ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/mattDisertation/oopse.tex
Revision: 1044
Committed: Mon Feb 9 21:44:01 2004 UTC (20 years, 4 months ago) by mmeineke
Content type: application/x-tex
File size: 48771 byte(s)
Log Message:
added some OOPSE sections

File Contents

# Content
1 \documentclass[11pt]{article}
2 \usepackage{amsmath}
3 \usepackage{amssymb}
4 \usepackage{endfloat}
5 \usepackage{berkeley}
6 \usepackage{listings}
7 \usepackage{epsf}
8 \usepackage[ref]{overcite}
9 \usepackage{setspace}
10 \usepackage{tabularx}
11 \pagestyle{plain}
12 \pagenumbering{arabic}
13 \oddsidemargin 0.0cm \evensidemargin 0.0cm
14 \topmargin -21pt \headsep 10pt
15 \textheight 9.0in \textwidth 6.5in
16 \brokenpenalty=10000
17 \renewcommand{\baselinestretch}{1.2}
18 \renewcommand\citemid{\ } % no comma in optional reference note
19
20 \begin{document}
21 \lstset{language=C,float,frame=tblr,frameround=tttt}
22 \renewcommand{\lstlistingname}{Scheme}
23 \title{{\sc oopse}: An Open Source Object-Oriented Parallel Simulation
24 Engine for Molecular Dynamics}
25
26 \author{Matthew A. Meineke, Charles F. Vardeman II, Teng Lin, Christopher J. Fennell and J. Daniel Gezelter\\
27 Department of Chemistry and Biochemistry\\
28 University of Notre Dame\\
29 Notre Dame, Indiana 46556}
30
31 \date{\today}
32 \maketitle
33
34 \begin{abstract}
35 We detail the capabilities of a new open-source parallel simulation
36 package ({\sc oopse}) that can perform molecular dynamics simulations
37 on atom types that are missing from other popular packages. In
38 particular, {\sc oopse} is capable of performing orientational
39 dynamics on dipolar systems, and it can handle simulations of metallic
40 systems using the embedded atom method ({\sc eam}).
41 \end{abstract}
42
43 \newpage
44
45 \section{\label{sec:intro}Introduction}
46
47 \begin{itemize}
48
49 \item Need for package / Niche to fill
50
51 \item Design Goal
52
53 \item Open Source
54
55 \item Discussion of Paper Layout
56
57 \end{itemize}
58
59 \section{\label{sec:empiricalEnergy}The Empirical Energy Functions}
60
61 \subsection{\label{sec:atomsMolecules}Atoms, Molecules and Rigid Bodies}
62
63 The basic unit of an {\sc oopse} simulation is the atom. The
64 parameters describing the atom are generalized to make the atom as
65 flexible a representation as possible. They may represent specific
66 atoms of an element, or be used for collections of atoms such as
67 methyl and carbonyl groups. The atoms are also capable of having
68 directional components associated with them (\emph{e.g.}~permanent
69 dipoles). Charges on atoms are not currently supported by {\sc oopse}.
70
71 \begin{lstlisting}[caption={[Specifier for molecules and atoms] A sample specification of the simple Ar molecule},label=sch:AtmMole]
72 molecule{
73 name = "Ar";
74 nAtoms = 1;
75 atom[0]{
76 type="Ar";
77 position( 0.0, 0.0, 0.0 );
78 }
79 }
80 \end{lstlisting}
81
82 Atoms can be collected into secondary srtructures such as rigid bodies
83 or molecules. The molecule is a way for {\sc oopse} to keep track of
84 the atoms in a simulation in logical manner. Molecular units store the
85 identities of all the atoms associated with themselves, and are
86 responsible for the evaluation of their own internal interactions
87 (\emph{i.e.}~bonds, bends, and torsions). Scheme \ref{sch:AtmMole}
88 shws how one creates a molecule in the \texttt{.mdl} files. The
89 position of the atoms given in the declaration are relative to the
90 origin of the molecule, and is used when creating a system containing
91 the molecule.
92
93 As stated previously, one of the features that sets {\sc oopse} apart
94 from most of the current molecular simulation packages is the ability
95 to handle rigid body dynamics. Rigid bodies are non-spherical
96 particles or collections of particles that have a constant internal
97 potential and move collectively.\cite{Goldstein01} They are not
98 included in most simulation packages because of the requirement to
99 propagate the orientational degrees of freedom. Until recently,
100 integrators which propagate orientational motion have been lacking.
101
102 Moving a rigid body involves determination of both the force and
103 torque applied by the surroundings, which directly affect the
104 translational and rotational motion in turn. In order to accumulate
105 the total force on a rigid body, the external forces and torques must
106 first be calculated for all the internal particles. The total force on
107 the rigid body is simply the sum of these external forces.
108 Accumulation of the total torque on the rigid body is more complex
109 than the force in that it is the torque applied on the center of mass
110 that dictates rotational motion. The torque on rigid body {\it i} is
111 \begin{equation}
112 \boldsymbol{\tau}_i=
113 \sum_{a}(\mathbf{r}_{ia}-\mathbf{r}_i)\times \mathbf{f}_{ia}
114 + \boldsymbol{\tau}_{ia},
115 \label{eq:torqueAccumulate}
116 \end{equation}
117 where $\boldsymbol{\tau}_i$ and $\mathbf{r}_i$ are the torque on and
118 position of the center of mass respectively, while $\mathbf{f}_{ia}$,
119 $\mathbf{r}_{ia}$, and $\boldsymbol{\tau}_{ia}$ are the force on,
120 position of, and torque on the component particles of the rigid body.
