
1770 z-constrained molecule $\alpha$. The forces of the z constrained
1771 molecule are then set to:
1772 \begin{equation}
1773 F_{\alpha i} = F_{\alpha i} -
1774 \frac{m_{\alpha i} G_{\alpha}}{\sum_i m_{\alpha i}}.
1775 \end{equation}
1776 Here, $m_{\alpha i}$ is the mass of atom $i$ in the z-constrained
1777 molecule. Having rescaled the forces, the velocities must also be
1778 rescaled to subtract out any center of mass velocity in the z
1779 direction.
1780 \begin{equation}
1781 v_{\alpha i} = v_{\alpha i} -
1782 \frac{\sum_i m_{\alpha i} v_{\alpha i}}{\sum_i m_{\alpha i}},
1783 \end{equation}
1784 where $v_{\alpha i}$ is the velocity of atom $i$ in the z direction.
1785 Lastly, all of the accumulated z constrained forces must be subtracted
1786 from the system to keep the system center of mass from drifting.
1787 \begin{equation}
1788 F_{\beta i} = F_{\beta i} - \frac{m_{\beta i} \sum_{\alpha} G_{\alpha}}
1789 {\sum_{\beta}\sum_i m_{\beta i}},
1790 \end{equation}
1791 where $\beta$ are all of the unconstrained molecules in the
1792 system. Similarly, the velocities of the unconstrained molecules must
1793 also be scaled.
1794 \begin{equation}
1795 v_{\beta i} = v_{\beta i} + \sum_{\alpha}
1796 \frac{\sum_i m_{\alpha i} v_{\alpha i}}{\sum_i m_{\alpha i}}.
1797 \end{equation}
1798
1799 At the very beginning of the simulation, the molecules may not be at their
1800 constrained positions. To move a z-constrained molecule to its specified
1801 position, a simple harmonic potential is used
1802 \begin{equation}
1803 U(t)=\frac{1}{2}k_{\text{Harmonic}}(z(t)-z_{\text{cons}})^{2},%
1804 \end{equation}
1805 where $k_{\text{Harmonic}}$ is the harmonic force constant, $z(t)$ is the
1806 current $z$ coordinate of the center of mass of the constrained molecule, and
1807 $z_{\text{cons}}$ is the constrained position. The harmonic force operating
1808 on the z-constrained molecule at time $t$ can be calculated by
1809 \begin{equation}
1810 F_{z_{\text{Harmonic}}}(t)=-\frac{\partial U(t)}{\partial z(t)}=
1811 -k_{\text{Harmonic}}(z(t)-z_{\text{cons}}).
1812 \end{equation}
1813
1814 \section{\label{oopseSec:design}Program Design}
1815
1816 \subsection{\label{sec:architecture} {\sc oopse} Architecture}
1817
1818 The core of OOPSE is divided into two main object libraries:
1819 \texttt{libBASS} and \texttt{libmdtools}. \texttt{libBASS} is the
1820 library developed around the parsing engine and \texttt{libmdtools}
1821 is the software library developed around the simulation engine. These
1822 two libraries are designed to encompass all the basic functions and
1823 tools that {\sc oopse} provides. Utility programs, such as the
1824 property analyzers, need only link against the software libraries to
1825 gain access to parsing, force evaluation, and input / output
1826 routines.
1827
1828 Contained in \texttt{libBASS} are all the routines associated with
1829 reading and parsing the \texttt{.bass} input files. Given a
1830 \texttt{.bass} file, \texttt{libBASS} will open it and any associated
1831 \texttt{.mdl} files; then create structures in memory that are
1832 templates of all the molecules specified in the input files. In
1833 addition, any simulation parameters set in the \texttt{.bass} file
1834 will be placed in a structure for later query by the controlling
1835 program.
1836
1837 Located in \texttt{libmdtools} are all other routines necessary to a
1838 Molecular Dynamics simulation. The library uses the main data
1839 structures returned by \texttt{libBASS} to initialize the various
1840 parts of the simulation: the atom structures and positions, the force
1841 field, the integrator, \emph{et cetera}. After initialization, the
1842 library can be used to perform a variety of tasks: integrate a
1843 Molecular Dynamics trajectory, query phase space information from a
1844 specific frame of a completed trajectory, or even recalculate force or
1845 energetic information about specific frames from a completed
1846 trajectory.
