ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/mattDisertation/oopse.tex
Revision: 1071
Committed: Thu Feb 26 21:55:43 2004 UTC (20 years, 4 months ago) by mmeineke
Content type: application/x-tex
File size: 66434 byte(s)
Log Message:
added citations to the bib, and started on a honking huge section of RATTLE in the oopse chapter.

File Contents

# Content
1 \chapter{\label{chapt:oopse}OOPSE: AN OPEN SOURCE OBJECT-ORIENTED PARALLEL SIMULATION ENGINE FOR MOLECULAR DYNAMICS}
2
3
4
5 %% \begin{abstract}
6 %% We detail the capabilities of a new open-source parallel simulation
7 %% package ({\sc oopse}) that can perform molecular dynamics simulations
8 %% on atom types that are missing from other popular packages. In
9 %% particular, {\sc oopse} is capable of performing orientational
10 %% dynamics on dipolar systems, and it can handle simulations of metallic
11 %% systems using the embedded atom method ({\sc eam}).
12 %% \end{abstract}
13
14 \lstset{language=C,frame=TB,basicstyle=\small,basicstyle=\ttfamily, %
15 xleftmargin=0.5in, xrightmargin=0.5in,captionpos=b, %
16 abovecaptionskip=0.5cm, belowcaptionskip=0.5cm}
17
18 \section{\label{oopseSec:foreword}Foreword}
19
20 In this chapter, I present and detail the capabilities of the open
21 source simulation package {\sc oopse}. It is important to note, that a
22 simulation package of this size and scope would not have been possible
23 without the collaborative efforts of my colleagues: Charles
24 F.~Vardeman II, Teng Lin, Christopher J.~Fennell and J.~Daniel
25 Gezelter. Although my contributions to {\sc oopse} are major,
26 consideration of my work apart from the others would not give a
27 complete description to the package's capabilities. As such, all
28 contributions to {\sc oopse} to date are presented in this chapter.
29
30 Charles Vardeman is responsible for the parallelization of the long
31 range forces in {\sc oopse} (Sec.~\ref{oopseSec:parallelization}) as
32 well as the inclusion of the embedded-atom potential for transition
33 metals (Sec.~\ref{oopseSec:eam}). Teng Lin's contributions include
34 refinement of the periodic boundary conditions
35 (Sec.~\ref{oopseSec:pbc}), the z-constraint method
36 (Sec.~\ref{oopseSec:zcons}), refinement of the property analysis
37 programs (Sec.~\ref{oopseSec:props}), and development in the extended
38 system integrators (Sec.~\ref{oopseSec:noseHooverThermo}). Christopher
39 Fennell worked on the symplectic integrator
40 (Sec.~\ref{oopseSec:integrate}) and the refinement of the {\sc ssd}
41 water model (Sec.~\ref{oopseSec:SSD}). Daniel Gezelter lent his
42 talents in the development of the extended system integrators
43 (Sec.~\ref{oopseSec:noseHooverThermo}) as well as giving general
44 direction and oversight to the entire project. My responsibilities
45 covered the creation and specification of {\sc bass}
46 (Sec.~\ref{oopseSec:IOfiles}), the original development of the single
47 processor version of {\sc oopse}, contributions to the extended state
48 integrators (Sec.~\ref{oopseSec:noseHooverThermo}), the implementation
49 of the Lennard-Jones (Sec.~\ref{sec:LJPot}) and {\sc duff}
50 (Sec.~\ref{oopseSec:DUFF}) force fields, and initial implementation of
51 the property analysis (Sec.~\ref{oopseSec:props}) and system
52 initialization (Sec.~\ref{oopseSec:initCoords}) utility programs. {\sc
53 oopse}, like many other Molecular Dynamics programs, is a work in
54 progress, and will continue to be so for many graduate student
55 lifetimes.
56
57 \section{\label{sec:intro}Introduction}
58
59 When choosing to simulate a chemical system with molecular dynamics,
60 there are a variety of options available. For simple systems, one
61 might consider writing one's own programming code. However, as systems
62 grow larger and more complex, building and maintaining code for the
63 simulations becomes a time consuming task. In such cases it is usually
64 more convenient for a researcher to turn to pre-existing simulation
65 packages. These packages, such as {\sc amber}\cite{pearlman:1995} and
66 {\sc charmm}\cite{Brooks83}, provide powerful tools for researchers to
67 conduct simulations of their systems without spending their time
68 developing a code base to conduct their research. This then frees them
69 to perhaps explore experimental analogues to their models.
70
71 Despite their utility, problems with these packages arise when
72 researchers try to develop techniques or energetic models that the
73 code was not originally designed to simulate. Examples of uncommonly
74 implemented techniques and energetics include; dipole-dipole
75 interactions, rigid body dynamics, and metallic embedded
76 potentials. When faced with these obstacles, a researcher must either
77 develop their own code or license and extend one of the commercial
78 packages. What we have elected to do, is develop a package of
79 simulation code capable of implementing the types of models upon which
80 our research is based.
81
82 In developing {\sc oopse}, we have adhered to the precepts of Open
83 Source development, and are releasing our source code with a
84 permissive license. It is our intent that by doing so, other
85 researchers might benefit from our work, and add their own
86 contributions to the package. The license under which {\sc oopse} is
87 distributed allows any researcher to download and modify the source
88 code for their own use. In this way further development of {\sc oopse}
89 is not limited to only the models of interest to ourselves, but also
90 those of the community of scientists who contribute back to the
91 project.
92
93 We have structured this chapter to first discuss the empirical energy
94 functions that {\sc oopse } implements in
95 Sec.~\ref{oopseSec:empiricalEnergy}. Following that is a discussion of
96 the various input and output files associated with the package
97 (Sec.~\ref{oopseSec:IOfiles}). Sec.~\ref{oopseSec:mechanics}
98 elucidates the various Molecular Dynamics algorithms {\sc oopse}
99 implements in the integration of the Newtonian equations of
100 motion. Basic analysis of the trajectories obtained from the
101 simulation is discussed in Sec.~\ref{oopseSec:props}. Program design
102 considerations are presented in Sec.~\ref{oopseSec:design}. And
103 lastly, Sec.~\ref{oopseSec:conclusion} concludes the chapter.
104
105 \section{\label{oopseSec:empiricalEnergy}The Empirical Energy Functions}
106
107 \subsection{\label{oopseSec:atomsMolecules}Atoms, Molecules and Rigid Bodies}
108
109 The basic unit of an {\sc oopse} simulation is the atom. The
110 parameters describing the atom are generalized to make the atom as
111 flexible a representation as possible. They may represent specific
112 atoms of an element, or be used for collections of atoms such as
113 methyl and carbonyl groups. The atoms are also capable of having
114 directional components associated with them (\emph{e.g.}~permanent
115 dipoles). Charges, permanent dipoles, and Lennard-Jones parameters for
116 a given atom type are set in the force field parameter files.
117
118 \begin{lstlisting}[float,caption={[Specifier for molecules and atoms] A sample specification of an Ar molecule},label=sch:AtmMole]
119 molecule{
120 name = "Ar";
121 nAtoms = 1;
122 atom[0]{
123 type="Ar";
124 position( 0.0, 0.0, 0.0 );
125 }
126 }
127 \end{lstlisting}
128
129
130 Atoms can be collected into secondary structures such as rigid bodies
131 or molecules. The molecule is a way for {\sc oopse} to keep track of
132 the atoms in a simulation in logical manner. Molecular units store the
133 identities of all the atoms and rigid bodies associated with
134 themselves, and are responsible for the evaluation of their own
135 internal interactions (\emph{i.e.}~bonds, bends, and torsions). Scheme
136 \ref{sch:AtmMole} shows how one creates a molecule in a ``model'' or
137 \texttt{.mdl} file. The position of the atoms given in the
138 declaration are relative to the origin of the molecule, and is used
139 when creating a system containing the molecule.
140
141 As stated previously, one of the features that sets {\sc oopse} apart
142 from most of the current molecular simulation packages is the ability
143 to handle rigid body dynamics. Rigid bodies are non-spherical
144 particles or collections of particles that have a constant internal
145 potential and move collectively.\cite{Goldstein01} They are not
146 included in most simulation packages because of the algorithmic
147 complexity involved in propagating orientational degrees of
148 freedom. Until recently, integrators which propagate orientational
149 motion have been much worse than those available for translational
150 motion.
151
152 Moving a rigid body involves determination of both the force and
153 torque applied by the surroundings, which directly affect the
154 translational and rotational motion in turn. In order to accumulate
155 the total force on a rigid body, the external forces and torques must
156 first be calculated for all the internal particles. The total force on
157 the rigid body is simply the sum of these external forces.
158 Accumulation of the total torque on the rigid body is more complex
159 than the force because the torque is applied to the center of mass of
160 the rigid body. The torque on rigid body $i$ is
161 \begin{equation}
162 \boldsymbol{\tau}_i=
163 \sum_{a}\biggl[(\mathbf{r}_{ia}-\mathbf{r}_i)\times \mathbf{f}_{ia}
164 + \boldsymbol{\tau}_{ia}\biggr]
165 \label{eq:torqueAccumulate}
166 \end{equation}
167 where $\boldsymbol{\tau}_i$ and $\mathbf{r}_i$ are the torque on and
168 position of the center of mass respectively, while $\mathbf{f}_{ia}$,
169 $\mathbf{r}_{ia}$, and $\boldsymbol{\tau}_{ia}$ are the force on,
170 position of, and torque on the component particles of the rigid body.