121
122 The summation of the total torque is done in the body fixed axis of
123 the rigid body. In order to move between the space fixed and body
124 fixed coordinate axes, parameters describing the orientation must be
125 maintained for each rigid body. At a minimum, the rotation matrix
126 (\textbf{A}) can be described by the three Euler angles ($\phi,
127 \theta,$ and $\psi$), where the elements of \textbf{A} are composed of
128 trigonometric operations involving $\phi, \theta,$ and
129 $\psi$.\cite{Goldstein01} In order to avoid numerical instabilities
130 inherent in using the Euler angles, the four parameter ``quaternion''
131 scheme is often used. The elements of \textbf{A} can be expressed as
132 arithmetic operations involving the four quaternions ($q_0, q_1, q_2,$
133 and $q_3$).\cite{allen87:csl} Use of quaternions also leads to
134 performance enhancements, particularly for very small
135 systems.\cite{Evans77}
136
137 {\sc oopse} utilizes a relatively new scheme that propagates the
138 entire nine parameter rotation matrix internally. Further discussion
139 on this choice can be found in Sec.~\ref{sec:integrate}. An example
140 definition of a riged body can be seen in Scheme
141 \ref{sch:rigidBody}. The positions in the atom definitions are the
142 placements of the atoms relative to the origin of the rigid body,
143 which itself has a position relative to the origin of the molecule.
144
145 \begin{lstlisting}[caption={[Defining rigid bodies]A sample definition of a rigid body},label={sch:rigidBody}]
146 molecule{
147 name = "TIP3P_water";
148 nRigidBodies = 1;
149 rigidBody[0]{
150 nAtoms = 3;
151 atom[0]{
152 type = "O_TIP3P";
153 position( 0.0, 0.0, -0.06556 );
154 }
155 atom[1]{
156 type = "H_TIP3P";
157 position( 0.0, 0.75695, 0.52032 );
158 }
159 atom[2]{
160 type = "H_TIP3P";
161 position( 0.0, -0.75695, 0.52032 );
162 }
163 position( 0.0, 0.0, 0.0 );
164 orientation( 0.0, 0.0, 1.0 );
165 }
166 }
167 \end{lstlisting}
168
169 \subsection{\label{sec:LJPot}The Lennard Jones Potential}
170
171 The most basic force field implemented in {\sc oopse} is the
172 Lennard-Jones potential, which mimics the van der Waals interaction at
173 long distances, and uses an empirical repulsion at short
174 distances. The Lennard-Jones potential is given by:
175 \begin{equation}
176 V_{\text{LJ}}(r_{ij}) =
177 4\epsilon_{ij} \biggl[
178 \biggl(\frac{\sigma_{ij}}{r_{ij}}\biggr)^{12}
179 - \biggl(\frac{\sigma_{ij}}{r_{ij}}\biggr)^{6}
180 \biggr]
181 \label{eq:lennardJonesPot}
182 \end{equation}
183 Where $r_{ij}$ is the distance between particles $i$ and $j$,
184 $\sigma_{ij}$ scales the length of the interaction, and
185 $\epsilon_{ij}$ scales the well depth of the potential. Scheme
186 \ref{sch:LJFF} gives and example partial \texttt{.bass} file that
187 shows a system of 108 Ar particles simulated with the Lennard-Jones
188 force field.
189
190 \begin{lstlisting}[caption={[Invocation of the Lennard-Jones force field] A sample system using the Lennard-Jones force field.},label={sch:LJFF}]
191
192 /*
193 * The Ar molecule is specified
194 * external to the.bass file
195 */
196
197 #include "argon.mdl"
198
199 nComponents = 1;
200 component{
201 type = "Ar";
202 nMol = 108;
203 }
204
205 /*
206 * The initial configuration is generated
207 * before the simulation is invoked.
208 */
209
210 initialConfig = "./argon.init";
211
212 forceField = "LJ";
213 \end{lstlisting}
214
215 Because this potential is calculated between all pairs, the force
216 evaluation can become computationally expensive for large systems. To
217 keep the pair evaluations to a manageable number, {\sc oopse} employs
218 a cut-off radius.\cite{allen87:csl} The cutoff radius is set to be
219 $2.5\sigma_{ii}$, where $\sigma_{ii}$ is the largest Lennard-Jones
220 length parameter present in the simulation. Truncating the calculation
221 at $r_{\text{cut}}$ introduces a discontinuity into the potential
222 energy. To offset this discontinuity, the energy value at
223 $r_{\text{cut}}$ is subtracted from the potential. This causes the
224 potential to go to zero smoothly at the cut-off radius.
225
226 Interactions between dissimilar particles requires the generation of
227 cross term parameters for $\sigma$ and $\epsilon$. These are
228 calculated through the Lorentz-Berthelot mixing
229 rules:\cite{allen87:csl}
230 \begin{equation}
231 \sigma_{ij} = \frac{1}{2}[\sigma_{ii} + \sigma_{jj}]
232 \label{eq:sigmaMix}
233 \end{equation}
234 and
235 \begin{equation}
236 \epsilon_{ij} = \sqrt{\epsilon_{ii} \epsilon_{jj}}
237 \label{eq:epsilonMix}
238 \end{equation}
239
240
241
242 \subsection{\label{sec:DUFF}Dipolar Unified-Atom Force Field}
243
244 The dipolar unified-atom force field ({\sc duff}) was developed to
245 simulate lipid bilayers. The simulations require a model capable of
246 forming bilayers, while still being sufficiently computationally
247 efficient to allow large systems ($\approx$100's of phospholipids,
248 $\approx$1000's of waters) to be simulated for long times
249 ($\approx$10's of nanoseconds).
250
251 With this goal in mind, {\sc duff} has no point
252 charges. Charge-neutral distributions were replaced with dipoles,
253 while most atoms and groups of atoms were reduced to Lennard-Jones
254 interaction sites. This simplification cuts the length scale of long
255 range interactions from $\frac{1}{r}$ to $\frac{1}{r^3}$, allowing us
256 to avoid the computationally expensive Ewald sum. Instead, we can use
257 neighbor-lists, reaction field, and cutoff radii for the dipolar
258 interactions.
