ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/mattDisertation/oopse.tex
Revision: 1083
Committed: Fri Mar 5 00:07:19 2004 UTC (20 years, 4 months ago) by mmeineke
Content type: application/x-tex
File size: 69700 byte(s)
Log Message:
added several figures to the lipid paper, changed the ndthesis class around a little. some small fixes in oopse.

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 a lipid model in {\sc duff}]{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
635 required to embed a positively-charged core ion $i$ into a linear
636 superposition of spherically averaged atomic electron densities given
637 by $\rho_{i}$. $\phi_{ij}$ is a primarily repulsive pairwise
638 interaction between atoms $i$ and $j$. In the original formulation of
639 {\sc eam}\cite{Daw84}, $\phi_{ij}$ was an entirely repulsive term,
640 however in later refinements to {\sc eam} have shown that non-uniqueness
641 between $F$ and $\phi$ allow for more general forms for
642 $\phi$.\cite{Daw89} There is a cutoff distance, $r_{cut}$, which
643 limits the 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 \emph{et al}.~fit {\sc eam} potentials for the fcc
646 metals Cu, Ag, Au, Ni, Pd, Pt and alloys of these metals.\cite{FBD86}
647 These fits, are included in {\sc oopse}.
648
649 \subsection{\label{oopseSec:pbc}Periodic Boundary Conditions}
650
651 \newcommand{\roundme}{\operatorname{round}}
652
653 \textit{Periodic boundary conditions} are widely used to simulate bulk properties with a relatively small number of particles. The
654 simulation box is replicated throughout space to form an infinite
655 lattice. During the simulation, when a particle moves in the primary
656 cell, its image in other cells move in exactly the same direction with
657 exactly the same orientation. Thus, as a particle leaves the primary
658 cell, one of its images will enter through the opposite face. If the
659 simulation box is large enough to avoid ``feeling'' the symmetries of
660 the periodic lattice, surface effects can be ignored. The available
661 periodic cells in OOPSE are cubic, orthorhombic and parallelepiped. We
662 use a $3 \times 3$ matrix, $\mathsf{H}$, to describe the shape and
663 size of the simulation box. $\mathsf{H}$ is defined:
664 \begin{equation}
665 \mathsf{H} = ( \mathbf{h}_x, \mathbf{h}_y, \mathbf{h}_z )
666 \end{equation}
667 Where $\mathbf{h}_j$ is the column vector of the $j$th axis of the
668 box. During the course of the simulation both the size and shape of
669 the box can be changed to allow volume fluctuations when constraining
670 the pressure.
671
672 A real space vector, $\mathbf{r}$ can be transformed in to a box space
673 vector, $\mathbf{s}$, and back through the following transformations:
674 \begin{align}
675 \mathbf{s} &= \mathsf{H}^{-1} \mathbf{r} \\
676 \mathbf{r} &= \mathsf{H} \mathbf{s}
677 \end{align}
678 The vector $\mathbf{s}$ is now a vector expressed as the number of box
679 lengths in the $\mathbf{h}_x$, $\mathbf{h}_y$, and $\mathbf{h}_z$
680 directions. To find the minimum image of a vector $\mathbf{r}$, we
681 first convert it to its corresponding vector in box space, and then,
682 cast each element to lie on the in the range $[-0.5,0.5]$:
683 \begin{equation}
684 s_{i}^{\prime}=s_{i}-\roundme(s_{i})
685 \end{equation}
686 Where $s_i$ is the $i$th element of $\mathbf{s}$, and
687 $\roundme(s_i)$is given by
688 \begin{equation}
689 \roundme(x) =
690 \begin{cases}
691 \lfloor x+0.5 \rfloor & \text{if $x \ge 0$} \\
692 \lceil x-0.5 \rceil & \text{if $x < 0$ }
693 \end{cases}
694 \end{equation}
695 Here $\lfloor x \rfloor$ is the floor operator, and gives the largest
696 integer value that is not greater than $x$, and $\lceil x \rceil$ is
697 the ceiling operator, and gives the smallest integer that is not less
698 than $x$. For example, $\roundme(3.6)=4$, $\roundme(3.1)=3$,
699 $\roundme(-3.6)=-4$, $\roundme(-3.1)=-3$.
700
701 Finally, we obtain the minimum image coordinates $\mathbf{r}^{\prime}$ by
702 transforming back to real space,
703 \begin{equation}
704 \mathbf{r}^{\prime}=\mathsf{H}^{-1}\mathbf{s}^{\prime}%
705 \end{equation}
706 In this way, particles are allowed to diffuse freely in $\mathbf{r}$,
707 but their minimum images, $\mathbf{r}^{\prime}$ are used to compute
708 the inter-atomic forces.
709
710
711 \section{\label{oopseSec:IOfiles}Input and Output Files}
712
713 \subsection{{\sc bass} and Model Files}
714
715 Every {\sc oopse} simulation begins with a Bizarre Atom Simulation
716 Syntax ({\sc bass}) file. {\sc bass} is a script syntax that is parsed
717 by {\sc oopse} at runtime. The {\sc bass} file allows for the user to
718 completely describe the system they wish to simulate, as well as tailor
719 {\sc oopse}'s behavior during the simulation. {\sc bass} files are
720 denoted with the extension
721 \texttt{.bass}, an example file is shown in
722 Scheme~\ref{sch:bassExample}.
723
724 \begin{lstlisting}[float,caption={[An example of a complete {\sc bass} file] An example showing a complete {\sc bass} file.},label={sch:bassExample}]
725
726 molecule{
727 name = "Ar";
728 nAtoms = 1;
729 atom[0]{
730 type="Ar";
731 position( 0.0, 0.0, 0.0 );
732 }
733 }
734
735 nComponents = 1;
736 component{
737 type = "Ar";
738 nMol = 108;
739 }
740
741 initialConfig = "./argon.init";
742
743 forceField = "LJ";
744 ensemble = "NVE"; // specify the simulation ensemble
745 dt = 1.0; // the time step for integration
746 runTime = 1e3; // the total simulation run time
747 sampleTime = 100; // trajectory file frequency
748 statusTime = 50; // statistics file frequency
749
750 \end{lstlisting}
751
752 Within the \texttt{.bass} file it is necessary to provide a complete
753 description of the molecule before it is actually placed in the
754 simulation. The {\sc bass} syntax was originally developed with this
755 goal in mind, and allows for the specification of all the atoms in a
756 molecular prototype, as well as any bonds, bends, or torsions. These
757 descriptions can become lengthy for complex molecules, and it would be
758 inconvenient to duplicate the simulation at the beginning of each {\sc
759 bass} script. Addressing this issue {\sc bass} allows for the
760 inclusion of model files at the top of a \texttt{.bass} file. These
761 model files, denoted with the \texttt{.mdl} extension, allow the user
762 to describe a molecular prototype once, then simply include it into
763 each simulation containing that molecule. Returning to the example in
764 Scheme~\ref{sch:bassExample}, the \texttt{.mdl} file's contents would
765 be Scheme~\ref{sch:mdlExample}, and the new \texttt{.bass} file would
766 become Scheme~\ref{sch:bassExPrime}.