171
172 The summation of the total torque is done in the body fixed axis of
173 each rigid body. In order to move between the space fixed and body
174 fixed coordinate axes, parameters describing the orientation must be
175 maintained for each rigid body. At a minimum, the rotation matrix
176 (\textbf{A}) can be described by the three Euler angles ($\phi,
177 \theta,$ and $\psi$), where the elements of \textbf{A} are composed of
178 trigonometric operations involving $\phi, \theta,$ and
179 $\psi$.\cite{Goldstein01} In order to avoid numerical instabilities
180 inherent in using the Euler angles, the four parameter ``quaternion''
181 scheme is often used. The elements of \textbf{A} can be expressed as
182 arithmetic operations involving the four quaternions ($q_0, q_1, q_2,$
183 and $q_3$).\cite{allen87:csl} Use of quaternions also leads to
184 performance enhancements, particularly for very small
185 systems.\cite{Evans77}
186
187 {\sc oopse} utilizes a relatively new scheme that propagates the
188 entire nine parameter rotation matrix. Further discussion
189 on this choice can be found in Sec.~\ref{oopseSec:integrate}. An example
190 definition of a rigid body can be seen in Scheme
191 \ref{sch:rigidBody}. The positions in the atom definitions are the
192 placements of the atoms relative to the origin of the rigid body,
193 which itself has a position relative to the origin of the molecule.
194
195 \begin{lstlisting}[float,caption={[Defining rigid bodies]A sample definition of a rigid body},label={sch:rigidBody}]
196 molecule{
197 name = "TIP3P_water";
198 nRigidBodies = 1;
199 rigidBody[0]{
200 nAtoms = 3;
201 atom[0]{
202 type = "O_TIP3P";
203 position( 0.0, 0.0, -0.06556 );
204 }
205 atom[1]{
206 type = "H_TIP3P";
207 position( 0.0, 0.75695, 0.52032 );
208 }
209 atom[2]{
210 type = "H_TIP3P";
211 position( 0.0, -0.75695, 0.52032 );
212 }
213 position( 0.0, 0.0, 0.0 );
214 orientation( 0.0, 0.0, 1.0 );
215 }
216 }
217 \end{lstlisting}
218
219 \subsection{\label{sec:LJPot}The Lennard Jones Force Field}
220
221 The most basic force field implemented in {\sc oopse} is the
222 Lennard-Jones force field, which mimics the van der Waals interaction at
223 long distances, and uses an empirical repulsion at short
224 distances. The Lennard-Jones potential is given by:
225 \begin{equation}
226 V_{\text{LJ}}(r_{ij}) =
227 4\epsilon_{ij} \biggl[
228 \biggl(\frac{\sigma_{ij}}{r_{ij}}\biggr)^{12}
229 - \biggl(\frac{\sigma_{ij}}{r_{ij}}\biggr)^{6}
230 \biggr]
231 \label{eq:lennardJonesPot}
232 \end{equation}
233 Where $r_{ij}$ is the distance between particles $i$ and $j$,
234 $\sigma_{ij}$ scales the length of the interaction, and
235 $\epsilon_{ij}$ scales the well depth of the potential. Scheme
236 \ref{sch:LJFF} gives and example \texttt{.bass} file that
237 sets up a system of 108 Ar particles to be simulated using the
238 Lennard-Jones force field.
239
240 \begin{lstlisting}[float,caption={[Invocation of the Lennard-Jones force field] A sample system using the Lennard-Jones force field.},label={sch:LJFF}]
241
242 #include "argon.mdl"
243
244 nComponents = 1;
245 component{
246 type = "Ar";
247 nMol = 108;
248 }
249
250 initialConfig = "./argon.init";
251
252 forceField = "LJ";
253 \end{lstlisting}
254
255 Because this potential is calculated between all pairs, the force
256 evaluation can become computationally expensive for large systems. To
257 keep the pair evaluations to a manageable number, {\sc oopse} employs
258 a cut-off radius.\cite{allen87:csl} The cutoff radius can either be
259 specified in the \texttt{.bass} file, or left as its default value of
260 $2.5\sigma_{ii}$, where $\sigma_{ii}$ is the largest Lennard-Jones
261 length parameter present in the simulation. Truncating the calculation
262 at $r_{\text{cut}}$ introduces a discontinuity into the potential
263 energy and the force. To offset this discontinuity in the potential,
264 the energy value at $r_{\text{cut}}$ is subtracted from the
265 potential. This causes the potential to go to zero smoothly at the
266 cut-off radius, and preserves conservation of energy in integrating
267 the equations of motion.
268
269 Interactions between dissimilar particles requires the generation of
270 cross term parameters for $\sigma$ and $\epsilon$. These are
271 calculated through the Lorentz-Berthelot mixing
272 rules:\cite{allen87:csl}
273 \begin{equation}
274 \sigma_{ij} = \frac{1}{2}[\sigma_{ii} + \sigma_{jj}]
275 \label{eq:sigmaMix}
276 \end{equation}
277 and
278 \begin{equation}
279 \epsilon_{ij} = \sqrt{\epsilon_{ii} \epsilon_{jj}}
280 \label{eq:epsilonMix}
281 \end{equation}
282
283 \subsection{\label{oopseSec:DUFF}Dipolar Unified-Atom Force Field}
284
285 The dipolar unified-atom force field ({\sc duff}) was developed to
286 simulate lipid bilayers. The simulations require a model capable of
287 forming bilayers, while still being sufficiently computationally
288 efficient to allow large systems ($\sim$100's of phospholipids,
289 $\sim$1000's of waters) to be simulated for long times
290 ($\sim$10's of nanoseconds).
291
292 With this goal in mind, {\sc duff} has no point
293 charges. Charge-neutral distributions were replaced with dipoles,
294 while most atoms and groups of atoms were reduced to Lennard-Jones
295 interaction sites. This simplification cuts the length scale of long
296 range interactions from $\frac{1}{r}$ to $\frac{1}{r^3}$, and allows
297 us to avoid the computationally expensive Ewald sum. Instead, we can
298 use neighbor-lists and cutoff radii for the dipolar interactions, or
299 include a reaction field to mimic larger range interactions.
300
301 As an example, lipid head-groups in {\sc duff} are represented as
302 point dipole interaction sites. By placing a dipole at the head group
303 center of mass, our model mimics the charge separation found in common
304 phospholipids such as phosphatidylcholine.\cite{Cevc87} Additionally,
305 a large Lennard-Jones site is located at the pseudoatom's center of
306 mass. The model is illustrated by the red atom in
307 Fig.~\ref{oopseFig:lipidModel}. The water model we use to complement
308 the dipoles of the lipids is our reparameterization of the soft sticky
309 dipole (SSD) model of Ichiye
310 \emph{et al.}\cite{liu96:new_model}
311
312 \begin{figure}
313 \centering
314 \includegraphics[width=\linewidth]{lipidModel.eps}
315 \caption{A representation of the lipid model. $\phi$ is the torsion angle, $\theta$ %
316 is the bend angle, $\mu$ is the dipole moment of the head group, and n
317 is the chain length.}
318 \label{oopseFig:lipidModel}
319 \end{figure}
320
321 We have used a set of scalable parameters to model the alkyl groups
322 with Lennard-Jones sites. For this, we have borrowed parameters from
323 the TraPPE force field of Siepmann
324 \emph{et al}.\cite{Siepmann1998} TraPPE is a unified-atom
325 representation of n-alkanes, which is parametrized against phase
326 equilibria using Gibbs ensemble Monte Carlo simulation
327 techniques.\cite{Siepmann1998} One of the advantages of TraPPE is that
328 it generalizes the types of atoms in an alkyl chain to keep the number
329 of pseudoatoms to a minimum; the parameters for a unified atom such as
330 $\text{CH}_2$ do not change depending on what species are bonded to
331 it.
332
333 TraPPE also constrains all bonds to be of fixed length. Typically,
334 bond vibrations are the fastest motions in a molecular dynamic
335 simulation. Small time steps between force evaluations must be used to
336 ensure adequate energy conservation in the bond degrees of freedom. By
337 constraining the bond lengths, larger time steps may be used when
338 integrating the equations of motion. A simulation using {\sc duff} is
339 illustrated in Scheme \ref{sch:DUFF}.
340
341 \begin{lstlisting}[float,caption={[Invocation of {\sc duff}]Sample \texttt{.bass} file showing a simulation utilizing {\sc duff}},label={sch:DUFF}]
342
343 #include "water.mdl"
344 #include "lipid.mdl"
345
346 nComponents = 2;
347 component{
348 type = "simpleLipid_16";
349 nMol = 60;
350 }
351
352 component{
353 type = "SSD_water";
354 nMol = 1936;
355 }
356
357 initialConfig = "bilayer.init";
358
359 forceField = "DUFF";
360
361 \end{lstlisting}
362
363 \subsection{\label{oopseSec:energyFunctions}{\sc duff} Energy Functions}
364
365 The total potential energy function in {\sc duff} is
366 \begin{equation}
367 V = \sum^{N}_{I=1} V^{I}_{\text{Internal}}
368 + \sum^{N-1}_{I=1} \sum_{J>I} V^{IJ}_{\text{Cross}}
369 \label{eq:totalPotential}
370 \end{equation}
371 Where $V^{I}_{\text{Internal}}$ is the internal potential of molecule $I$:
372 \begin{equation}
373 V^{I}_{\text{Internal}} =
374 \sum_{\theta_{ijk} \in I} V_{\text{bend}}(\theta_{ijk})
375 + \sum_{\phi_{ijkl} \in I} V_{\text{torsion}}(\phi_{ijkl})
376 + \sum_{i \in I} \sum_{(j>i+4) \in I}
377 \biggl[ V_{\text{LJ}}(r_{ij}) + V_{\text{dipole}}
378 (\mathbf{r}_{ij},\boldsymbol{\Omega}_{i},\boldsymbol{\Omega}_{j})
379 \biggr]
380 \label{eq:internalPotential}
381 \end{equation}
382 Here $V_{\text{bend}}$ is the bend potential for all 1, 3 bonded pairs
383 within the molecule $I$, and $V_{\text{torsion}}$ is the torsion potential
384 for all 1, 4 bonded pairs. The pairwise portions of the internal
385 potential are excluded for pairs that are closer than three bonds,
386 i.e.~atom pairs farther away than a torsion are included in the
387 pair-wise loop.