259
260 As an example, lipid head-groups in {\sc duff} are represented as
261 point dipole interaction sites. By placing a dipole of 20.6~Debye at
262 the head group center of mass, our model mimics the head group of
263 phosphatidylcholine.\cite{Cevc87} Additionally, a large Lennard-Jones
264 site is located at the pseudoatom's center of mass. The model is
265 illustrated by the dark grey atom in Fig.~\ref{fig:lipidModel}. The
266 water model we use to complement the dipoles of the lipids is our
267 reparameterization of the soft sticky dipole (SSD) model of Ichiye
268 \emph{et al.}\cite{liu96:new_model}
269
270 \begin{figure}
271 \epsfxsize=\linewidth
272 \epsfbox{lipidModel.eps}
273 \caption{A representation of the lipid model. $\phi$ is the torsion angle, $\theta$ %
274 is the bend angle, $\mu$ is the dipole moment of the head group, and n
275 is the chain length.}
276 \label{fig:lipidModel}
277 \end{figure}
278
279 We have used a set of scalable parameters to model the alkyl groups
280 with Lennard-Jones sites. For this, we have borrowed parameters from
281 the TraPPE force field of Siepmann
282 \emph{et al}.\cite{Siepmann1998} TraPPE is a unified-atom
283 representation of n-alkanes, which is parametrized against phase
284 equilibria using Gibbs ensemble Monte Carlo simulation
285 techniques.\cite{Siepmann1998} One of the advantages of TraPPE is that
286 it generalizes the types of atoms in an alkyl chain to keep the number
287 of pseudoatoms to a minimum; the parameters for an atom such as
288 $\text{CH}_2$ do not change depending on what species are bonded to
289 it.
290
291 TraPPE also constrains all bonds to be of fixed length. Typically,
292 bond vibrations are the fastest motions in a molecular dynamic
293 simulation. Small time steps between force evaluations must be used to
294 ensure adequate sampling of the bond potential to ensure conservation
295 of energy. By constraining the bond lengths, larger time steps may be
296 used when integrating the equations of motion. A simulation using {\sc
297 duff} is illustrated in Scheme \ref{sch:DUFF}.
298
299 \begin{lstlisting}[caption={[Invocation of {\sc duff}]Sample \texttt{.bass} file showing a simulation utilizing {\sc duff}},label={sch:DUFF}]
300
301 #include "water.mdl"
302 #include "lipid.mdl"
303
304 nComponents = 2;
305 component{
306 type = "simpleLipid_16";
307 nMol = 60;
308 }
309
310 component{
311 type = "SSD_water";
312 nMol = 1936;
313 }
314
315 initialConfig = "bilayer.init";
316
317 forceField = "DUFF";
318
319 \end{lstlisting}
320
321 \subsubsection{\label{subSec:energyFunctions}{\sc duff} Energy Functions}
322
323 The total potential energy function in {\sc duff} is
324 \begin{equation}
325 V = \sum^{N}_{I=1} V^{I}_{\text{Internal}}
326 + \sum^{N}_{I=1} \sum_{J>I} V^{IJ}_{\text{Cross}}
327 \label{eq:totalPotential}
328 \end{equation}
329 Where $V^{I}_{\text{Internal}}$ is the internal potential of molecule $I$:
330 \begin{equation}
331 V^{I}_{\text{Internal}} =
332 \sum_{\theta_{ijk} \in I} V_{\text{bend}}(\theta_{ijk})
333 + \sum_{\phi_{ijkl} \in I} V_{\text{torsion}}(\phi_{ijkl})
334 + \sum_{i \in I} \sum_{(j>i+4) \in I}
335 \biggl[ V_{\text{LJ}}(r_{ij}) + V_{\text{dipole}}
336 (\mathbf{r}_{ij},\boldsymbol{\Omega}_{i},\boldsymbol{\Omega}_{j})
337 \biggr]
338 \label{eq:internalPotential}
339 \end{equation}
340 Here $V_{\text{bend}}$ is the bend potential for all 1, 3 bonded pairs
341 within the molecule $I$, and $V_{\text{torsion}}$ is the torsion potential
342 for all 1, 4 bonded pairs. The pairwise portions of the internal
343 potential are excluded for pairs that are closer than three bonds,
344 i.e.~atom pairs farther away than a torsion are included in the
345 pair-wise loop.
346
347
348 The bend potential of a molecule is represented by the following function:
349 \begin{equation}
350 V_{\text{bend}}(\theta_{ijk}) = k_{\theta}( \theta_{ijk} - \theta_0 )^2 \label{eq:bendPot}
351 \end{equation}
352 Where $\theta_{ijk}$ is the angle defined by atoms $i$, $j$, and $k$
353 (see Fig.~\ref{fig:lipidModel}), $\theta_0$ is the equilibrium
354 bond angle, and $k_{\theta}$ is the force constant which determines the
355 strength of the harmonic bend. The parameters for $k_{\theta}$ and
356 $\theta_0$ are borrowed from those in TraPPE.\cite{Siepmann1998}
357
358 The torsion potential and parameters are also borrowed from TraPPE. It is
359 of the form:
360 \begin{equation}
361 V_{\text{torsion}}(\phi) = c_1[1 + \cos \phi]
362 + c_2[1 + \cos(2\phi)]
363 + c_3[1 + \cos(3\phi)]
364 \label{eq:origTorsionPot}
365 \end{equation}
366 Here $\phi$ is the angle defined by four bonded neighbors $i$,
367 $j$, $k$, and $l$ (again, see Fig.~\ref{fig:lipidModel}). For
368 computational efficiency, the torsion potential has been recast after
369 the method of CHARMM,\cite{charmm1983} in which the angle series is
370 converted to a power series of the form:
371 \begin{equation}
372 V_{\text{torsion}}(\phi) =
373 k_3 \cos^3 \phi + k_2 \cos^2 \phi + k_1 \cos \phi + k_0
374 \label{eq:torsionPot}
375 \end{equation}
376 Where:
377 \begin{align*}
378 k_0 &= c_1 + c_3 \\
379 k_1 &= c_1 - 3c_3 \\
380 k_2 &= 2 c_2 \\
381 k_3 &= 4c_3
382 \end{align*}
383 By recasting the potential as a power series, repeated trigonometric
384 evaluations are avoided during the calculation of the potential energy.