767
768 \begin{lstlisting}[float,caption={An example \texttt{.mdl} file.},label={sch:mdlExample}]
769
770 molecule{
771 name = "Ar";
772 nAtoms = 1;
773 atom[0]{
774 type="Ar";
775 position( 0.0, 0.0, 0.0 );
776 }
777 }
778
779 \end{lstlisting}
780
781 \begin{lstlisting}[float,caption={Revised {\sc bass} example.},label={sch:bassExPrime}]
782
783 #include "argon.mdl"
784
785 nComponents = 1;
786 component{
787 type = "Ar";
788 nMol = 108;
789 }
790
791 initialConfig = "./argon.init";
792
793 forceField = "LJ";
794 ensemble = "NVE";
795 dt = 1.0;
796 runTime = 1e3;
797 sampleTime = 100;
798 statusTime = 50;
799
800 \end{lstlisting}
801
802 \subsection{\label{oopseSec:coordFiles}Coordinate Files}
803
804 The standard format for storage of a systems coordinates is a modified
805 xyz-file syntax, the exact details of which can be seen in
806 Scheme~\ref{sch:dumpFormat}. As all bonding and molecular information
807 is stored in the \texttt{.bass} and \texttt{.mdl} files, the
808 coordinate files are simply the complete set of coordinates for each
809 atom at a given simulation time. One important note, although the
810 simulation propagates the complete rotation matrix, directional
811 entities are written out using quanternions, to save space in the
812 output files.
813
814 \begin{lstlisting}[float,caption={[The format of the coordinate files]Shows the format of the coordinate files. The fist line is the number of atoms. The second line begins with the time stamp followed by the three $\mathsf{H}$ column vectors. It is important to note, that for extended system ensembles, additional information pertinent to the integrators may be stored on this line as well.. The next lines are the atomic coordinates for all atoms in the system. First is the name followed by position, velocity, quanternions, and lastly angular velocities.},label=sch:dumpFormat]
815
816 nAtoms
817 time; Hxx Hyx Hzx; Hxy Hyy Hzy; Hxz Hyz Hzz;
818 Name1 x y z vx vy vz q0 q1 q2 q3 jx jy jz
819 Name2 x y z vx vy vz q0 q1 q2 q3 jx jy jz
820 etc...
821
822 \end{lstlisting}
823
824
825 There are three major files used by {\sc oopse} written in the
826 coordinate format, they are as follows: the initialization file
827 (\texttt{.init}), the simulation trajectory file (\texttt{.dump}), and
828 the final coordinates of the simulation. The initialization file is
829 necessary for {\sc oopse} to start the simulation with the proper
830 coordinates, and is generated before the simulation run. The
831 trajectory file is created at the beginning of the simulation, and is
832 used to store snapshots of the simulation at regular intervals. The
833 first frame is a duplication of the
834 \texttt{.init} file, and each subsequent frame is appended to the file
835 at an interval specified in the \texttt{.bass} file with the
836 \texttt{sampleTime} flag. The final coordinate file is the end of run file. The
837 \texttt{.eor} file stores the final configuration of the system for a
838 given simulation. The file is updated at the same time as the
839 \texttt{.dump} file, however, it only contains the most recent
840 frame. In this way, an \texttt{.eor} file may be used as the
841 initialization file to a second simulation in order to continue a
842 simulation or recover one from a processor that has crashed during the
843 course of the run.
844
845 \subsection{\label{oopseSec:initCoords}Generation of Initial Coordinates}
846
847 As was stated in Sec.~\ref{oopseSec:coordFiles}, an initialization
848 file is needed to provide the starting coordinates for a
849 simulation. The {\sc oopse} package provides several system building
850 programs to aid in the creation of the \texttt{.init}
851 file. The programs use {\sc bass}, and will recognize
852 arguments and parameters in the \texttt{.bass} file that would
853 otherwise be ignored by the simulation.
854
855 \subsection{The Statistics File}
856
857 The last output file generated by {\sc oopse} is the statistics
858 file. This file records such statistical quantities as the
859 instantaneous temperature, volume, pressure, etc. It is written out
860 with the frequency specified in the \texttt{.bass} file with the
861 \texttt{statusTime} keyword. The file allows the user to observe the
862 system variables as a function of simulation time while the simulation
863 is in progress. One useful function the statistics file serves is to
864 monitor the conserved quantity of a given simulation ensemble, this
865 allows the user to observe the stability of the integrator. The
866 statistics file is denoted with the \texttt{.stat} file extension.
867
868 \section{\label{oopseSec:mechanics}Mechanics}
869
870 \subsection{\label{oopseSec:integrate}Integrating the Equations of Motion: the Symplectic Step Integrator}
871
872 Integration of the equations of motion was carried out using the
873 symplectic splitting method proposed by Dullweber \emph{et
874 al.}.\cite{Dullweber1997} The reason for the selection of this
875 integrator, is the poor energy conservation of rigid body systems
876 using quaternion dynamics. While quaternions work well for
877 orientational motion in alternate ensembles, the microcanonical
878 ensemble has a constant energy requirement that is quite sensitive to
879 errors in the equations of motion. The original implementation of {\sc
880 oopse} utilized quaternions for rotational motion propagation;
881 however, a detailed investigation showed that they resulted in a
882 steady drift in the total energy, something that has been observed by
883 others.\cite{Laird97}
884
885 The key difference in the integration method proposed by Dullweber
886 \emph{et al}.~({\sc dlm}) is that the entire rotation matrix is propagated from
887 one time step to the next. In the past, this would not have been a
888 feasible option, since the rotation matrix for a single body is nine
889 elements long as opposed to three or four elements for Euler angles
890 and quaternions respectively. System memory has become much less of an
891 issue in recent times, and the {\sc dlm} method has used memory in
892 exchange for substantial benefits in energy conservation.