388
389
390 The bend potential of a molecule is represented by the following function:
391 \begin{equation}
392 V_{\text{bend}}(\theta_{ijk}) = k_{\theta}( \theta_{ijk} - \theta_0 )^2 \label{eq:bendPot}
393 \end{equation}
394 Where $\theta_{ijk}$ is the angle defined by atoms $i$, $j$, and $k$
395 (see Fig.~\ref{oopseFig:lipidModel}), $\theta_0$ is the equilibrium
396 bond angle, and $k_{\theta}$ is the force constant which determines the
397 strength of the harmonic bend. The parameters for $k_{\theta}$ and
398 $\theta_0$ are borrowed from those in TraPPE.\cite{Siepmann1998}
399
400 The torsion potential and parameters are also borrowed from TraPPE. It is
401 of the form:
402 \begin{equation}
403 V_{\text{torsion}}(\phi) = c_1[1 + \cos \phi]
404 + c_2[1 + \cos(2\phi)]
405 + c_3[1 + \cos(3\phi)]
406 \label{eq:origTorsionPot}
407 \end{equation}
408 Where:
409 \begin{equation}
410 \cos\phi = (\hat{\mathbf{r}}_{ij} \times \hat{\mathbf{r}}_{jk}) \cdot
411 (\hat{\mathbf{r}}_{jk} \times \hat{\mathbf{r}}_{kl})
412 \label{eq:torsPhi}
413 \end{equation}
414 Here, $\hat{\mathbf{r}}_{\alpha\beta}$ are the set of unit bond
415 vectors between atoms $i$, $j$, $k$, and $l$. For computational
416 efficiency, the torsion potential has been recast after the method of
417 {\sc charmm},\cite{Brooks83} in which the angle series is converted to
418 a power series of the form:
419 \begin{equation}
420 V_{\text{torsion}}(\phi) =
421 k_3 \cos^3 \phi + k_2 \cos^2 \phi + k_1 \cos \phi + k_0
422 \label{eq:torsionPot}
423 \end{equation}
424 Where:
425 \begin{align*}
426 k_0 &= c_1 + c_3 \\
427 k_1 &= c_1 - 3c_3 \\
428 k_2 &= 2 c_2 \\
429 k_3 &= 4c_3
430 \end{align*}
431 By recasting the potential as a power series, repeated trigonometric
432 evaluations are avoided during the calculation of the potential energy.
433
434
435 The cross potential between molecules $I$ and $J$, $V^{IJ}_{\text{Cross}}$, is
436 as follows:
437 \begin{equation}
438 V^{IJ}_{\text{Cross}} =
439 \sum_{i \in I} \sum_{j \in J}
440 \biggl[ V_{\text{LJ}}(r_{ij}) + V_{\text{dipole}}
441 (\mathbf{r}_{ij},\boldsymbol{\Omega}_{i},\boldsymbol{\Omega}_{j})
442 + V_{\text{sticky}}
443 (\mathbf{r}_{ij},\boldsymbol{\Omega}_{i},\boldsymbol{\Omega}_{j})
444 \biggr]
445 \label{eq:crossPotentail}
446 \end{equation}
447 Where $V_{\text{LJ}}$ is the Lennard Jones potential,
448 $V_{\text{dipole}}$ is the dipole dipole potential, and
449 $V_{\text{sticky}}$ is the sticky potential defined by the SSD model
450 (Sec.~\ref{oopseSec:SSD}). Note that not all atom types include all
451 interactions.
452
453 The dipole-dipole potential has the following form:
454 \begin{equation}
455 V_{\text{dipole}}(\mathbf{r}_{ij},\boldsymbol{\Omega}_{i},
456 \boldsymbol{\Omega}_{j}) = \frac{|\mu_i||\mu_j|}{4\pi\epsilon_{0}r_{ij}^{3}} \biggl[
457 \boldsymbol{\hat{u}}_{i} \cdot \boldsymbol{\hat{u}}_{j}
458 -
459 3(\boldsymbol{\hat{u}}_i \cdot \hat{\mathbf{r}}_{ij}) %
460 (\boldsymbol{\hat{u}}_j \cdot \hat{\mathbf{r}}_{ij}) \biggr]
461 \label{eq:dipolePot}
462 \end{equation}
463 Here $\mathbf{r}_{ij}$ is the vector starting at atom $i$ pointing
464 towards $j$, and $\boldsymbol{\Omega}_i$ and $\boldsymbol{\Omega}_j$
465 are the orientational degrees of freedom for atoms $i$ and $j$
466 respectively. $|\mu_i|$ is the magnitude of the dipole moment of atom
467 $i$, $\boldsymbol{\hat{u}}_i$ is the standard unit orientation vector
468 of $\boldsymbol{\Omega}_i$, and $\boldsymbol{\hat{r}}_{ij}$ is the
469 unit vector pointing along $\mathbf{r}_{ij}$
470 ($\boldsymbol{\hat{r}}_{ij}=\mathbf{r}_{ij}/|\mathbf{r}_{ij}|$).
471
472 To improve computational efficiency of the dipole-dipole interactions,
473 {\sc oopse} employs an electrostatic cutoff radius. This parameter can
474 be set in the \texttt{.bass} file, and controls the length scale over
475 which dipole interactions are felt. To compensate for the
476 discontinuity in the potential and the forces at the cutoff radius, we
477 have implemented a switching function to smoothly scale the
478 dipole-dipole interaction at the cutoff.
479 \begin{equation}
480 S(r_{ij}) =
481 \begin{cases}
482 1 & \text{if $r_{ij} \le r_t$},\\
483 \frac{(r_{\text{cut}} + 2r_{ij} - 3r_t)(r_{\text{cut}} - r_{ij})^2}
484 {(r_{\text{cut}} - r_t)^2}
485 & \text{if $r_t < r_{ij} \le r_{\text{cut}}$}, \\
486 0 & \text{if $r_{ij} > r_{\text{cut}}$.}
487 \end{cases}
488 \label{eq:dipoleSwitching}
489 \end{equation}
490 Here $S(r_{ij})$ scales the potential at a given $r_{ij}$, and $r_t$
491 is the taper radius some given thickness less than the electrostatic
492 cutoff. The switching thickness can be set in the \texttt{.bass} file.
493
494 \subsection{\label{oopseSec:SSD}The {\sc duff} Water Models: SSD/E and SSD/RF}
495
496 In the interest of computational efficiency, the default solvent used
497 by {\sc oopse} is the extended Soft Sticky Dipole (SSD/E) water
498 model.\cite{Gezelter04} The original SSD was developed by Ichiye
499 \emph{et al.}\cite{liu96:new_model} as a modified form of the hard-sphere
500 water model proposed by Bratko, Blum, and
501 Luzar.\cite{Bratko85,Bratko95} It consists of a single point dipole
502 with a Lennard-Jones core and a sticky potential that directs the
503 particles to assume the proper hydrogen bond orientation in the first
504 solvation shell. Thus, the interaction between two SSD water molecules
505 \emph{i} and \emph{j} is given by the potential
506 \begin{equation}
507 V_{ij} =
508 V_{ij}^{LJ} (r_{ij})\ + V_{ij}^{dp}
509 (\mathbf{r}_{ij},\boldsymbol{\Omega}_i,\boldsymbol{\Omega}_j)\ +
510 V_{ij}^{sp}
511 (\mathbf{r}_{ij},\boldsymbol{\Omega}_i,\boldsymbol{\Omega}_j),
512 \label{eq:ssdPot}
513 \end{equation}
514 where the $\mathbf{r}_{ij}$ is the position vector between molecules
515 \emph{i} and \emph{j} with magnitude equal to the distance $r_{ij}$, and
516 $\boldsymbol{\Omega}_i$ and $\boldsymbol{\Omega}_j$ represent the
517 orientations of the respective molecules. The Lennard-Jones and dipole
518 parts of the potential are given by equations \ref{eq:lennardJonesPot}
519 and \ref{eq:dipolePot} respectively. The sticky part is described by
520 the following,
521 \begin{equation}
522 u_{ij}^{sp}(\mathbf{r}_{ij},\boldsymbol{\Omega}_i,\boldsymbol{\Omega}_j)=
523 \frac{\nu_0}{2}[s(r_{ij})w(\mathbf{r}_{ij},
524 \boldsymbol{\Omega}_i,\boldsymbol{\Omega}_j) +
525 s^\prime(r_{ij})w^\prime(\mathbf{r}_{ij},
526 \boldsymbol{\Omega}_i,\boldsymbol{\Omega}_j)]\ ,
527 \label{eq:stickyPot}
528 \end{equation}
529 where $\nu_0$ is a strength parameter for the sticky potential, and
530 $s$ and $s^\prime$ are cubic switching functions which turn off the
531 sticky interaction beyond the first solvation shell. The $w$ function
532 can be thought of as an attractive potential with tetrahedral
533 geometry:
534 \begin{equation}
535 w({\bf r}_{ij},{\bf \Omega}_i,{\bf \Omega}_j)=
536 \sin\theta_{ij}\sin2\theta_{ij}\cos2\phi_{ij},
537 \label{eq:stickyW}
538 \end{equation}
539 while the $w^\prime$ function counters the normal aligned and
540 anti-aligned structures favored by point dipoles:
541 \begin{equation}
542 w^\prime({\bf r}_{ij},{\bf \Omega}_i,{\bf \Omega}_j)=
543 (\cos\theta_{ij}-0.6)^2(\cos\theta_{ij}+0.8)^2-w^0,
544 \label{eq:stickyWprime}
545 \end{equation}
546 It should be noted that $w$ is proportional to the sum of the $Y_3^2$
547 and $Y_3^{-2}$ spherical harmonics (a linear combination which
548 enhances the tetrahedral geometry for hydrogen bonded structures),
549 while $w^\prime$ is a purely empirical function. A more detailed
550 description of the functional parts and variables in this potential
551 can be found in the original SSD
552 articles.\cite{liu96:new_model,liu96:monte_carlo,chandra99:ssd_md,Ichiye03}
553
554 Since SSD/E is a single-point {\it dipolar} model, the force
555 calculations are simplified significantly relative to the standard
556 {\it charged} multi-point models. In the original Monte Carlo
557 simulations using this model, Ichiye {\it et al.} reported that using
558 SSD decreased computer time by a factor of 6-7 compared to other
559 models.\cite{liu96:new_model} What is most impressive is that these savings
560 did not come at the expense of accurate depiction of the liquid state
561 properties. Indeed, SSD/E maintains reasonable agreement with the Head-Gordon
562 diffraction data for the structural features of liquid
563 water.\cite{hura00,liu96:new_model} Additionally, the dynamical properties
564 exhibited by SSD/E agree with experiment better than those of more
565 computationally expensive models (like TIP3P and
566 SPC/E).\cite{chandra99:ssd_md} The combination of speed and accurate depiction
567 of solvent properties makes SSD/E a very attractive model for the
568 simulation of large scale biochemical simulations.