385
386
387 The cross potential between molecules $I$ and $J$, $V^{IJ}_{\text{Cross}}$, is
388 as follows:
389 \begin{equation}
390 V^{IJ}_{\text{Cross}} =
391 \sum_{i \in I} \sum_{j \in J}
392 \biggl[ V_{\text{LJ}}(r_{ij}) + V_{\text{dipole}}
393 (\mathbf{r}_{ij},\boldsymbol{\Omega}_{i},\boldsymbol{\Omega}_{j})
394 + V_{\text{sticky}}
395 (\mathbf{r}_{ij},\boldsymbol{\Omega}_{i},\boldsymbol{\Omega}_{j})
396 \biggr]
397 \label{eq:crossPotentail}
398 \end{equation}
399 Where $V_{\text{LJ}}$ is the Lennard Jones potential,
400 $V_{\text{dipole}}$ is the dipole dipole potential, and
401 $V_{\text{sticky}}$ is the sticky potential defined by the SSD model
402 (Sec.~\ref{sec:SSD}). Note that not all atom types include all
403 interactions.
404
405 The dipole-dipole potential has the following form:
406 \begin{equation}
407 V_{\text{dipole}}(\mathbf{r}_{ij},\boldsymbol{\Omega}_{i},
408 \boldsymbol{\Omega}_{j}) = \frac{|\mu_i||\mu_j|}{4\pi\epsilon_{0}r_{ij}^{3}} \biggl[
409 \boldsymbol{\hat{u}}_{i} \cdot \boldsymbol{\hat{u}}_{j}
410 -
411 \frac{3(\boldsymbol{\hat{u}}_i \cdot \mathbf{r}_{ij}) %
412 (\boldsymbol{\hat{u}}_j \cdot \mathbf{r}_{ij}) }
413 {r^{2}_{ij}} \biggr]
414 \label{eq:dipolePot}
415 \end{equation}
416 Here $\mathbf{r}_{ij}$ is the vector starting at atom $i$ pointing
417 towards $j$, and $\boldsymbol{\Omega}_i$ and $\boldsymbol{\Omega}_j$
418 are the orientational degrees of freedom for atoms $i$ and $j$
419 respectively. $|\mu_i|$ is the magnitude of the dipole moment of atom
420 $i$, $\boldsymbol{\hat{u}}_i$ is the standard unit orientation
421 vector of $\boldsymbol{\Omega}_i$, and $\boldsymbol{\hat{r}}_{ij}$ is
422 the unit vector pointing along $\mathbf{r}_{ij}$.
423
424
425 \subsubsection{\label{sec:SSD}The {\sc duff} Water Models: SSD/E and SSD/RF}
426
427 In the interest of computational efficiency, the default solvent used
428 by {\sc oopse} is the extended Soft Sticky Dipole (SSD/E) water
429 model.\cite{Gezelter04} The original SSD was developed by Ichiye
430 \emph{et al.}\cite{liu96:new_model} as a modified form of the hard-sphere
431 water model proposed by Bratko, Blum, and
432 Luzar.\cite{Bratko85,Bratko95} It consists of a single point dipole
433 with a Lennard-Jones core and a sticky potential that directs the
434 particles to assume the proper hydrogen bond orientation in the first
435 solvation shell. Thus, the interaction between two SSD water molecules
436 \emph{i} and \emph{j} is given by the potential
437 \begin{equation}
438 V_{ij} =
439 V_{ij}^{LJ} (r_{ij})\ + V_{ij}^{dp}
440 (\mathbf{r}_{ij},\boldsymbol{\Omega}_i,\boldsymbol{\Omega}_j)\ +
441 V_{ij}^{sp}
442 (\mathbf{r}_{ij},\boldsymbol{\Omega}_i,\boldsymbol{\Omega}_j),
443 \label{eq:ssdPot}
444 \end{equation}
445 where the $\mathbf{r}_{ij}$ is the position vector between molecules
446 \emph{i} and \emph{j} with magnitude equal to the distance $r_{ij}$, and
447 $\boldsymbol{\Omega}_i$ and $\boldsymbol{\Omega}_j$ represent the
448 orientations of the respective molecules. The Lennard-Jones and dipole
449 parts of the potential are given by equations \ref{eq:lennardJonesPot}
450 and \ref{eq:dipolePot} respectively. The sticky part is described by
451 the following,
452 \begin{equation}
453 u_{ij}^{sp}(\mathbf{r}_{ij},\boldsymbol{\Omega}_i,\boldsymbol{\Omega}_j)=
454 \frac{\nu_0}{2}[s(r_{ij})w(\mathbf{r}_{ij},
455 \boldsymbol{\Omega}_i,\boldsymbol{\Omega}_j) +
456 s^\prime(r_{ij})w^\prime(\mathbf{r}_{ij},
457 \boldsymbol{\Omega}_i,\boldsymbol{\Omega}_j)]\ ,
458 \label{eq:stickyPot}
459 \end{equation}
460 where $\nu_0$ is a strength parameter for the sticky potential, and
461 $s$ and $s^\prime$ are cubic switching functions which turn off the
462 sticky interaction beyond the first solvation shell. The $w$ function
463 can be thought of as an attractive potential with tetrahedral
464 geometry:
465 \begin{equation}
466 w({\bf r}_{ij},{\bf \Omega}_i,{\bf \Omega}_j)=
467 \sin\theta_{ij}\sin2\theta_{ij}\cos2\phi_{ij},
468 \label{eq:stickyW}
469 \end{equation}
470 while the $w^\prime$ function counters the normal aligned and
471 anti-aligned structures favored by point dipoles:
472 \begin{equation}
473 w^\prime({\bf r}_{ij},{\bf \Omega}_i,{\bf \Omega}_j)=
474 (\cos\theta_{ij}-0.6)^2(\cos\theta_{ij}+0.8)^2-w^0,
475 \label{eq:stickyWprime}
476 \end{equation}
477 It should be noted that $w$ is proportional to the sum of the $Y_3^2$
478 and $Y_3^{-2}$ spherical harmonics (a linear combination which
479 enhances the tetrahedral geometry for hydrogen bonded structures),
480 while $w^\prime$ is a purely empirical function. A more detailed
481 description of the functional parts and variables in this potential
482 can be found in the original SSD
483 articles.\cite{liu96:new_model,liu96:monte_carlo,chandra99:ssd_md,Ichiye03}
484
485 Since SSD is a single-point {\it dipolar} model, the force