1847
1848 With these core libraries in place, several programs have been
1849 developed to utilize the routines provided by \texttt{libBASS} and
1850 \texttt{libmdtools}. The main program of the package is \texttt{oopse}
1851 and the corresponding parallel version \texttt{oopse\_MPI}. These two
1852 programs will take the \texttt{.bass} file, and create (and integrate)
1853 the simulation specified in the script. The two analysis programs
1854 \texttt{staticProps} and \texttt{dynamicProps} utilize the core
1855 libraries to initialize and read in trajectories from previously
1856 completed simulations, in addition to the ability to use functionality
1857 from \texttt{libmdtools} to recalculate forces and energies at key
1858 frames in the trajectories. Lastly, the family of system building
1859 programs (Sec.~\ref{oopseSec:initCoords}) also use the libraries to
1860 store and output the system configurations they create.
1861
1862 \subsection{\label{oopseSec:parallelization} Parallelization of {\sc oopse}}
1863
1864 Although processor power is continually growing roughly following
1865 Moore's Law, it is still unreasonable to simulate systems of more then
1866 a 1000 atoms on a single processor. To facilitate study of larger
1867 system sizes or smaller systems on long time scales in a reasonable
1868 period of time, parallel methods were developed allowing multiple
1869 CPU's to share the simulation workload. Three general categories of
1870 parallel decomposition methods have been developed including atomic,
1871 spatial and force decomposition methods.
1872
1873 Algorithmically simplest of the three methods is atomic decomposition
1874 where N particles in a simulation are split among P processors for the
1875 duration of the simulation. Computational cost scales as an optimal
1876 $\mathcal{O}(N/P)$ for atomic decomposition. Unfortunately all
1877 processors must communicate positions and forces with all other
1878 processors at every force evaluation, leading communication costs to
1879 scale as an unfavorable $\mathcal{O}(N)$, \emph{independent of the
1880 number of processors}. This communication bottleneck led to the
1881 development of spatial and force decomposition methods in which
1882 communication among processors scales much more favorably. Spatial or
1883 domain decomposition divides the physical spatial domain into 3D boxes
1884 in which each processor is responsible for calculation of forces and
1885 positions of particles located in its box. Particles are reassigned to
1886 different processors as they move through simulation space. To
1887 calculate forces on a given particle, a processor must know the
1888 positions of particles within some cutoff radius located on nearby
1889 processors instead of the positions of particles on all
1890 processors. Both communication between processors and computation
1891 scale as $\mathcal{O}(N/P)$ in the spatial method. However, spatial
1892 decomposition adds algorithmic complexity to the simulation code and
1893 is not very efficient for small N since the overall communication
1894 scales as the surface to volume ratio $\mathcal{O}(N/P)^{2/3}$ in
1895 three dimensions.
1896
1897 The parallelization method used in {\sc oopse} is the force
1898 decomposition method. Force decomposition assigns particles to
1899 processors based on a block decomposition of the force
1900 matrix. Processors are split into an optimally square grid forming row
1901 and column processor groups. Forces are calculated on particles in a
1902 given row by particles located in that processors column
1903 assignment. Force decomposition is less complex to implement than the
1904 spatial method but still scales computationally as $\mathcal{O}(N/P)$
1905 and scales as $\mathcal{O}(N/\sqrt{P})$ in communication
1906 cost. Plimpton has also found that force decompositions scale more
1907 favorably than spatial decompositions for systems up to 10,000 atoms
1908 and favorably compete with spatial methods up to 100,000
1909 atoms.\cite{plimpton95}
1910
1911 \section{\label{oopseSec:conclusion}Conclusion}
1912
1913 We have presented the design and implementation of our open source
1914 simulation package {\sc oopse}. The package offers novel capabilities
1915 to the field of Molecular Dynamics simulation packages in the form of
1916 dipolar force fields, and symplectic integration of rigid body
1917 dynamics. It is capable of scaling across multiple processors through
1918 the use of force based decomposition using MPI. It also implements
1919 several advanced integrators allowing the end user control over
1920 temperature and pressure. In addition, it is capable of integrating
1921 constrained dynamics through both the {\sc rattle} algorithm and the
1922 z-constraint method.
1923
1924 These features are all brought together in a single open-source
1925 program. This allows researchers to not only benefit from
1926 {\sc oopse}, but also contribute to {\sc oopse}'s development as
1927 well.
1928
1929
1930 \newpage
1931 \section{Acknowledgments}
1932 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}.
1933
1934 \bibliographystyle{achemso}
1935 \bibliography{oopsePaper}
1936
1937 \end{document}