569
570 Recent constant pressure simulations revealed issues in the original
571 SSD model that led to lower than expected densities at all target
572 pressures.\cite{Ichiye03,Gezelter04} The default model in {\sc oopse}
573 is therefore SSD/E, a density corrected derivative of SSD that
574 exhibits improved liquid structure and transport behavior. If the use
575 of a reaction field long-range interaction correction is desired, it
576 is recommended that the parameters be modified to those of the SSD/RF
577 model. Solvent parameters can be easily modified in an accompanying
578 \texttt{.bass} file as illustrated in the scheme below. A table of the
579 parameter values and the drawbacks and benefits of the different
580 density corrected SSD models can be found in
581 reference~\cite{Gezelter04}.
582
583 \begin{lstlisting}[float,caption={[A simulation of {\sc ssd} water]An example file showing a simulation including {\sc ssd} water.},label={sch:ssd}]
584
585 #include "water.mdl"
586
587 nComponents = 1;
588 component{
589 type = "SSD_water";
590 nMol = 864;
591 }
592
593 initialConfig = "liquidWater.init";
594
595 forceField = "DUFF";
596
597 /*
598 * The following two flags set the cutoff
599 * radius for the electrostatic forces
600 * as well as the skin thickness of the switching
601 * function.
602 */
603
604 electrostaticCutoffRadius = 9.2;
605 electrostaticSkinThickness = 1.38;
606
607 \end{lstlisting}
608
609
610 \subsection{\label{oopseSec:eam}Embedded Atom Method}
611
612 There are Molecular Dynamics packages which have the
613 capacity to simulate metallic systems, including some that have
614 parallel computational abilities\cite{plimpton93}. Potentials that
615 describe bonding transition metal
616 systems\cite{Finnis84,Ercolessi88,Chen90,Qi99,Ercolessi02} have an
617 attractive interaction which models ``Embedding''
618 a positively charged metal ion in the electron density due to the
619 free valance ``sea'' of electrons created by the surrounding atoms in
620 the system. A mostly-repulsive pairwise part of the potential
621 describes the interaction of the positively charged metal core ions
622 with one another. A particular potential description called the
623 Embedded Atom Method\cite{Daw84,FBD86,johnson89,Lu97}({\sc eam}) that has
624 particularly wide adoption has been selected for inclusion in {\sc oopse}. A
625 good review of {\sc eam} and other metallic potential formulations was written
626 by Voter.\cite{voter}
627
628 The {\sc eam} potential has the form:
629 \begin{eqnarray}
630 V & = & \sum_{i} F_{i}\left[\rho_{i}\right] + \sum_{i} \sum_{j \neq i}
631 \phi_{ij}({\bf r}_{ij}) \\
632 \rho_{i} & = & \sum_{j \neq i} f_{j}({\bf r}_{ij})
633 \end{eqnarray}
634 where $F_{i} $ is the embedding function that equates the energy required to embed a
635 positively-charged core ion $i$ into a linear superposition of
636 spherically averaged atomic electron densities given by
637 $\rho_{i}$. $\phi_{ij}$ is a primarily repulsive pairwise interaction
638 between atoms $i$ and $j$. In the original formulation of
639 {\sc eam}\cite{Daw84}, $\phi_{ij}$ was an entirely repulsive term, however
640 in later refinements to EAM have shown that non-uniqueness between $F$
641 and $\phi$ allow for more general forms for $\phi$.\cite{Daw89}
642 There is a cutoff distance, $r_{cut}$, which limits the
643 summations in the {\sc eam} equation to the few dozen atoms
644 surrounding atom $i$ for both the density $\rho$ and pairwise $\phi$
645 interactions. Foiles et al. fit EAM potentials for fcc metals Cu, Ag, Au, Ni, Pd, Pt and alloys of these metals\cite{FBD86}. These potential fits are in the DYNAMO 86 format and are included with {\sc oopse}.
646
647
648 \subsection{\label{oopseSec:pbc}Periodic Boundary Conditions}
649
650 \newcommand{\roundme}{\operatorname{round}}
651
652 \textit{Periodic boundary conditions} are widely used to simulate bulk properties with a relatively small number of particles. The
653 simulation box is replicated throughout space to form an infinite
654 lattice. During the simulation, when a particle moves in the primary
655 cell, its image in other cells move in exactly the same direction with
656 exactly the same orientation. Thus, as a particle leaves the primary
657 cell, one of its images will enter through the opposite face. If the
658 simulation box is large enough to avoid ``feeling'' the symmetries of
659 the periodic lattice, surface effects can be ignored. The available
660 periodic cells in OOPSE are cubic, orthorhombic and parallelepiped. We
661 use a $3 \times 3$ matrix, $\mathbf{H}$, to describe the shape and
662 size of the simulation box. $\mathbf{H}$ is defined:
663 \begin{equation}
664 \mathbf{H} = ( \mathbf{h}_x, \mathbf{h}_y, \mathbf{h}_z )
665 \end{equation}
666 Where $\mathbf{h}_j$ is the column vector of the $j$th axis of the
667 box. During the course of the simulation both the size and shape of
668 the box can be changed to allow volume fluctations when constraining
669 the pressure.
670
671 A real space vector, $\mathbf{r}$ can be transformed in to a box space
672 vector, $\mathbf{s}$, and back through the following transformations:
673 \begin{align}
674 \mathbf{s} &= \mathbf{H}^{-1} \mathbf{r} \\
675 \mathbf{r} &= \mathbf{H} \mathbf{s}
676 \end{align}
677 The vector $\mathbf{s}$ is now a vector expressed as the number of box
678 lengths in the $\mathbf{h}_x$, $\mathbf{h}_y$, and $\mathbf{h}_z$
679 directions. To find the minimum image of a vector $\mathbf{r}$, we
680 first convert it to its corresponding vector in box space, and then,
681 cast each element to lie on the in the range $[-0.5,0.5]$:
682 \begin{equation}
683 s_{i}^{\prime}=s_{i}-\roundme(s_{i})
684 \end{equation}
685 Where $s_i$ is the $i$th element of $\mathbf{s}$, and
686 $\roundme(s_i)$is given by
687 \begin{equation}
688 \roundme(x) =
689 \begin{cases}
690 \lfloor x+0.5 \rfloor & \text{if $x \ge 0$} \\
691 \lceil x-0.5 \rceil & \text{if $x < 0$ }
692 \end{cases}
693 \end{equation}
694 Here $\lfloor x \rfloor$ is the floor operator, and gives the largest
695 integer value that is not greater than $x$, and $\lceil x \rceil$ is
696 the ceiling operator, and gives the smallest integer that is not less
697 than $x$. For example, $\roundme(3.6)=4$, $\roundme(3.1)=3$,
698 $\roundme(-3.6)=-4$, $\roundme(-3.1)=-3$.
699
700 Finally, we obtain the minimum image coordinates $\mathbf{r}^{\prime}$ by
701 transforming back to real space,
702 \begin{equation}
703 \mathbf{r}^{\prime}=\mathbf{H}^{-1}\mathbf{s}^{\prime}%
704 \end{equation}
705 In this way, particles are allowed to diffuse freely in $\mathbf{r}$,
706 but their minimum images, $\mathbf{r}^{\prime}$ are used to compute
707 the interatomic forces.