486 calculations are simplified significantly relative to the standard
487 {\it charged} multi-point models. In the original Monte Carlo
488 simulations using this model, Ichiye {\it et al.} reported that using
489 SSD decreased computer time by a factor of 6-7 compared to other
490 models.\cite{liu96:new_model} What is most impressive is that these savings
491 did not come at the expense of accurate depiction of the liquid state
492 properties. Indeed, SSD maintains reasonable agreement with the Soper
493 diffraction data for the structural features of liquid
494 water.\cite{Soper86,liu96:new_model} Additionally, the dynamical properties
495 exhibited by SSD agree with experiment better than those of more
496 computationally expensive models (like TIP3P and
497 SPC/E).\cite{chandra99:ssd_md} The combination of speed and accurate depiction
498 of solvent properties makes SSD a very attractive model for the
499 simulation of large scale biochemical simulations.
500
501 Recent constant pressure simulations revealed issues in the original
502 SSD model that led to lower than expected densities at all target
503 pressures.\cite{Ichiye03,Gezelter04} The default model in {\sc oopse}
504 is therefore SSD/E, a density corrected derivative of SSD that
505 exhibits improved liquid structure and transport behavior. If the use
506 of a reaction field long-range interaction correction is desired, it
507 is recommended that the parameters be modified to those of the SSD/RF
508 model. Solvent parameters can be easily modified in an accompanying
509 {\sc BASS} file as illustrated in the scheme below. A table of the
510 parameter values and the drawbacks and benefits of the different
511 density corrected SSD models can be found in reference
512 \ref{Gezelter04}.
513
514 \begin{lstlisting}[caption={[A simulation of {\sc ssd} water]An example file showing a simulation including {\sc ssd} water.},label={sch:ssd}]
515
516 #include "water.mdl"
517
518 nComponents = 1;
519 component{
520 type = "SSD_water";
521 nMol = 864;
522 }
523
524 initialConfig = "liquidWater.init";
525
526 forceField = "DUFF";
527
528 /*
529 * The reactionField flag toggles reaction
530 * field corrections.
531 */
532
533 reactionField = false; // defaults to false
534 dielectric = 80.0; // dielectric for reaction field
535
536 /*
537 * The following two flags set the cutoff
538 * radius for the electrostatic forces
539 * as well as the skin thickness of the switching
540 * function.
541 */
542
543 electrostaticCutoffRadius = 9.2;
544 electrostaticSkinThickness = 1.38;
545
546 \end{lstlisting}
547
548
549 \subsection{\label{sec:eam}Embedded Atom Method}
550
551 Several other molecular dynamics packages\cite{dynamo86} exist which have the
552 capacity to simulate metallic systems, including some that have
553 parallel computational abilities\cite{plimpton93}. Potentials that
554 describe bonding transition metal
555 systems\cite{Finnis84,Ercolessi88,Chen90,Qi99,Ercolessi02} have a
556 attractive interaction which models ``Embedding''
557 a positively charged metal ion in the electron density due to the
558 free valance ``sea'' of electrons created by the surrounding atoms in
559 the system. A mostly repulsive pairwise part of the potential
560 describes the interaction of the positively charged metal core ions
561 with one another. A particular potential description called the
562 Embedded Atom Method\cite{Daw84,FBD86,johnson89,Lu97}({\sc eam}) that has
563 particularly wide adoption has been selected for inclusion in {\sc oopse}. A
564 good review of {\sc eam} and other metallic potential formulations was done
565 by Voter.\cite{voter}
566
567 The {\sc eam} potential has the form:
568 \begin{eqnarray}
569 V & = & \sum_{i} F_{i}\left[\rho_{i}\right] + \sum_{i} \sum_{j \neq i}
570 \phi_{ij}({\bf r}_{ij}) \\
571 \rho_{i} & = & \sum_{j \neq i} f_{j}({\bf r}_{ij})
572 \end{eqnarray}S
573
574 where $F_{i} $ is the embedding function that equates the energy required to embed a
575 positively-charged core ion $i$ into a linear superposition of
576 spherically averaged atomic electron densities given by
577 $\rho_{i}$. $\phi_{ij}$ is a primarily repulsive pairwise interaction
578 between atoms $i$ and $j$. In the original formulation of
579 {\sc eam} cite{Daw84}, $\phi_{ij}$ was an entirely repulsive term, however
580 in later refinements to EAM have shown that non-uniqueness between $F$
581 and $\phi$ allow for more general forms for $\phi$.\cite{Daw89}
582 There is a cutoff distance, $r_{cut}$, which limits the
583 summations in the {\sc eam} equation to the few dozen atoms
584 surrounding atom $i$ for both the density $\rho$ and pairwise $\phi$
585 interactions. Foiles et al. fit EAM potentials for fcc metals Cu, Ag, Au, Ni, Pd, Pt and alloys of these metals\cite{FDB86}. These potential fits are in the DYNAMO 86 format and are included with {\sc oopse}.