893
894 The {\sc dlm} method allows for Verlet style integration of both
895 linear and angular motion of rigid bodies. In the integration method,
896 the orientational propagation involves a sequence of matrix
897 evaluations to update the rotation matrix.\cite{Dullweber1997} These
898 matrix rotations are more costly computationally than the simpler
899 arithmetic quaternion propagation. With the same time step, a 1000 SSD
900 particle simulation shows an average 7\% increase in computation time
901 using the {\sc dlm} method in place of quaternions. This cost is more
902 than justified when comparing the energy conservation of the two
903 methods as illustrated in Fig.~\ref{timestep}.
904
905 \begin{figure}
906 \centering
907 \includegraphics[width=\linewidth]{timeStep.eps}
908 \caption[Energy conservation for quaternion versus {\sc dlm} dynamics]{Energy conservation using quaternion based integration versus
909 the {\sc dlm} method with
910 increasing time step. For each time step, the dotted line is total
911 energy using the {\sc dlm} integrator, and the solid line comes
912 from the quaternion integrator. The larger time step plots are shifted
913 up from the true energy baseline for clarity.}
914 \label{timestep}
915 \end{figure}
916
917 In Fig.~\ref{timestep}, the resulting energy drift at various time
918 steps for both the {\sc dlm} and quaternion integration schemes
919 is compared. All of the 1000 SSD particle simulations started with the
920 same configuration, and the only difference was the method for
921 handling rotational motion. At time steps of 0.1 and 0.5 fs, both
922 methods for propagating particle rotation conserve energy fairly well,
923 with the quaternion method showing a slight energy drift over time in
924 the 0.5 fs time step simulation. At time steps of 1 and 2 fs, the
925 energy conservation benefits of the {\sc dlm} method are clearly
926 demonstrated. Thus, while maintaining the same degree of energy
927 conservation, one can take considerably longer time steps, leading to
928 an overall reduction in computation time.
929
930 Energy drift in these SSD particle simulations was unnoticeable for
931 time steps up to three femtoseconds. A slight energy drift on the
932 order of 0.012 kcal/mol per nanosecond was observed at a time step of
933 four femtoseconds, and as expected, this drift increases dramatically
934 with increasing time step.
935
936
937 \subsection{\label{sec:extended}Extended Systems for other Ensembles}
938
939
940 {\sc oopse} implements a
941
942
943 \subsection{\label{oopseSec:noseHooverThermo}Nose-Hoover Thermostatting}
944
945 To mimic the effects of being in a constant temperature ({\sc nvt})
946 ensemble, {\sc oopse} uses the Nose-Hoover extended system
947 approach.\cite{Hoover85} In this method, the equations of motion for
948 the particle positions and velocities are
949 \begin{eqnarray}
950 \dot{{\bf r}} & = & {\bf v} \\
951 \dot{{\bf v}} & = & \frac{{\bf f}}{m} - \chi {\bf v}
952 \label{eq:nosehoovereom}
953 \end{eqnarray}
954
955 $\chi$ is an ``extra'' variable included in the extended system, and
956 it is propagated using the first order equation of motion
957 \begin{equation}
958 \dot{\chi} = \frac{1}{\tau_{T}} \left( \frac{T}{T_{target}} - 1 \right)
959 \label{eq:nosehooverext}
960 \end{equation}
961 where $T_{target}$ is the target temperature for the simulation, and
962 $\tau_{T}$ is a time constant for the thermostat.
963
964 To select the Nose-Hoover {\sc nvt} ensemble, the {\tt ensemble = NVT;}
965 command would be used in the simulation's {\sc bass} file. There is
966 some subtlety in choosing values for $\tau_{T}$, and it is usually set
967 to values of a few ps. Within a {\sc bass} file, $\tau_{T}$ could be
968 set to 1 ps using the {\tt tauThermostat = 1000; } command.
969
970 \subsection{\label{oopseSec:rattle}The {\sc rattle} Method for Bond
971 Constraints}
972
973 In order to satisfy the constraints of fixed bond lengths within {\sc
974 oopse}, we have implemented the {\sc rattle} algorithm of
975 Andersen.\cite{andersen83} The algorithm is a velocity verlet
976 formulation of the {\sc shake} method\cite{ryckaert77} of iteratively
977 solving the Lagrange multipliers of constraint. The system of lagrange
978 multipliers allows one to reformulate the equations of motion with
979 explicit constraint forces.\cite{fowles99:lagrange}
980
981 Consider a system described by coordinates $q_1$ and $q_2$ subject to an
982 equation of constraint:
983 \begin{equation}
984 \sigma(q_1, q_2,t) = 0
985 \label{oopseEq:lm1}
986 \end{equation}
987 The Lagrange formulation of the equations of motion can be written:
988 \begin{equation}
989 \delta\int_{t_1}^{t_2}L\, dt =
990 \int_{t_1}^{t_2} \sum_i \biggl [ \frac{\partial L}{\partial q_i}
991 - \frac{d}{dt}\biggl(\frac{\partial L}{\partial \dot{q}_i}