708
709
710 \section{\label{oopseSec:IOfiles}Input and Output Files}
711
712 \subsection{{\sc bass} and Model Files}
713
714 Every {\sc oopse} simulation begins with a Bizarre Atom Simulation
715 Syntax ({\sc bass}) file. {\sc bass} is a script syntax that is parsed
716 by {\sc oopse} at runtime. The {\sc bass} file allows for the user to
717 completely describe the system they wish to simulate, as well as tailor
718 {\sc oopse}'s behavior during the simulation. {\sc bass} files are
719 denoted with the extension
720 \texttt{.bass}, an example file is shown in
721 Scheme~\ref{sch:bassExample}.
722
723 \begin{lstlisting}[float,caption={[An example of a complete {\sc bass} file] An example showing a complete {\sc bass} file.},label={sch:bassExample}]
724
725 molecule{
726 name = "Ar";
727 nAtoms = 1;
728 atom[0]{
729 type="Ar";
730 position( 0.0, 0.0, 0.0 );
731 }
732 }
733
734 nComponents = 1;
735 component{
736 type = "Ar";
737 nMol = 108;
738 }
739
740 initialConfig = "./argon.init";
741
742 forceField = "LJ";
743 ensemble = "NVE"; // specify the simulation enesemble
744 dt = 1.0; // the time step for integration
745 runTime = 1e3; // the total simulation run time
746 sampleTime = 100; // trajectory file frequency
747 statusTime = 50; // statistics file frequency
748
749 \end{lstlisting}
750
751 Within the \texttt{.bass} file it is necessary to provide a complete
752 description of the molecule before it is actually placed in the
753 simulation. The {\sc bass} syntax was originally developed with this
754 goal in mind, and allows for the specification of all the atoms in a
755 molecular prototype, as well as any bonds, bends, or torsions. These
756 descriptions can become lengthy for complex molecules, and it would be
757 inconvenient to duplicate the simulation at the beginning of each {\sc
758 bass} script. Addressing this issue {\sc bass} allows for the
759 inclusion of model files at the top of a \texttt{.bass} file. These
760 model files, denoted with the \texttt{.mdl} extension, allow the user
761 to describe a molecular prototype once, then simply include it into
762 each simulation containing that molecule. Returning to the example in
763 Scheme~\ref{sch:bassExample}, the \texttt{.mdl} file's contents would
764 be Scheme~\ref{sch:mdlExample}, and the new \texttt{.bass} file would
765 become Scheme~\ref{sch:bassExPrime}.
766
767 \begin{lstlisting}[float,caption={An example \texttt{.mdl} file.},label={sch:mdlExample}]
768
769 molecule{
770 name = "Ar";
771 nAtoms = 1;
772 atom[0]{
773 type="Ar";
774 position( 0.0, 0.0, 0.0 );
775 }
776 }
777
778 \end{lstlisting}
779
780 \begin{lstlisting}[float,caption={Revised {\sc bass} example.},label={sch:bassExPrime}]
781
782 #include "argon.mdl"
783
784 molecule{
785 name = "Ar";
786 nAtoms = 1;
787 atom[0]{
788 type="Ar";
789 position( 0.0, 0.0, 0.0 );
790 }
791 }
792
793 nComponents = 1;
794 component{
795 type = "Ar";
796 nMol = 108;
797 }
798
799 initialConfig = "./argon.init";
800
801 forceField = "LJ";
802 ensemble = "NVE";
803 dt = 1.0;
804 runTime = 1e3;
805 sampleTime = 100;
806 statusTime = 50;
807
808 \end{lstlisting}
809
810 \subsection{\label{oopseSec:coordFiles}Coordinate Files}
811
812 The standard format for storage of a systems coordinates is a modified
813 xyz-file syntax, the exact details of which can be seen in
814 Scheme~\ref{sch:dumpFormat}. As all bonding and molecular information
815 is stored in the \texttt{.bass} and \texttt{.mdl} files, the
816 coordinate files are simply the complete set of coordinates for each
817 atom at a given simulation time. One important note, although the
818 simulation propagates the complete rotation matrix, directional
819 entities are written out using quanternions, to save space in the
820 output files.
821
822 \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 $\mathbf{H}$ column vectors. The next lines are the atomic coordinates for all atoms in the system. First is the name followed by position, velocity, quanternions, and lastly angular momentum.},label=sch:dumpFormat]
823
824 nAtoms
825 time; Hxx Hyx Hzx; Hxy Hyy Hzy; Hxz Hyz Hzz;
826 Name1 x y z vx vy vz q0 q1 q2 q3 jx jy jz
827 Name2 x y z vx vy vz q0 q1 q2 q3 jx jy jz
828 etc...
829
830 \end{lstlisting}
831
832
833 There are three major files used by {\sc oopse} written in the
834 coordinate format, they are as follows: the initialization file
835 (\texttt{.init}), the simulation trajectory file (\texttt{.dump}), and
836 the final coordinates of the simulation. The initialization file is
837 necessary for {\sc oopse} to start the simulation with the proper
838 coordinates, and is generated before the simulation run. The
839 trajectory file is created at the beginning of the simulation, and is
840 used to store snapshots of the simulation at regular intervals. The
841 first frame is a duplication of the
842 \texttt{.init} file, and each subsequent frame is appended to the file
843 at an interval specified in the \texttt{.bass} file with the
844 \texttt{sampleTime} flag. The final coordinate file is the end of run file. The
845 \texttt{.eor} file stores the final configuration of the system for a
846 given simulation. The file is updated at the same time as the
847 \texttt{.dump} file, however, it only contains the most recent
848 frame. In this way, an \texttt{.eor} file may be used as the
849 initialization file to a second simulation in order to continue a
850 simulation or recover one from a processor that has crashed during the
851 course of the run.
852
853 \subsection{\label{oopseSec:initCoords}Generation of Initial Coordinates}
854
855 As was stated in Sec.~\ref{oopseSec:coordFiles}, an initialization
856 file is needed to provide the starting coordinates for a
857 simulation. The {\sc oopse} package provides a program called
858 \texttt{sysBuilder} to aid in the creation of the \texttt{.init}
859 file. \texttt{sysBuilder} uses {\sc bass}, and will recognize
860 arguments and parameters in the \texttt{.bass} file that would
861 otherwise be ignored by the simulation.
862
863 \subsection{The Statistics File}
864
865 The last output file generated by {\sc oopse} is the statistics
866 file. This file records such statistical quantities as the
867 instantaneous temperature, volume, pressure, etc. It is written out
868 with the frequency specified in the \texttt{.bass} file with the
869 \texttt{statusTime} keyword. The file allows the user to observe the
870 system variables as a function of simulation time while the simulation
871 is in progress. One useful function the statistics file serves is to
872 monitor the conserved quantity of a given simulation ensemble, this
873 allows the user to observe the stability of the integrator. The
874 statistics file is denoted with the \texttt{.stat} file extension.
875
876 \section{\label{oopseSec:mechanics}Mechanics}
877
878 \subsection{\label{oopseSec:integrate}Integrating the Equations of Motion: the Symplectic Step Integrator}
879
880 Integration of the equations of motion was carried out using the
881 symplectic splitting method proposed by Dullweber \emph{et
882 al.}.\cite{Dullweber1997} The reason for the selection of this
883 integrator, is the poor energy conservation of rigid body systems
884 using quaternion dynamics. While quaternions work well for
885 orientational motion in alternate ensembles, the microcanonical
886 ensemble has a constant energy requirement that is quite sensitive to
887 errors in the equations of motion. The original implementation of {\sc
888 oopse} utilized quaternions for rotational motion propagation;
889 however, a detailed investigation showed that they resulted in a
890 steady drift in the total energy, something that has been observed by
891 others.\cite{Laird97}
892
893 The key difference in the integration method proposed by Dullweber
894 \emph{et al}.~({\sc dlm}) is that the entire rotation matrix is propagated from
895 one time step to the next. In the past, this would not have been a
896 feasible option, since the rotation matrix for a single body is nine
897 elements long as opposed to three or four elements for Euler angles
898 and quaternions respectively. System memory has become much less of an
899 issue in recent times, and the {\sc dlm} method has used memory in
900 exchange for substantial benefits in energy conservation.
901
902 The {\sc dlm} method allows for Verlet style integration of both
903 linear and angular motion of rigid bodies. In the integration method,
904 the orientational propagation involves a sequence of matrix
905 evaluations to update the rotation matrix.\cite{Dullweber1997} These
906 matrix rotations are more costly computationally than the simpler
907 arithmetic quaternion propagation. With the same time step, a 1000 SSD
908 particle simulation shows an average 7\% increase in computation time
909 using the {\sc dlm} method in place of quaternions. This cost is more
910 than justified when comparing the energy conservation of the two
911 methods as illustrated in Fig.~\ref{timestep}.
912
913 \begin{figure}
914 \centering
915 \includegraphics[width=\linewidth]{timeStep.eps}
916 \caption[Energy conservation for quaternion versus {\sc dlm} dynamics]{Energy conservation using quaternion based integration versus
917 the {\sc dlm} method with
918 increasing time step. For each time step, the dotted line is total
919 energy using the {\sc dlm} integrator, and the solid line comes
920 from the quaternion integrator. The larger time step plots are shifted
921 up from the true energy baseline for clarity.}
922 \label{timestep}
923 \end{figure}
924
925 In Fig.~\ref{timestep}, the resulting energy drift at various time
926 steps for both the {\sc dlm} and quaternion integration schemes
927 is compared. All of the 1000 SSD particle simulations started with the
928 same configuration, and the only difference was the method for
929 handling rotational motion. At time steps of 0.1 and 0.5 fs, both
930 methods for propagating particle rotation conserve energy fairly well,
931 with the quaternion method showing a slight energy drift over time in
932 the 0.5 fs time step simulation. At time steps of 1 and 2 fs, the
933 energy conservation benefits of the {\sc dlm} method are clearly
934 demonstrated. Thus, while maintaining the same degree of energy
935 conservation, one can take considerably longer time steps, leading to
936 an overall reduction in computation time.
937
938 Energy drift in these SSD particle simulations was unnoticeable for
939 time steps up to three femtoseconds. A slight energy drift on the
940 order of 0.012 kcal/mol per nanosecond was observed at a time step of
941 four femtoseconds, and as expected, this drift increases dramatically
942 with increasing time step.