586
587
588 \subsection{\label{Sec:pbc}Periodic Boundary Conditions}
589
590 \newcommand{\roundme}{\operatorname{round}}
591
592 \textit{Periodic boundary conditions} are widely used to simulate truly
593 macroscopic systems with a relatively small number of particles. The
594 simulation box is replicated throughout space to form an infinite lattice.
595 During the simulation, when a particle moves in the primary cell, its image in
596 other boxes move in exactly the same direction with exactly the same
597 orientation.Thus, as a particle leaves the primary cell, one of its images
598 will enter through the opposite face.If the simulation box is large enough to
599 avoid \textquotedblleft feeling\textquotedblright\ the symmetries of the
600 periodic lattice, surface effects can be ignored. Cubic, orthorhombic and
601 parallelepiped are the available periodic cells In OOPSE. We use a matrix to
602 describe the property of the simulation box. Therefore, both the size and
603 shape of the simulation box can be changed during the simulation. The
604 transformation from box space vector $\mathbf{s}$ to its corresponding real
605 space vector $\mathbf{r}$ is defined by
606 \begin{equation}
607 \mathbf{r}=\underline{\mathbf{H}}\cdot\mathbf{s}%
608 \end{equation}
609
610
611 where $H=(h_{x},h_{y},h_{z})$ is a transformation matrix made up of the three
612 box axis vectors. $h_{x},h_{y}$ and $h_{z}$ represent the three sides of the
613 simulation box respectively.
614
615 To find the minimum image of a vector $\mathbf{r}$, we convert the real vector
616 to its corresponding vector in box space first, \bigskip%
617 \begin{equation}
618 \mathbf{s}=\underline{\mathbf{H}}^{-1}\cdot\mathbf{r}%
619 \end{equation}
620 And then, each element of $\mathbf{s}$ is wrapped to lie between -0.5 to 0.5,
621 \begin{equation}
622 s_{i}^{\prime}=s_{i}-\roundme(s_{i})
623 \end{equation}
624 where
625
626 %
627
628 \begin{equation}
629 \roundme(x)=\left\{
630 \begin{array}{cc}%
631 \lfloor{x+0.5}\rfloor & \text{if \ }x\geqslant0\\
632 \lceil{x-0.5}\rceil & \text{otherwise}%
633 \end{array}
634 \right.
635 \end{equation}
636
637
638 For example, $\roundme(3.6)=4$,$\roundme(3.1)=3$, $\roundme(-3.6)=-4$, $\roundme(-3.1)=-3$.
639
640 Finally, we obtain the minimum image coordinates $\mathbf{r}^{\prime}$ by
641 transforming back to real space,%
642
643 \begin{equation}
644 \mathbf{r}^{\prime}=\underline{\mathbf{H}}^{-1}\cdot\mathbf{s}^{\prime}%
645 \end{equation}
646
647
648 \section{Input and Output Files}
649
650 \subsection{{\sc bass} and Model Files}
651
652 Every {\sc oopse} simuation begins with a {\sc bass} file. {\sc bass}
653 (\underline{B}izarre \underline{A}tom \underline{S}imulation
654 \underline{S}yntax) is a script syntax that is parsed by {\sc oopse} at
655 runtime. The {\sc bass} file allows for the user to completely describe the
656 system they are to simulate, as well as tailor {\sc oopse}'s behavior during
657 the simulation. {\sc bass} files are denoted with the extension
658 \texttt{.bass}, an example file is shown in
659 Fig.~\ref{fig:bassExample}.
660
661 \begin{figure}
662
663 \centering
664 \framebox[\linewidth]{\rule{0cm}{0.75\linewidth}I'm a {\sc bass} file!}
665 \caption{Here is an example \texttt{.bass} file}
666 \label{fig:bassExample}
667 \end{figure}
668
669 Within the \texttt{.bass} file it is neccassary to provide a complete
670 description of the molecule before it is actually placed in the
671 simulation. The {\sc bass} syntax was originally developed with this goal in
672 mind, and allows for the specification of all the atoms in a molecular
673 prototype, as well as any bonds, bends, or torsions. These
674 descriptions can become lengthy for complex molecules, and it would be
675 inconvient to duplicate the simulation at the begining of each {\sc bass}
676 script. Addressing this issue {\sc bass} allows for the inclusion of model
677 files at the top of a \texttt{.bass} file. These model files, denoted
678 with the \texttt{.mdl} extension, allow the user to describe a
679 molecular prototype once, then simply include it into each simulation
680 containing that molecule.
681
682 \subsection{\label{subSec:coordFiles}Coordinate Files}
683
684 The standard format for storage of a systems coordinates is a modified
685 xyz-file syntax, the exact details of which can be seen in
686 App.~\ref{appCoordFormat}. As all bonding and molecular information is
687 stored in the \texttt{.bass} and \texttt{.mdl} files, the coordinate
688 files are simply the complete set of coordinates for each atom at a
689 given simulation time.