992 \biggr ) \biggr] \delta q_i \, dt = 0
993 \label{oopseEq:lm2}
994 \end{equation}
995 Here, $\delta q_i$ is not independent for each $q$, as $q_1$ and $q_2$
996 are linked by $\sigma$. However, $\sigma$ is fixed at any given
997 instant of time, giving:
998 \begin{align}
999 \delta\sigma &= \biggl( \frac{\partial\sigma}{\partial q_1} \delta q_1 %
1000 + \frac{\partial\sigma}{\partial q_2} \delta q_2 \biggr) = 0 \\
1001 %
1002 \frac{\partial\sigma}{\partial q_1} \delta q_1 &= %
1003 - \frac{\partial\sigma}{\partial q_2} \delta q_2 \\
1004 %
1005 \delta q_2 &= - \biggl(\frac{\partial\sigma}{\partial q_1} \bigg / %
1006 \frac{\partial\sigma}{\partial q_2} \biggr) \delta q_1
1007 \end{align}
1008 Substituted back into Eq.~\ref{oopseEq:lm2},
1009 \begin{equation}
1010 \int_{t_1}^{t_2}\biggl [ \biggl(\frac{\partial L}{\partial q_1}
1011 - \frac{d}{dt}\,\frac{\partial L}{\partial \dot{q}_1}
1012 \biggr)
1013 - \biggl( \frac{\partial L}{\partial q_1}
1014 - \frac{d}{dt}\,\frac{\partial L}{\partial \dot{q}_1}
1015 \biggr) \biggl(\frac{\partial\sigma}{\partial q_1} \bigg / %
1016 \frac{\partial\sigma}{\partial q_2} \biggr)\biggr] \delta q_1 \, dt = 0
1017 \label{oopseEq:lm3}
1018 \end{equation}
1019 Leading to,
1020 \begin{equation}
1021 \frac{\biggl(\frac{\partial L}{\partial q_1}
1022 - \frac{d}{dt}\,\frac{\partial L}{\partial \dot{q}_1}
1023 \biggr)}{\frac{\partial\sigma}{\partial q_1}} =
1024 \frac{\biggl(\frac{\partial L}{\partial q_2}
1025 - \frac{d}{dt}\,\frac{\partial L}{\partial \dot{q}_2}
1026 \biggr)}{\frac{\partial\sigma}{\partial q_2}}
1027 \label{oopseEq:lm4}
1028 \end{equation}
1029 This relation can only be statisfied, if both are equal to a single
1030 function $-\lambda(t)$,
1031 \begin{align}
1032 \frac{\biggl(\frac{\partial L}{\partial q_1}
1033 - \frac{d}{dt}\,\frac{\partial L}{\partial \dot{q}_1}
1034 \biggr)}{\frac{\partial\sigma}{\partial q_1}} &= -\lambda(t) \\
1035 %
1036 \frac{\partial L}{\partial q_1}
1037 - \frac{d}{dt}\,\frac{\partial L}{\partial \dot{q}_1} &=
1038 -\lambda(t)\,\frac{\partial\sigma}{\partial q_1} \\
1039 %
1040 \frac{\partial L}{\partial q_1}
1041 - \frac{d}{dt}\,\frac{\partial L}{\partial \dot{q}_1}
1042 + \mathcal{G}_i &= 0
1043 \end{align}
1044 Where $\mathcal{G}_i$, the force of constraint on $i$, is:
1045 \begin{equation}
1046 \mathcal{G}_i = \lambda(t)\,\frac{\partial\sigma}{\partial q_1}
1047 \label{oopseEq:lm5}
1048 \end{equation}
1049
1050 In a simulation, this would involve the solution of a set of $(m + n)$
1051 number of equations. Where $m$ is the number of constraints, and $n$
1052 is the number of constrained coordinates. In practice, this is not
1053 done, as the matrix inversion necessary to solve the system of
1054 equations would be very time consuming to solve. Additionally, the
1055 numerical error in the solution of the set of $\lambda$'s would be
1056 compounded by the error inherent in propagating by the Velocity Verlet
1057 algorithm ($\Delta t^4$). The Verlet propagation error is negligible
1058 in an unconstrained system, as one is interested in the statistics of
1059 the run, and not that the run be numerically exact to the ``true''
1060 integration. This relates back to the ergodic hypothesis that a time
1061 integral of a valid trajectory will still give the correct ensemble
1062 average. However, in the case of constraints, if the equations of
1063 motion leave the ``true'' trajectory, they are departing from the
1064 constrained surface. The method that is used, is to iteratively solve
1065 for $\lambda(t)$ at each time step.
1066
1067 In {\sc rattle} the equations of motion are modified subject to the
1068 following two constraints:
1069 \begin{align}
1070 \sigma_{ij}[\mathbf{r}(t)] \equiv
1071 [ \mathbf{r}_i(t) - \mathbf{r}_j(t)]^2 - d_{ij}^2 &= 0 %
1072 \label{oopseEq:c1} \\
1073 %
1074 [\mathbf{\dot{r}}_i(t) - \mathbf{\dot{r}}_j(t)] \cdot
1075 [\mathbf{r}_i(t) - \mathbf{r}_j(t)] &= 0 \label{oopseEq:c2}
1076 \end{align}
1077 Eq.~\ref{oopseEq:c1} is the set of bond constraints, where $d_{ij}$ is
1078 the constrained distance between atom $i$ and
1079 $j$. Eq.~\ref{oopseEq:c2} constrains the velocities of $i$ and $j$ to
1080 be perpendicular to the bond vector, so that the bond can neither grow
1081 nor shrink. The constrained dynamics equations become:
1082 \begin{equation}
1083 m_i \mathbf{\ddot{r}}_i = \mathbf{F}_i + \mathbf{\mathcal{G}}_i
1084 \label{oopseEq:r1}
1085 \end{equation}
1086 Where,$\mathbf{\mathcal{G}}_i$ are the forces of constraint on $i$,
1087 and are defined:
1088 \begin{equation}
1089 \mathbf{\mathcal{G}}_i = - \sum_j \lambda_{ij}(t)\,\nabla \sigma_{ij}
1090 \label{oopseEq:r2}
1091 \end{equation}
1092
1093 In Velocity Verlet, if $\Delta t = h$, the propagation can be written:
1094 \begin{align}
1095 \mathbf{r}_i(t+h) &=