943
944
945 \subsection{\label{sec:extended}Extended Systems for other Ensembles}
946
947
948 {\sc oopse} implements a
949
950
951 \subsection{\label{oopseSec:noseHooverThermo}Nose-Hoover Thermostatting}
952
953 To mimic the effects of being in a constant temperature ({\sc nvt})
954 ensemble, {\sc oopse} uses the Nose-Hoover extended system
955 approach.\cite{Hoover85} In this method, the equations of motion for
956 the particle positions and velocities are
957 \begin{eqnarray}
958 \dot{{\bf r}} & = & {\bf v} \\
959 \dot{{\bf v}} & = & \frac{{\bf f}}{m} - \chi {\bf v}
960 \label{eq:nosehoovereom}
961 \end{eqnarray}
962
963 $\chi$ is an ``extra'' variable included in the extended system, and
964 it is propagated using the first order equation of motion
965 \begin{equation}
966 \dot{\chi} = \frac{1}{\tau_{T}} \left( \frac{T}{T_{target}} - 1 \right)
967 \label{eq:nosehooverext}
968 \end{equation}
969 where $T_{target}$ is the target temperature for the simulation, and
970 $\tau_{T}$ is a time constant for the thermostat.
971
972 To select the Nose-Hoover {\sc nvt} ensemble, the {\tt ensemble = NVT;}
973 command would be used in the simulation's {\sc bass} file. There is
974 some subtlety in choosing values for $\tau_{T}$, and it is usually set
975 to values of a few ps. Within a {\sc bass} file, $\tau_{T}$ could be
976 set to 1 ps using the {\tt tauThermostat = 1000; } command.
977
978 \subsection{\label{oopseSec:rattle}The {\sc rattle} Method for Bond
979 Constraints}
980
981 In order to satisfy the constraints of fixed bond lengths within {\sc
982 oopse}, we have implemented the {\sc rattle} algorithm of
983 Andersen.\cite{andersen83} The algorithm is a velocity verlet
984 formulation of the {\sc shake} method\cite{ryckaert77} of iteratively
985 solving the Lagrange multipliers of constraint. The system of lagrange
986 multipliers allows one to reformulate the equations of motion with
987 explicit constraint forces on the equations of
988 motion.\cite{fowles99:lagrange}
989
990 Consider a system described by qoordinates $q_1$ and $q_2$ subject to an
991 equation of constraint:
992 \begin{equation}
993 \sigma(q_1, q_2,t) = 0
994 \label{oopseEq:lm1}
995 \end{equation}
996 The Lagrange formulation of the equations of motion can be written:
997 \begin{equation}
998 \delta\int_{t_1}^{t_2}L\, dt =
999 \int_{t_1}^{t_2} \sum_i \biggl [ \frac{\partial L}{\partial q_i}
1000 - \frac{d}{dt}\biggl(\frac{\partial L}{\partial \dot{q}_i}
1001 \biggr ) \biggr] \delta q_i \, dt = 0
1002 \label{oopseEq:lm2}
1003 \end{equation}
1004 Here, $\delta q_i$ is not independent for each $q$, as $q_1$ and $q_2$
1005 are linked by $\sigma$. However, $\sigma$ is fixed at any given
1006 instant of time, giving:
1007 \begin{align}
1008 \delta\sigma &= \biggl( \frac{\partial\sigma}{\partial q_1} \delta q_1 %
1009 + \frac{\partial\sigma}{\partial q_2} \delta q_2 \biggr) = 0 \\
1010 %
1011 \frac{\partial\sigma}{\partial q_1} \delta q_1 &= %
1012 - \frac{\partial\sigma}{\partial q_2} \delta q_2 \\
1013 %
1014 \delta q_2 &= - \biggl(\frac{\partial\sigma}{\partial q_1} \bigg / %
1015 \frac{\partial\sigma}{\partial q_2} \biggr) \delta q_1
1016 \end{align}
1017 Substituted back into Eq.~\ref{oopseEq:lm2},
1018 \begin{equation}
1019 \int_{t_1}^{t_2}\biggl [ \biggl(\frac{\partial L}{\partial q_1}
1020 - \frac{d}{dt}\,\frac{\partial L}{\partial \dot{q}_1}
1021 \biggr)
1022 - \biggl( \frac{\partial L}{\partial q_1}
1023 - \frac{d}{dt}\,\frac{\partial L}{\partial \dot{q}_1}
1024 \biggr) \biggl(\frac{\partial\sigma}{\partial q_1} \bigg / %
1025 \frac{\partial\sigma}{\partial q_2} \biggr)\biggr] \delta q_1 \, dt = 0
1026 \label{oopseEq:lm3}
1027 \end{equation}
1028 Leading to,
1029 \begin{equation}
1030 \frac{\biggl(\frac{\partial L}{\partial q_1}
1031 - \frac{d}{dt}\,\frac{\partial L}{\partial \dot{q}_1}
1032 \biggr)}{\frac{\partial\sigma}{\partial q_1}} =
1033 \frac{\biggl(\frac{\partial L}{\partial q_2}
1034 - \frac{d}{dt}\,\frac{\partial L}{\partial \dot{q}_2}
1035 \biggr)}{\frac{\partial\sigma}{\partial q_2}}
1036 \label{oopseEq:lm4}
1037 \end{equation}
1038 This relation can only be statisfied, if both are equal to a single
1039 function $-\lambda(t)$,
1040 \begin{align}
1041 \frac{\biggl(\frac{\partial L}{\partial q_1}
1042 - \frac{d}{dt}\,\frac{\partial L}{\partial \dot{q}_1}
1043 \biggr)}{\frac{\partial\sigma}{\partial q_1}} &= -\lambda(t) \\
1044 %
1045 \frac{\partial L}{\partial q_1}
1046 - \frac{d}{dt}\,\frac{\partial L}{\partial \dot{q}_1} &=
1047 -\lambda(t)\,\frac{\partial\sigma}{\partial q_1} \\
1048 %
1049 \frac{\partial L}{\partial q_1}
1050 - \frac{d}{dt}\,\frac{\partial L}{\partial \dot{q}_1}
1051 + \mathcal{G}_i &= 0
1052 \end{align}
1053 Where $\mathcal{G}_i$, the force of constraint on $i$, is:
1054 \begin{equation}
1055 \mathcal{G}_i = \lambda(t)\,\frac{\partial\sigma}{\partial q_1}
1056 \label{oopseEq:lm5}
1057 \end{equation}
1058
1059 In a simulation, this would involve the solution of a set of $(m + n)$
1060 number of equations. Where $m$ is the number of constraints, and $n$
1061 is the number of constrained coordinates. In practice, this is not
1062 done, as the matrix inversion neccassary to solve the system of
1063 equations would be very time consuming to solve. Additionally, the
1064 numerical error in the solution of the set of $\lambda$'s would be
1065 compounded by the error inherent in propagating by the Velocity Verlet
1066 algorithm ($\Delta t^4$). The verlet propagation error is negligible
1067 in an unconstrained system, as one is interested in the statisitics of
1068 the run, and not that the run be numerically exact to the ``true''
1069 integration. This relates back to the ergodic hypothesis that a time
1070 integral of a valid trajectory will still give the correct enesemble
1071 average. However, in the case of constraints, if the equations of
1072 motion leave the ``true'' trajectory, they are departing from the
1073 constrained surface. The method that is used, is to iteratively solve
1074 for $\lambda(t)$ at each time step.
1075
1076 In {\sc rattle} the equations of motion are modified subject to the
1077 following two constraints:
1078 \begin{align}
1079 \sigma_{ij}[\mathbf{r}(t)] \equiv
1080 [ \mathbf{r}_i(t) - \mathbf{r}_j(t)]^2 - d_{ij}^2 &= 0 %
1081 \label{oopseEq:c1} \\
1082 %
1083 [\mathbf{\dot{r}}_i(t) - \mathbf{\dot{r}}_j(t)] \cdot
1084 [\mathbf{r}_i(t) - \mathbf{r}_j(t)] &= 0 \label{oopseEq:c2}
1085 \end{align}
1086 Eq.~\ref{oopseEq:c1} is the set of bond constraints, where $d_{ij}$ is
1087 the constrained distance between atom $i$ and
1088 $j$. Eq.~\ref{oopseEq:c2} constrains the velocities of $i$ and $j$ to
1089 be perpindicular to the bond vector, so that the bond can neither grow
1090 nor shrink. The constrained dynamics equations become:
1091 \begin{equation}
1092 m_i \mathbf{\ddot{r}}_i = \mathbf{F}_i + \mathbf{\mathcal{G}}_i
1093 \label{oopseEq:r1}
1094 \end{equation}
1095 Where,
1096 \begin{equation}
1097 \mathbf{\mathcal{G}}_i = - \sum_j \lambda_{ij}(t)\,\nabla \sigma_{ij}
1098 \label{oopseEq:r2}
1099 \end{equation}
1100
1101 In Velocity Verlet, if $\Delta t = h$, the propagation can be written:
1102 \begin{align}
1103 \mathbf{r}_i(t+h) &=
1104 \mathbf{r}_i(t) + h\mathbf{\dot{r}}(t) +
1105 \frac{h^2}{2m_i}\,\Bigl[ \mathbf{F}_i(t) +
1106 \mathbf{\mathcal{G}}_{Ri}(t) \Bigr] \label{oopseEq:vv1} \\
1107 %
1108 \mathbf{\dot{r}}_i(t+h) &=
1109 \mathbf{\dot{r}}_i(t) + \frac{h}{2m_i}
1110 \Bigl[ \mathbf{F}_i(t) + \mathbf{\mathcal{G}}_{Ri}(t) +
1111 \mathbf{F}_i(t+h) + \mathbf{\mathcal{G}}_{Vi}(t+h) \Bigr] %
1112 \label{oopseEq:vv2}
1113 \end{align}
1114
1115
1116
1117 \subsection{\label{oopseSec:zcons}Z-Constraint Method}
1118
1119 Based on fluctuation-dissipation theorem, a force auto-correlation
1120 method was developed to investigate the dynamics of ions inside the ion
1121 channels.\cite{Roux91} Time-dependent friction coefficient can be calculated
1122 from the deviation of the instantaneous force from its mean force.