690
691 There are three major files used by {\sc oopse} written in the coordinate
692 format, they are as follows: the initialization file, the simulation
693 trajectory file, and the final coordinates of the simulation. The
694 initialization file is neccassary for {\sc oopse} to start the simulation
695 with the proper coordinates. It is typically denoted with the
696 extension \texttt{.init}. The trajectory file is created at the
697 beginning of the simulation, and is used to store snapshots of the
698 simulation at regular intervals. The first frame is a duplication of
699 the \texttt{.init} file, and each subsequent frame is appended to the
700 file at an interval specified in the \texttt{.bass} file. The
701 trajectory file is given the extension \texttt{.dump}. The final
702 coordinate file is the end of run or \texttt{.eor} file. The
703 \texttt{.eor} file stores the final configuration of teh system for a
704 given simulation. The file is updated at the same time as the
705 \texttt{.dump} file. However, it only contains the most recent
706 frame. In this way, an \texttt{.eor} file may be used as the
707 initialization file to a second simulation in order to continue or
708 recover the previous simulation.
709
710 \subsection{Generation of Initial Coordinates}
711
712 As was stated in Sec.~\ref{subSec:coordFiles}, an initialization file
713 is needed to provide the starting coordinates for a simulation. The
714 {\sc oopse} package provides a program called \texttt{sysBuilder} to aid in
715 the creation of the \texttt{.init} file. \texttt{sysBuilder} is {\sc bass}
716 aware, and will recognize arguments and parameters in the
717 \texttt{.bass} file that would otherwise be ignored by the
718 simulation. The program itself is under contiunual development, and is
719 offered here as a helper tool only.
720
721 \subsection{The Statistics File}
722
723 The last output file generated by {\sc oopse} is the statistics file. This
724 file records such statistical quantities as the instantaneous
725 temperature, volume, pressure, etc. It is written out with the
726 frequency specified in the \texttt{.bass} file. The file allows the
727 user to observe the system variables as a function od simulation time
728 while the simulation is in progress. One useful function the
729 statistics file serves is to monitor the conserved quantity of a given
730 simulation ensemble, this allows the user to observe the stability of
731 the integrator. The statistics file is denoted with the \texttt{.stat}
732 file extension.
733
734 \section{\label{sec:mechanics}Mechanics}
735
736 \subsection{\label{integrate}Integrating the Equations of Motion: the Symplectic Step Integrator}
737
738 Integration of the equations of motion was carried out using the
739 symplectic splitting method proposed by Dullweber \emph{et
740 al.}.\cite{Dullweber1997} The reason for this integrator selection
741 deals with poor energy conservation of rigid body systems using
742 quaternions. While quaternions work well for orientational motion in
743 alternate ensembles, the microcanonical ensemble has a constant energy
744 requirement that is quite sensitive to errors in the equations of
745 motion. The original implementation of this code utilized quaternions
746 for rotational motion propagation; however, a detailed investigation
747 showed that they resulted in a steady drift in the total energy,
748 something that has been observed by others.\cite{Laird97}
749
750 The key difference in the integration method proposed by Dullweber
751 \emph{et al.} is that the entire rotation matrix is propagated from
752 one time step to the next. In the past, this would not have been as
753 feasible a option, being that the rotation matrix for a single body is
754 nine elements long as opposed to 3 or 4 elements for Euler angles and
755 quaternions respectively. System memory has become much less of an
756 issue in recent times, and this has resulted in substantial benefits
757 in energy conservation. There is still the issue of 5 or 6 additional
758 elements for describing the orientation of each particle, which will
759 increase dump files substantially. Simply translating the rotation
760 matrix into its component Euler angles or quaternions for storage
761 purposes relieves this burden.
762
763 The symplectic splitting method allows for Verlet style integration of
764 both linear and angular motion of rigid bodies. In the integration
765 method, the orientational propagation involves a sequence of matrix
766 evaluations to update the rotation matrix.\cite{Dullweber1997} These
767 matrix rotations end up being more costly computationally than the
768 simpler arithmetic quaternion propagation. With the same time step, a
769 1000 SSD particle simulation shows an average 7\% increase in
770 computation time using the symplectic step method in place of
771 quaternions. This cost is more than justified when comparing the
772 energy conservation of the two methods as illustrated in figure
773 \ref{timestep}.
774
775 \begin{figure}
776 \epsfxsize=6in
777 \epsfbox{timeStep.epsi}
778 \caption{Energy conservation using quaternion based integration versus
779 the symplectic step method proposed by Dullweber \emph{et al.} with
780 increasing time step. For each time step, the dotted line is total
781 energy using the symplectic step integrator, and the solid line comes
782 from the quaternion integrator. The larger time step plots are shifted
783 up from the true energy baseline for clarity.}
784 \label{timestep}
785 \end{figure}
786
787 In figure \ref{timestep}, the resulting energy drift at various time
788 steps for both the symplectic step and quaternion integration schemes
789 is compared. All of the 1000 SSD particle simulations started with the
790 same configuration, and the only difference was the method for
791 handling rotational motion. At time steps of 0.1 and 0.5 fs, both
792 methods for propagating particle rotation conserve energy fairly well,
793 with the quaternion method showing a slight energy drift over time in
794 the 0.5 fs time step simulation. At time steps of 1 and 2 fs, the
795 energy conservation benefits of the symplectic step method are clearly
796 demonstrated. Thus, while maintaining the same degree of energy
797 conservation, one can take considerably longer time steps, leading to
798 an overall reduction in computation time.
799
800 Energy drift in these SSD particle simulations was unnoticeable for
801 time steps up to three femtoseconds. A slight energy drift on the
802 order of 0.012 kcal/mol per nanosecond was observed at a time step of
803 four femtoseconds, and as expected, this drift increases dramatically
804 with increasing time step. To insure accuracy in the constant energy
805 simulations, time steps were set at 2 fs and kept at this value for
806 constant pressure simulations as well.