1096 \mathbf{r}_i(t) + h\mathbf{\dot{r}}(t) +
1097 \frac{h^2}{2m_i}\,\Bigl[ \mathbf{F}_i(t) +
1098 \mathbf{\mathcal{G}}_{Ri}(t) \Bigr] \label{oopseEq:vv1} \\
1099 %
1100 \mathbf{\dot{r}}_i(t+h) &=
1101 \mathbf{\dot{r}}_i(t) + \frac{h}{2m_i}
1102 \Bigl[ \mathbf{F}_i(t) + \mathbf{\mathcal{G}}_{Ri}(t) +
1103 \mathbf{F}_i(t+h) + \mathbf{\mathcal{G}}_{Vi}(t+h) \Bigr] %
1104 \label{oopseEq:vv2}
1105 \end{align}
1106 Where:
1107 \begin{align}
1108 \mathbf{\mathcal{G}}_{Ri}(t) &=
1109 -2 \sum_j \lambda_{Rij}(t) \mathbf{r}_{ij}(t) \\
1110 %
1111 \mathbf{\mathcal{G}}_{Vi}(t+h) &=
1112 -2 \sum_j \lambda_{Vij}(t+h) \mathbf{r}(t+h)
1113 \end{align}
1114 Next, define:
1115 \begin{align}
1116 g_{ij} &= h \lambda_{Rij}(t) \\
1117 k_{ij} &= h \lambda_{Vij}(t+h) \\
1118 \mathbf{q}_i &= \mathbf{\dot{r}}_i(t) + \frac{h}{2m_i} \mathbf{F}_i(t)
1119 - \frac{1}{m_i}\sum_j g_{ij}\mathbf{r}_{ij}(t)
1120 \end{align}
1121 Using these definitions, Eq.~\ref{oopseEq:vv1} and \ref{oopseEq:vv2}
1122 can be rewritten as,
1123 \begin{align}
1124 \mathbf{r}_i(t+h) &= \mathbf{r}_i(t) + h \mathbf{q}_i \\
1125 %
1126 \mathbf{\dot{r}}(t+h) &= \mathbf{q}_i + \frac{h}{2m_i}\mathbf{F}_i(t+h)
1127 -\frac{1}{m_i}\sum_j k_{ij} \mathbf{r}_{ij}(t+h)
1128 \end{align}
1129
1130 To integrate the equations of motion, the {\sc rattle} algorithm first
1131 solves for $\mathbf{r}(t+h)$. Let,
1132 \begin{equation}
1133 \mathbf{q}_i = \mathbf{\dot{r}}(t) + \frac{h}{2m_i}\mathbf{F}_i(t)
1134 \end{equation}
1135 Here $\mathbf{q}_i$ corresponds to an initial unconstrained move. Next
1136 pick a constraint $j$, and let,
1137 \begin{equation}
1138 \mathbf{s} = \mathbf{r}_i(t) + h\mathbf{q}_i(t)
1139 - \mathbf{r}_j(t) + h\mathbf{q}_j(t)
1140 \label{oopseEq:ra1}
1141 \end{equation}
1142 If
1143 \begin{equation}
1144 \Big| |\mathbf{s}|^2 - d_{ij}^2 \Big| > \text{tolerance},
1145 \end{equation}
1146 then the constraint is unsatisfied, and corrections are made to the
1147 positions. First we define a test corrected configuration as,
1148 \begin{align}
1149 \mathbf{r}_i^T(t+h) = \mathbf{r}_i(t) + h\biggl[\mathbf{q}_i -
1150 g_{ij}\,\frac{\mathbf{r}_{ij}(t)}{m_i} \biggr] \\
1151 %
1152 \mathbf{r}_j^T(t+h) = \mathbf{r}_j(t) + h\biggl[\mathbf{q}_j +
1153 g_{ij}\,\frac{\mathbf{r}_{ij}(t)}{m_j} \biggr]
1154 \end{align}
1155 And we chose $g_{ij}$ such that, $|\mathbf{r}_i^T - \mathbf{r}_j^T|^2
1156 = d_{ij}^2$. Solving the quadratic for $g_{ij}$ we obtain the
1157 approximation,
1158 \begin{equation}
1159 g_{ij} = \frac{(s^2 - d^2)}{2h[\mathbf{s}\cdot\mathbf{r}_{ij}(t)]
1160 (\frac{1}{m_i} + \frac{1}{m_j})}
1161 \end{equation}
1162 Although not an exact solution for $g_{ij}$, as this is an iterative
1163 scheme overall, the eventual solution will converge. With a trial
1164 $g_{ij}$, the new $\mathbf{q}$'s become,
1165 \begin{align}
1166 \mathbf{q}_i &= \mathbf{q}^{\text{old}}_i - g_{ij}\,
1167 \frac{\mathbf{r}_{ij}(t)}{m_i} \\
1168 %
1169 \mathbf{q}_j &= \mathbf{q}^{\text{old}}_j + g_{ij}\,
1170 \frac{\mathbf{r}_{ij}(t)}{m_j}
1171 \end{align}
1172 The whole algorithm is then repeated from Eq.~\ref{oopseEq:ra1} until
1173 all constraints are satisfied.
1174
1175 The second step of {\sc rattle}, is to then update the velocities. The
1176 step starts with,
1177 \begin{equation}
1178 \mathbf{\dot{r}}_i(t+h) = \mathbf{q}_i + \frac{h}{2m_i}\mathbf{F}_i(t+h)
1179 \end{equation}
1180 Next we pick a constraint $j$, and calculate the dot product $\ell$.
1181 \begin{equation}
1182 \ell = \mathbf{r}_{ij}(t+h) \cdot \mathbf{\dot{r}}_{ij}(t+h)
1183 \label{oopseEq:rv1}
1184 \end{equation}
1185 Here if constraint Eq.~\ref{oopseEq:c2} holds, $\ell$ should be
1186 zero. Therefore if $\ell$ is greater than some tolerance, then
1187 corrections are made to the $i$ and $j$ velocities.
1188 \begin{align}
1189 \mathbf{\dot{r}}_i^T &= \mathbf{\dot{r}}_i(t+h) - k_{ij}
1190 \frac{\mathbf{\dot{r}}_{ij}(t+h)}{m_i} \\
1191 %
1192 \mathbf{\dot{r}}_j^T &= \mathbf{\dot{r}}_j(t+h) + k_{ij}
1193 \frac{\mathbf{\dot{r}}_{ij}(t+h)}{m_j}
1194 \end{align}
1195 Like in the previous step, we select a value for $k_{ij}$ such that
1196 $\ell$ is zero.
1197 \begin{equation}
1198 k_{ij} = \frac{\ell}{d^2_{ij}(\frac{1}{m_i} + \frac{1}{m_j})}
1199 \end{equation}
1200 The test velocities, $\mathbf{\dot{r}}^T_i$ and
1201 $\mathbf{\dot{r}}^T_j$, then replace their respective velocities, and
1202 the algorithm is iterated from Eq.~\ref{oopseEq:rv1} until all
1203 constraints are satisfied.
1204
1205
1206 \subsection{\label{oopseSec:zcons}Z-Constraint Method}
1207
1208 Based on the fluctuation-dissipation theorem, a force auto-correlation
1209 method was developed by Roux and Karplus to investigate the dynamics
1210 of ions inside ion channels.\cite{Roux91} The time-dependent friction
1211 coefficient can be calculated from the deviation of the instantaneous
1212 force from its mean force.