1123
1124 %
1125
1126 \begin{equation}
1127 \xi(z,t)=\langle\delta F(z,t)\delta F(z,0)\rangle/k_{B}T
1128 \end{equation}
1129 where%
1130 \begin{equation}
1131 \delta F(z,t)=F(z,t)-\langle F(z,t)\rangle
1132 \end{equation}
1133
1134
1135 If the time-dependent friction decay rapidly, static friction coefficient can
1136 be approximated by%
1137
1138 \begin{equation}
1139 \xi^{static}(z)=\int_{0}^{\infty}\langle\delta F(z,t)\delta F(z,0)\rangle dt
1140 \end{equation}
1141
1142
1143 Hence, diffusion constant can be estimated by
1144 \begin{equation}
1145 D(z)=\frac{k_{B}T}{\xi^{static}(z)}=\frac{(k_{B}T)^{2}}{\int_{0}^{\infty
1146 }\langle\delta F(z,t)\delta F(z,0)\rangle dt}%
1147 \end{equation}
1148
1149
1150 \bigskip Z-Constraint method, which fixed the z coordinates of the molecules
1151 with respect to the center of the mass of the system, was proposed to obtain
1152 the forces required in force auto-correlation method.\cite{Marrink94} However,
1153 simply resetting the coordinate will move the center of the mass of the whole
1154 system. To avoid this problem, a new method was used at {\sc oopse}. Instead of
1155 resetting the coordinate, we reset the forces of z-constraint molecules as
1156 well as subtract the total constraint forces from the rest of the system after
1157 force calculation at each time step.
1158 \begin{align}
1159 F_{\alpha i}&=0\\
1160 V_{\alpha i}&=V_{\alpha i}-\frac{\sum\limits_{i}M_{_{\alpha i}}V_{\alpha i}}{\sum\limits_{i}M_{_{\alpha i}}}\\
1161 F_{\alpha i}&=F_{\alpha i}-\frac{M_{_{\alpha i}}}{\sum\limits_{\alpha}\sum\limits_{i}M_{_{\alpha i}}}\sum\limits_{\beta}F_{\beta}\\
1162 V_{\alpha i}&=V_{\alpha i}-\frac{\sum\limits_{\alpha}\sum\limits_{i}M_{_{\alpha i}}V_{\alpha i}}{\sum\limits_{\alpha}\sum\limits_{i}M_{_{\alpha i}}}
1163 \end{align}
1164
1165 At the very beginning of the simulation, the molecules may not be at its
1166 constraint position. To move the z-constraint molecule to the specified
1167 position, a simple harmonic potential is used%
1168
1169 \begin{equation}
1170 U(t)=\frac{1}{2}k_{Harmonic}(z(t)-z_{cons})^{2}%
1171 \end{equation}
1172 where $k_{Harmonic}$\bigskip\ is the harmonic force constant, $z(t)$ is
1173 current z coordinate of the center of mass of the z-constraint molecule, and
1174 $z_{cons}$ is the restraint position. Therefore, the harmonic force operated
1175 on the z-constraint molecule at time $t$ can be calculated by%
1176 \begin{equation}
1177 F_{z_{Harmonic}}(t)=-\frac{\partial U(t)}{\partial z(t)}=-k_{Harmonic}%
1178 (z(t)-z_{cons})
1179 \end{equation}
1180 Worthy of mention, other kinds of potential functions can also be used to
1181 drive the z-constraint molecule.
1182
1183 \section{\label{oopseSec:props}Trajectory Analysis}
1184
1185 \subsection{\label{oopseSec:staticProps}Static Property Analysis}
1186
1187 The static properties of the trajectories are analyzed with the
1188 program \texttt{staticProps}. The code is capable of calculating the following
1189 pair correlations between species A and B:
1190 \begin{itemize}
1191 \item $g_{\text{AB}}(r)$: Eq.~\ref{eq:gofr}
1192 \item $g_{\text{AB}}(r, \cos \theta)$: Eq.~\ref{eq:gofrCosTheta}
1193 \item $g_{\text{AB}}(r, \cos \omega)$: Eq.~\ref{eq:gofrCosOmega}
1194 \item $g_{\text{AB}}(x, y, z)$: Eq.~\ref{eq:gofrXYZ}
1195 \item $\langle \cos \omega \rangle_{\text{AB}}(r)$:
1196 Eq.~\ref{eq:cosOmegaOfR}
1197 \end{itemize}
1198
1199 The first pair correlation, $g_{\text{AB}}(r)$, is defined as follows:
1200 \begin{equation}
1201 g_{\text{AB}}(r) = \frac{V}{N_{\text{A}}N_{\text{B}}}\langle %%
1202 \sum_{i \in \text{A}} \sum_{j \in \text{B}} %%
1203 \delta( r - |\mathbf{r}_{ij}|) \rangle \label{eq:gofr}
1204 \end{equation}
1205 Where $\mathbf{r}_{ij}$ is the vector
1206 \begin{equation*}
1207 \mathbf{r}_{ij} = \mathbf{r}_j - \mathbf{r}_i \notag
1208 \end{equation*}
1209 and $\frac{V}{N_{\text{A}}N_{\text{B}}}$ normalizes the average over
1210 the expected pair density at a given $r$.
1211
1212 The next two pair correlations, $g_{\text{AB}}(r, \cos \theta)$ and
1213 $g_{\text{AB}}(r, \cos \omega)$, are similar in that they are both two
1214 dimensional histograms. Both use $r$ for the primary axis then a
1215 $\cos$ for the secondary axis ($\cos \theta$ for
1216 Eq.~\ref{eq:gofrCosTheta} and $\cos \omega$ for
1217 Eq.~\ref{eq:gofrCosOmega}). This allows for the investigator to
1218 correlate alignment on directional entities. $g_{\text{AB}}(r, \cos
1219 \theta)$ is defined as follows:
1220 \begin{equation}
1221 g_{\text{AB}}(r, \cos \theta) = \frac{V}{N_{\text{A}}N_{\text{B}}}\langle
1222 \sum_{i \in \text{A}} \sum_{j \in \text{B}}
1223 \delta( \cos \theta - \cos \theta_{ij})
1224 \delta( r - |\mathbf{r}_{ij}|) \rangle
1225 \label{eq:gofrCosTheta}
1226 \end{equation}
1227 Where
1228 \begin{equation*}
1229 \cos \theta_{ij} = \mathbf{\hat{i}} \cdot \mathbf{\hat{r}}_{ij}
1230 \end{equation*}
1231 Here $\mathbf{\hat{i}}$ is the unit directional vector of species $i$
1232 and $\mathbf{\hat{r}}_{ij}$ is the unit vector associated with vector
1233 $\mathbf{r}_{ij}$.
1234
1235 The second two dimensional histogram is of the form:
1236 \begin{equation}
1237 g_{\text{AB}}(r, \cos \omega) =
1238 \frac{V}{N_{\text{A}}N_{\text{B}}}\langle
1239 \sum_{i \in \text{A}} \sum_{j \in \text{B}}
1240 \delta( \cos \omega - \cos \omega_{ij})
1241 \delta( r - |\mathbf{r}_{ij}|) \rangle \label{eq:gofrCosOmega}
1242 \end{equation}
1243 Here
1244 \begin{equation*}
1245 \cos \omega_{ij} = \mathbf{\hat{i}} \cdot \mathbf{\hat{j}}
1246 \end{equation*}
1247 Again, $\mathbf{\hat{i}}$ and $\mathbf{\hat{j}}$ are the unit
1248 directional vectors of species $i$ and $j$.
1249
1250 The static analysis code is also cable of calculating a three
1251 dimensional pair correlation of the form:
1252 \begin{equation}\label{eq:gofrXYZ}
1253 g_{\text{AB}}(x, y, z) =
1254 \frac{V}{N_{\text{A}}N_{\text{B}}}\langle
1255 \sum_{i \in \text{A}} \sum_{j \in \text{B}}
1256 \delta( x - x_{ij})
1257 \delta( y - y_{ij})
1258 \delta( z - z_{ij}) \rangle
1259 \end{equation}
1260 Where $x_{ij}$, $y_{ij}$, and $z_{ij}$ are the $x$, $y$, and $z$
1261 components respectively of vector $\mathbf{r}_{ij}$.
1262
1263 The final pair correlation is similar to
1264 Eq.~\ref{eq:gofrCosOmega}. $\langle \cos \omega
1265 \rangle_{\text{AB}}(r)$ is calculated in the following way:
1266 \begin{equation}\label{eq:cosOmegaOfR}
1267 \langle \cos \omega \rangle_{\text{AB}}(r) =
1268 \langle \sum_{i \in \text{A}} \sum_{j \in \text{B}}
1269 (\cos \omega_{ij}) \delta( r - |\mathbf{r}_{ij}|) \rangle
1270 \end{equation}
1271 Here $\cos \omega_{ij}$ is defined in the same way as in
1272 Eq.~\ref{eq:gofrCosOmega}. This equation is a single dimensional pair
1273 correlation that gives the average correlation of two directional
1274 entities as a function of their distance from each other.
1275
1276 All static properties are calculated on a frame by frame basis. The
1277 trajectory is read a single frame at a time, and the appropriate
1278 calculations are done on each frame. Once one frame is finished, the
1279 next frame is read in, and a running average of the property being
1280 calculated is accumulated in each frame. The program allows for the
1281 user to specify more than one property be calculated in single run,
1282 preventing the need to read a file multiple times.
1283
1284 \subsection{\label{dynamicProps}Dynamic Property Analysis}
1285
1286 The dynamic properties of a trajectory are calculated with the program
1287 \texttt{dynamicProps}. The program will calculate the following properties:
1288 \begin{gather}
1289 \langle | \mathbf{r}(t) - \mathbf{r}(0) |^2 \rangle \label{eq:rms}\\
1290 \langle \mathbf{v}(t) \cdot \mathbf{v}(0) \rangle \label{eq:velCorr} \\
1291 \langle \mathbf{j}(t) \cdot \mathbf{j}(0) \rangle \label{eq:angularVelCorr}
1292 \end{gather}
1293
1294 Eq.~\ref{eq:rms} is the root mean square displacement
1295 function. Eq.~\ref{eq:velCorr} and Eq.~\ref{eq:angularVelCorr} are the
1296 velocity and angular velocity correlation functions respectively. The
1297 latter is only applicable to directional species in the simulation.