1213 \begin{equation}
1214 \xi(z,t)=\langle\delta F(z,t)\delta F(z,0)\rangle/k_{B}T
1215 \end{equation}
1216 where%
1217 \begin{equation}
1218 \delta F(z,t)=F(z,t)-\langle F(z,t)\rangle
1219 \end{equation}
1220
1221
1222 If the time-dependent friction decays rapidly, the static friction
1223 coefficient can be approximated by
1224 \begin{equation}
1225 \xi^{static}(z)=\int_{0}^{\infty}\langle\delta F(z,t)\delta F(z,0)\rangle dt
1226 \end{equation}
1227 Therefore, the diffusion constant can then be estimated by
1228 \begin{equation}
1229 D(z)=\frac{k_{B}T}{\xi^{static}(z)}=\frac{(k_{B}T)^{2}}{\int_{0}^{\infty
1230 }\langle\delta F(z,t)\delta F(z,0)\rangle dt}%
1231 \end{equation}
1232
1233 The Z-Constraint method, which fixes the z coordinates of the
1234 molecules with respect to the center of the mass of the system, has
1235 been a method suggested to obtain the forces required for the force
1236 auto-correlation calculation.\cite{Marrink94} However, simply resetting the
1237 coordinate will move the center of the mass of the whole system. To
1238 avoid this problem, a new method was used in {\sc oopse}. Instead of
1239 resetting the coordinate, we reset the forces of z-constraint
1240 molecules as well as subtract the total constraint forces from the
1241 rest of the system after force calculation at each time step.
1242 \begin{align}
1243 F_{\alpha i}&=0\\
1244 V_{\alpha i}&=V_{\alpha i}-\frac{\sum\limits_{i}M_{_{\alpha i}}V_{\alpha i}}{\sum\limits_{i}M_{_{\alpha i}}}\\
1245 F_{\alpha i}&=F_{\alpha i}-\frac{M_{_{\alpha i}}}{\sum\limits_{\alpha}\sum\limits_{i}M_{_{\alpha i}}}\sum\limits_{\beta}F_{\beta}\\
1246 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}}}
1247 \end{align}
1248
1249 At the very beginning of the simulation, the molecules may not be at their
1250 constrained positions. To move a z-constrained molecule to its specified
1251 position, a simple harmonic potential is used
1252 \begin{equation}
1253 U(t)=\frac{1}{2}k_{\text{Harmonic}}(z(t)-z_{\text{cons}})^{2}%
1254 \end{equation}
1255 where $k_{\text{Harmonic}}$ is the harmonic force constant, $z(t)$ is the
1256 current $z$ coordinate of the center of mass of the constrained molecule, and
1257 $z_{\text{cons}}$ is the constrained position. The harmonic force operating
1258 on the z-constrained molecule at time $t$ can be calculated by
1259 \begin{equation}
1260 F_{z_{\text{Harmonic}}}(t)=-\frac{\partial U(t)}{\partial z(t)}=
1261 -k_{\text{Harmonic}}(z(t)-z_{\text{cons}})
1262 \end{equation}
1263
1264 \section{\label{oopseSec:props}Trajectory Analysis}
1265
1266 \subsection{\label{oopseSec:staticProps}Static Property Analysis}
1267
1268 The static properties of the trajectories are analyzed with the
1269 program \texttt{staticProps}. The code is capable of calculating a
1270 number of pair correlations between species A and B. Some of which
1271 only apply to directional entities. The summary of pair correlations
1272 can be found in Table~\ref{oopseTb:gofrs}
1273
1274 \begin{table}
1275 \caption[The list of pair correlations in \texttt{staticProps}]{The different pair correlations in \texttt{staticProps} along with whether atom A or B must be directional.}
1276 \label{oopseTb:gofrs}
1277 \begin{center}
1278 \begin{tabular}{|l|c|c|}
1279 \hline
1280 Name & Equation & Directional Atom \\ \hline
1281 $g_{\text{AB}}(r)$ & Eq.~\ref{eq:gofr} & neither \\ \hline
1282 $g_{\text{AB}}(r, \cos \theta)$ & Eq.~\ref{eq:gofrCosTheta} & A \\ \hline
1283 $g_{\text{AB}}(r, \cos \omega)$ & Eq.~\ref{eq:gofrCosOmega} & both \\ \hline
1284 $g_{\text{AB}}(x, y, z)$ & Eq.~\ref{eq:gofrXYZ} & neither \\ \hline
1285 $\langle \cos \omega \rangle_{\text{AB}}(r)$ & Eq.~\ref{eq:cosOmegaOfR} &%
1286 both \\ \hline
1287 \end{tabular}
1288 \end{center}
1289 \end{table}
1290
1291 The first pair correlation, $g_{\text{AB}}(r)$, is defined as follows:
1292 \begin{equation}
1293 g_{\text{AB}}(r) = \frac{V}{N_{\text{A}}N_{\text{B}}}\langle %%
1294 \sum_{i \in \text{A}} \sum_{j \in \text{B}} %%
1295 \delta( r - |\mathbf{r}_{ij}|) \rangle \label{eq:gofr}
1296 \end{equation}
1297 Where $\mathbf{r}_{ij}$ is the vector
1298 \begin{equation*}
1299 \mathbf{r}_{ij} = \mathbf{r}_j - \mathbf{r}_i \notag
1300 \end{equation*}
1301 and $\frac{V}{N_{\text{A}}N_{\text{B}}}$ normalizes the average over
1302 the expected pair density at a given $r$.
1303
1304 The next two pair correlations, $g_{\text{AB}}(r, \cos \theta)$ and
1305 $g_{\text{AB}}(r, \cos \omega)$, are similar in that they are both two
1306 dimensional histograms. Both use $r$ for the primary axis then a
1307 $\cos$ for the secondary axis ($\cos \theta$ for
1308 Eq.~\ref{eq:gofrCosTheta} and $\cos \omega$ for
1309 Eq.~\ref{eq:gofrCosOmega}). This allows for the investigator to
1310 correlate alignment on directional entities. $g_{\text{AB}}(r, \cos
1311 \theta)$ is defined as follows:
1312 \begin{equation}
1313 g_{\text{AB}}(r, \cos \theta) = \frac{V}{N_{\text{A}}N_{\text{B}}}\langle
1314 \sum_{i \in \text{A}} \sum_{j \in \text{B}}
1315 \delta( \cos \theta - \cos \theta_{ij})
1316 \delta( r - |\mathbf{r}_{ij}|) \rangle
1317 \label{eq:gofrCosTheta}
1318 \end{equation}
1319 Where
1320 \begin{equation*}
1321 \cos \theta_{ij} = \mathbf{\hat{i}} \cdot \mathbf{\hat{r}}_{ij}
1322 \end{equation*}
1323 Here $\mathbf{\hat{i}}$ is the unit directional vector of species $i$
1324 and $\mathbf{\hat{r}}_{ij}$ is the unit vector associated with vector
1325 $\mathbf{r}_{ij}$.
1326
1327 The second two dimensional histogram is of the form:
1328 \begin{equation}
1329 g_{\text{AB}}(r, \cos \omega) =
1330 \frac{V}{N_{\text{A}}N_{\text{B}}}\langle
1331 \sum_{i \in \text{A}} \sum_{j \in \text{B}}
1332 \delta( \cos \omega - \cos \omega_{ij})
1333 \delta( r - |\mathbf{r}_{ij}|) \rangle \label{eq:gofrCosOmega}
1334 \end{equation}
1335 Here
1336 \begin{equation*}
1337 \cos \omega_{ij} = \mathbf{\hat{i}} \cdot \mathbf{\hat{j}}
1338 \end{equation*}
1339 Again, $\mathbf{\hat{i}}$ and $\mathbf{\hat{j}}$ are the unit
1340 directional vectors of species $i$ and $j$.
1341
1342 The static analysis code is also cable of calculating a three
1343 dimensional pair correlation of the form:
1344 \begin{equation}\label{eq:gofrXYZ}
1345 g_{\text{AB}}(x, y, z) =
1346 \frac{V}{N_{\text{A}}N_{\text{B}}}\langle
1347 \sum_{i \in \text{A}} \sum_{j \in \text{B}}
1348 \delta( x - x_{ij})
1349 \delta( y - y_{ij})
1350 \delta( z - z_{ij}) \rangle
1351 \end{equation}
1352 Where $x_{ij}$, $y_{ij}$, and $z_{ij}$ are the $x$, $y$, and $z$
1353 components respectively of vector $\mathbf{r}_{ij}$.
1354
1355 The final pair correlation is similar to
1356 Eq.~\ref{eq:gofrCosOmega}. $\langle \cos \omega
1357 \rangle_{\text{AB}}(r)$ is calculated in the following way:
1358 \begin{equation}\label{eq:cosOmegaOfR}
1359 \langle \cos \omega \rangle_{\text{AB}}(r) =
1360 \langle \sum_{i \in \text{A}} \sum_{j \in \text{B}}
1361 (\cos \omega_{ij}) \delta( r - |\mathbf{r}_{ij}|) \rangle
1362 \end{equation}
1363 Here $\cos \omega_{ij}$ is defined in the same way as in
1364 Eq.~\ref{eq:gofrCosOmega}. This equation is a single dimensional pair
1365 correlation that gives the average correlation of two directional
1366 entities as a function of their distance from each other.
1367
1368 \subsection{\label{dynamicProps}Dynamic Property Analysis}
1369
1370 The dynamic properties of a trajectory are calculated with the program
1371 \texttt{dynamicProps}. The program calculates the following properties:
1372 \begin{gather}
1373 \langle | \mathbf{r}(t) - \mathbf{r}(0) |^2 \rangle \label{eq:rms}\\
1374 \langle \mathbf{v}(t) \cdot \mathbf{v}(0) \rangle \label{eq:velCorr} \\
1375 \langle \mathbf{j}(t) \cdot \mathbf{j}(0) \rangle \label{eq:angularVelCorr}
1376 \end{gather}
1377
1378 Eq.~\ref{eq:rms} is the root mean square displacement function. Which
1379 allows one to observe the average displacement of an atom as a
1380 function of time. The quantity is useful when calculating diffusion
1381 coefficients because of the Einstein Relation, which is valid at long
1382 times.\cite{allen87:csl}
1383 \begin{equation}
1384 2tD = \langle | \mathbf{r}(t) - \mathbf{r}(0) |^2 \rangle
1385 \label{oopseEq:einstein}
1386 \end{equation}
1387
1388 Eq.~\ref{eq:velCorr} and \ref{eq:angularVelCorr} are the translational
1389 velocity and angular velocity correlation functions respectively. The
1390 latter is only applicable to directional species in the
1391 simulation. The velocity autocorrelation functions are useful when
1392 determining vibrational information about the system of interest.
1393
1394 \section{\label{oopseSec:design}Program Design}
1395
1396 \subsection{\label{sec:architecture} {\sc oopse} Architecture}
1397
1398 The core of OOPSE is divided into two main object libraries:
1399 \texttt{libBASS} and \texttt{libmdtools}. \texttt{libBASS} is the
1400 library developed around the parsing engine and \texttt{libmdtools}
1401 is the software library developed around the simulation engine. These
1402 two libraries are designed to encompass all the basic functions and
1403 tools that {\sc oopse} provides. Utility programs, such as the
1404 property analyzers, need only link against the software libraries to
1405 gain access to parsing, force evaluation, and input / output
1406 routines.
1407
1408 Contained in \texttt{libBASS} are all the routines associated with
1409 reading and parsing the \texttt{.bass} input files. Given a
1410 \texttt{.bass} file, \texttt{libBASS} will open it and any associated
1411 \texttt{.mdl} files; then create structures in memory that are
1412 templates of all the molecules specified in the input files. In
1413 addition, any simulation parameters set in the \texttt{.bass} file
1414 will be placed in a structure for later query by the controlling
1415 program.
1416
1417 Located in \texttt{libmdtools} are all other routines necessary to a
1418 Molecular Dynamics simulation. The library uses the main data
1419 structures returned by \texttt{libBASS} to initialize the various
1420 parts of the simulation: the atom structures and positions, the force
1421 field, the integrator, \emph{et cetera}. After initialization, the
1422 library can be used to perform a variety of tasks: integrate a
1423 Molecular Dynamics trajectory, query phase space information from a
1424 specific frame of a completed trajectory, or even recalculate force or
1425 energetic information about specific frames from a completed
1426 trajectory.