1298
1299 The \texttt{dynamicProps} program handles he file in a manner different from
1300 \texttt{staticProps}. As the properties calculated by this program are time
1301 dependent, multiple frames must be read in simultaneously by the
1302 program. For small trajectories this is no problem, and the entire
1303 trajectory is read into memory. However, for long trajectories of
1304 large systems, the files can be quite large. In order to accommodate
1305 large files, \texttt{dynamicProps} adopts a scheme whereby two blocks of memory
1306 are allocated to read in several frames each.
1307
1308 In this two block scheme, the correlation functions are first
1309 calculated within each memory block, then the cross correlations
1310 between the frames contained within the two blocks are
1311 calculated. Once completed, the memory blocks are incremented, and the
1312 process is repeated. A diagram illustrating the process is shown in
1313 Fig.~\ref{oopseFig:dynamicPropsMemory}. As was the case with
1314 \texttt{staticProps}, multiple properties may be calculated in a
1315 single run to avoid multiple reads on the same file.
1316
1317
1318
1319 \section{\label{oopseSec:design}Program Design}
1320
1321 \subsection{\label{sec:architecture} {\sc oopse} Architecture}
1322
1323 The core of OOPSE is divided into two main object libraries:
1324 \texttt{libBASS} and \texttt{libmdtools}. \texttt{libBASS} is the
1325 library developed around the parsing engine and \texttt{libmdtools}
1326 is the software library developed around the simulation engine. These
1327 two libraries are designed to encompass all the basic functions and
1328 tools that {\sc oopse} provides. Utility programs, such as the
1329 property analyzers, need only link against the software libraries to
1330 gain access to parsing, force evaluation, and input / output
1331 routines.
1332
1333 Contained in \texttt{libBASS} are all the routines associated with
1334 reading and parsing the \texttt{.bass} input files. Given a
1335 \texttt{.bass} file, \texttt{libBASS} will open it and any associated
1336 \texttt{.mdl} files; then create structures in memory that are
1337 templates of all the molecules specified in the input files. In
1338 addition, any simulation parameters set in the \texttt{.bass} file
1339 will be placed in a structure for later query by the controlling
1340 program.
1341
1342 Located in \texttt{libmdtools} are all other routines necessary to a
1343 Molecular Dynamics simulation. The library uses the main data
1344 structures returned by \texttt{libBASS} to initialize the various
1345 parts of the simulation: the atom structures and positions, the force
1346 field, the integrator, \emph{et cetera}. After initialization, the
1347 library can be used to perform a variety of tasks: integrate a
1348 Molecular Dynamics trajectory, query phase space information from a
1349 specific frame of a completed trajectory, or even recalculate force or
1350 energetic information about specific frames from a completed
1351 trajectory.
1352
1353 With these core libraries in place, several programs have been
1354 developed to utilize the routines provided by \texttt{libBASS} and
1355 \texttt{libmdtools}. The main program of the package is \texttt{oopse}
1356 and the corresponding parallel version \texttt{oopse\_MPI}. These two
1357 programs will take the \texttt{.bass} file, and create then integrate
1358 the simulation specified in the script. The two analysis programs
1359 \texttt{staticProps} and \texttt{dynamicProps} utilize the core
1360 libraries to initialize and read in trajectories from previously
1361 completed simulations, in addition to the ability to use functionality
1362 from \texttt{libmdtools} to recalculate forces and energies at key
1363 frames in the trajectories. Lastly, the family of system building
1364 programs (Sec.~\ref{oopseSec:initCoords}) also use the libraries to
1365 store and output the system configurations they create.
1366
1367 \subsection{\label{oopseSec:parallelization} Parallelization of {\sc oopse}}
1368
1369 Although processor power is continually growing month by month, it is
1370 still unreasonable to simulate systems of more then a 1000 atoms on a
1371 single processor. To facilitate study of larger system sizes or
1372 smaller systems on long time scales in a reasonable period of time,
1373 parallel methods were developed allowing multiple CPU's to share the
1374 simulation workload. Three general categories of parallel
1375 decomposition method's have been developed including atomic, spatial
1376 and force decomposition methods.
1377
1378 Algorithmically simplest of the three method's is atomic decomposition
1379 where N particles in a simulation are split among P processors for the
1380 duration of the simulation. Computational cost scales as an optimal
1381 $O(N/P)$ for atomic decomposition. Unfortunately all processors must
1382 communicate positions and forces with all other processors leading
1383 communication to scale as an unfavorable $O(N)$ independent of the
1384 number of processors. This communication bottleneck led to the
1385 development of spatial and force decomposition methods in which
1386 communication among processors scales much more favorably. Spatial or
1387 domain decomposition divides the physical spatial domain into 3D boxes
1388 in which each processor is responsible for calculation of forces and
1389 positions of particles located in its box. Particles are reassigned to
1390 different processors as they move through simulation space. To
1391 calculate forces on a given particle, a processor must know the
1392 positions of particles within some cutoff radius located on nearby
1393 processors instead of the positions of particles on all
1394 processors. Both communication between processors and computation
1395 scale as $O(N/P)$ in the spatial method. However, spatial
1396 decomposition adds algorithmic complexity to the simulation code and
1397 is not very efficient for small N since the overall communication
1398 scales as the surface to volume ratio $(N/P)^{2/3}$ in three
1399 dimensions.
1400
1401 Force decomposition assigns particles to processors based on a block
1402 decomposition of the force matrix. Processors are split into a
1403 optimally square grid forming row and column processor groups. Forces
1404 are calculated on particles in a given row by particles located in
1405 that processors column assignment. Force decomposition is less complex
1406 to implement then the spatial method but still scales computationally
1407 as $O(N/P)$ and scales as $(N/\sqrt{p})$ in communication
1408 cost. Plimpton also found that force decompositions scales more
1409 favorably then spatial decomposition up to 10,000 atoms and favorably
1410 competes with spatial methods for up to 100,000 atoms.
1411
1412 \subsection{\label{oopseSec:memAlloc}Memory Issues in Trajectory Analysis}
1413
1414 For large simulations, the trajectory files can sometimes reach sizes
1415 in excess of several gigabytes. In order to effectively analyze that
1416 amount of data+, two memory management schemes have been devised for
1417 \texttt{staticProps} and for \texttt{dynamicProps}. The first scheme,
1418 developed for \texttt{staticProps}, is the simplest. As each frame's
1419 statistics are calculated independent of each other, memory is
1420 allocated for each frame, then freed once correlation calculations are
1421 complete for the snapshot. To prevent multiple passes through a
1422 potentially large file, \texttt{staticProps} is capable of calculating
1423 all requested correlations per frame with only a single pair loop in
1424 each frame and a single read through of the file.
1425
1426 The second, more advanced memory scheme, is used by
1427 \texttt{dynamicProps}. Here, the program must have multiple frames in
1428 memory to calculate time dependent correlations. In order to prevent a
1429 situation where the program runs out of memory due to large
1430 trajectories, the user is able to specify that the trajectory be read
1431 in blocks. The number of frames in each block is specified by the
1432 user, and upon reading a block of the trajectory,
1433 \texttt{dynamicProps} will calculate all of the time correlation frame
1434 pairs within the block. After in block correlations are complete, a
1435 second block of the trajectory is read, and the cross correlations are
1436 calculated between the two blocks. this second block is then freed and
1437 then incremented and the process repeated until the end of the
1438 trajectory. Once the end is reached, the first block is freed then
1439 incremented, and the again the internal time correlations are
1440 calculated. The algorithm with the second block is then repeated with
1441 the new origin block, until all frame pairs have been correlated in
1442 time. This process is illustrated in
1443 Fig.~\ref{oopseFig:dynamicPropsMemory}.
1444
1445 \begin{figure}
1446 \centering
1447 \includegraphics[width=\linewidth]{dynamicPropsMem.eps}
1448 \caption[A representation of the block correlations in \texttt{dynamicProps}]{This diagram illustrates the memory management used by \texttt{dynamicProps}, which follows the scheme: $\sum^{N_{\text{memory blocks}}}_{i=1}[ \operatorname{self}(i) + \sum^{N_{\text{memory blocks}}}_{j>i} \operatorname{cross}(i,j)]$. The shaded region represents the self correlation of the memory block, and the open blocks are read one at a time and the cross correlations between blocks are calculated.}
1449 \label{oopseFig:dynamicPropsMemory}
1450 \end{figure}
1451
1452 \subsection{\label{openSource}Open Source and Distribution License}
1453
1454 \section{\label{oopseSec:conclusion}Conclusion}
1455
1456 We have presented the design and implementation of our open source
1457 simulation package {\sc oopse}. The package offers novel
1458 capabilities to the field of Molecular Dynamics simulation packages in
1459 the form of dipolar force fields, and symplectic integration of rigid
1460 body dynamics. It is capable of scaling across multiple processors
1461 through the use of MPI. It also implements several integration
1462 ensembles allowing the end user control over temperature and
1463 pressure. In addition, it is capable of integrating constrained
1464 dynamics through both the {\sc rattle} algorithm and the z-constraint
1465 method.
1466
1467 These features are all brought together in a single open-source
1468 development package. This allows researchers to not only benefit from
1469 {\sc oopse}, but also contribute to {\sc oopse}'s development as
1470 well.Documentation and source code for {\sc oopse} can be downloaded
1471 from \texttt{http://www.openscience.org/oopse/}.
1472