1427
1428 With these core libraries in place, several programs have been
1429 developed to utilize the routines provided by \texttt{libBASS} and
1430 \texttt{libmdtools}. The main program of the package is \texttt{oopse}
1431 and the corresponding parallel version \texttt{oopse\_MPI}. These two
1432 programs will take the \texttt{.bass} file, and create (and integrate)
1433 the simulation specified in the script. The two analysis programs
1434 \texttt{staticProps} and \texttt{dynamicProps} utilize the core
1435 libraries to initialize and read in trajectories from previously
1436 completed simulations, in addition to the ability to use functionality
1437 from \texttt{libmdtools} to recalculate forces and energies at key
1438 frames in the trajectories. Lastly, the family of system building
1439 programs (Sec.~\ref{oopseSec:initCoords}) also use the libraries to
1440 store and output the system configurations they create.
1441
1442 \subsection{\label{oopseSec:parallelization} Parallelization of {\sc oopse}}
1443
1444 Although processor power is continually growing roughly following
1445 Moore's Law, it is still unreasonable to simulate systems of more then
1446 a 1000 atoms on a single processor. To facilitate study of larger
1447 system sizes or smaller systems on long time scales in a reasonable
1448 period of time, parallel methods were developed allowing multiple
1449 CPU's to share the simulation workload. Three general categories of
1450 parallel decomposition methods have been developed including atomic,
1451 spatial and force decomposition methods.
1452
1453 Algorithmically simplest of the three methods is atomic decomposition
1454 where N particles in a simulation are split among P processors for the
1455 duration of the simulation. Computational cost scales as an optimal
1456 $O(N/P)$ for atomic decomposition. Unfortunately all processors must
1457 communicate positions and forces with all other processors at every
1458 force evaluation, leading communication costs to scale as an
1459 unfavorable $O(N)$, \emph{independent of the number of processors}. This
1460 communication bottleneck led to the development of spatial and force
1461 decomposition methods in which communication among processors scales
1462 much more favorably. Spatial or domain decomposition divides the
1463 physical spatial domain into 3D boxes in which each processor is
1464 responsible for calculation of forces and positions of particles
1465 located in its box. Particles are reassigned to different processors
1466 as they move through simulation space. To calculate forces on a given
1467 particle, a processor must know the positions of particles within some
1468 cutoff radius located on nearby processors instead of the positions of
1469 particles on all processors. Both communication between processors and
1470 computation scale as $O(N/P)$ in the spatial method. However, spatial
1471 decomposition adds algorithmic complexity to the simulation code and
1472 is not very efficient for small N since the overall communication
1473 scales as the surface to volume ratio $(N/P)^{2/3}$ in three
1474 dimensions.
1475
1476 The parallelization method used in {\sc oopse} is the force
1477 decomposition method. Force decomposition assigns particles to
1478 processors based on a block decomposition of the force
1479 matrix. Processors are split into an optimally square grid forming row
1480 and column processor groups. Forces are calculated on particles in a
1481 given row by particles located in that processors column
1482 assignment. Force decomposition is less complex to implement than the
1483 spatial method but still scales computationally as $O(N/P)$ and scales
1484 as $O(N/\sqrt{P})$ in communication cost. Plimpton has also found that
1485 force decompositions scale more favorably than spatial decompositions
1486 for systems up to 10,000 atoms and favorably compete with spatial
1487 methods up to 100,000 atoms.\cite{plimpton95}
1488
1489 \subsection{\label{oopseSec:memAlloc}Memory Issues in Trajectory Analysis}
1490
1491 For large simulations, the trajectory files can sometimes reach sizes
1492 in excess of several gigabytes. In order to effectively analyze that
1493 amount of data, two memory management schemes have been devised for
1494 \texttt{staticProps} and for \texttt{dynamicProps}. The first scheme,
1495 developed for \texttt{staticProps}, is the simplest. As each frame's
1496 statistics are calculated independent of each other, memory is
1497 allocated for each frame, then freed once correlation calculations are
1498 complete for the snapshot. To prevent multiple passes through a
1499 potentially large file, \texttt{staticProps} is capable of calculating
1500 all requested correlations per frame with only a single pair loop in
1501 each frame and a single read of the file.
1502
1503 The second, more advanced memory scheme, is used by
1504 \texttt{dynamicProps}. Here, the program must have multiple frames in
1505 memory to calculate time dependent correlations. In order to prevent a
1506 situation where the program runs out of memory due to large
1507 trajectories, the user is able to specify that the trajectory be read
1508 in blocks. The number of frames in each block is specified by the
1509 user, and upon reading a block of the trajectory,
1510 \texttt{dynamicProps} will calculate all of the time correlation frame
1511 pairs within the block. After in-block correlations are complete, a
1512 second block of the trajectory is read, and the cross correlations are
1513 calculated between the two blocks. this second block is then freed and
1514 then incremented and the process repeated until the end of the
1515 trajectory. Once the end is reached, the first block is freed then
1516 incremented, and the again the internal time correlations are
1517 calculated. The algorithm with the second block is then repeated with
1518 the new origin block, until all frame pairs have been correlated in
1519 time. This process is illustrated in
1520 Fig.~\ref{oopseFig:dynamicPropsMemory}.
1521
1522 \begin{figure}
1523 \centering
1524 \includegraphics[width=\linewidth]{dynamicPropsMem.eps}
1525 \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.}
1526 \label{oopseFig:dynamicPropsMemory}
1527 \end{figure}
1528
1529 \section{\label{oopseSec:conclusion}Conclusion}
1530
1531 We have presented the design and implementation of our open source
1532 simulation package {\sc oopse}. The package offers novel capabilities
1533 to the field of Molecular Dynamics simulation packages in the form of
1534 dipolar force fields, and symplectic integration of rigid body
1535 dynamics. It is capable of scaling across multiple processors through
1536 the use of force based decomposition using MPI. It also implements
1537 several advanced integrators allowing the end user control over
1538 temperature and pressure. In addition, it is capable of integrating
1539 constrained dynamics through both the {\sc rattle} algorithm and the
1540 z-constraint method.
1541
1542 These features are all brought together in a single open-source
1543 program. Allowing researchers to not only benefit from
1544 {\sc oopse}, but also contribute to {\sc oopse}'s development as
1545 well.Documentation and source code for {\sc oopse} can be downloaded
1546 from \texttt{http://www.openscience.org/oopse/}.
1547