807
808
809 \subsection{\label{sec:extended}Extended Systems for other Ensembles}
810
811
812 {\sc oopse} implements a
813
814
815 \subsubsection{\label{sec:noseHooverThermo}Nose-Hoover Thermostatting}
816
817 To mimic the effects of being in a constant temperature ({\sc nvt})
818 ensemble, {\sc oopse} uses the Nose-Hoover extended system
819 approach.\cite{Hoover85} In this method, the equations of motion for
820 the particle positions and velocities are
821 \begin{eqnarray}
822 \dot{{\bf r}} & = & {\bf v} \\
823 \dot{{\bf v}} & = & \frac{{\bf f}}{m} - \chi {\bf v}
824 \label{eq:nosehoovereom}
825 \end{eqnarray}
826
827 $\chi$ is an ``extra'' variable included in the extended system, and
828 it is propagated using the first order equation of motion
829 \begin{equation}
830 \dot{\chi} = \frac{1}{\tau_{T}} \left( \frac{T}{T_{target}} - 1 \right)
831 \label{eq:nosehooverext}
832 \end{equation}
833 where $T_{target}$ is the target temperature for the simulation, and
834 $\tau_{T}$ is a time constant for the thermostat.
835
836 To select the Nose-Hoover {\sc nvt} ensemble, the {\tt ensemble = NVT;}
837 command would be used in the simulation's {\sc bass} file. There is
838 some subtlety in choosing values for $\tau_{T}$, and it is usually set
839 to values of a few ps. Within a {\sc bass} file, $\tau_{T}$ could be
840 set to 1 ps using the {\tt tauThermostat = 1000; } command.
841
842
843 \subsection{\label{Sec:zcons}Z-Constraint Method}
844
845 Based on fluctuatin-dissipation theorem,\bigskip\ force auto-correlation
846 method was developed to investigate the dynamics of ions inside the ion
847 channels.\cite{Roux91} Time-dependent friction coefficient can be calculated
848 from the deviation of the instaneous force from its mean force.
849
850 %
851
852 \begin{equation}
853 \xi(z,t)=\langle\delta F(z,t)\delta F(z,0)\rangle/k_{B}T
854 \end{equation}
855
856
857 where%
858 \begin{equation}
859 \delta F(z,t)=F(z,t)-\langle F(z,t)\rangle
860 \end{equation}
861
862
863 If the time-dependent friction decay rapidly, static friction coefficient can
864 be approximated by%
865
866 \begin{equation}
867 \xi^{static}(z)=\int_{0}^{\infty}\langle\delta F(z,t)\delta F(z,0)\rangle dt
868 \end{equation}
869
870
871 Hence, diffusion constant can be estimated by
872 \begin{equation}
873 D(z)=\frac{k_{B}T}{\xi^{static}(z)}=\frac{(k_{B}T)^{2}}{\int_{0}^{\infty
874 }\langle\delta F(z,t)\delta F(z,0)\rangle dt}%
875 \end{equation}
876
877
878 \bigskip Z-Constraint method, which fixed the z coordinates of the molecules
879 with respect to the center of the mass of the system, was proposed to obtain
880 the forces required in force auto-correlation method.\cite{Marrink94} However,
881 simply resetting the coordinate will move the center of the mass of the whole
882 system. To avoid this problem, a new method was used at {\sc oopse}. Instead of
883 resetting the coordinate, we reset the forces of z-constraint molecules as
884 well as subtract the total constraint forces from the rest of the system after
885 force calculation at each time step.
886 \begin{verbatim}
887 $F_{\alpha i}=0$
888 $V_{\alpha i}=V_{\alpha i}-\frac{\sum\limits_{i}M_{_{\alpha i}}V_{\alpha i}}{\sum\limits_{i}M_{_{\alpha i}}}$
889 $F_{\alpha i}=F_{\alpha i}-\frac{M_{_{\alpha i}}}{\sum\limits_{\alpha}\sum\limits_{i}M_{_{\alpha i}}}\sum\limits_{\beta}F_{\beta}$
890 $V_{\alpha i}=V_{\alpha i}-\frac{\sum\limits_{\alpha}\sum\limits_{i}M_{_{\alpha i}}V_{\alpha i}}{\sum\limits_{\alpha}\sum\limits_{i}M_{_{\alpha i}}}$
891 \end{verbatim}
892
893 At the very beginning of the simulation, the molecules may not be at its
894 constraint position. To move the z-constraint molecule to the specified
895 position, a simple harmonic potential is used%
896
897 \begin{equation}
898 U(t)=\frac{1}{2}k_{Harmonic}(z(t)-z_{cons})^{2}%
899 \end{equation}
900 where $k_{Harmonic}$\bigskip\ is the harmonic force constant, $z(t)$ is
901 current z coordinate of the center of mass of the z-constraint molecule, and
902 $z_{cons}$ is the restraint position. Therefore, the harmonic force operated
903 on the z-constraint molecule at time $t$ can be calculated by%
904 \begin{equation}
905 F_{z_{Harmonic}}(t)=-\frac{\partial U(t)}{\partial z(t)}=-k_{Harmonic}%
906 (z(t)-z_{cons})
907 \end{equation}
908 Worthy of mention, other kinds of potential functions can also be used to
909 drive the z-constraint molecule.
910
911 \section{\label{sec:analysis}Trajectory Analysis}
912
913 \subsection{\label{subSec:staticProps}Static Property Analysis}
914
915 The static properties of the trajectories are analyzed with the
916 program \texttt{staticProps}. The code is capable of calculating the following
917 pair correlations between species A and B:
918 \begin{itemize}
919 \item $g_{\text{AB}}(r)$: Eq.~\ref{eq:gofr}
920 \item $g_{\text{AB}}(r, \cos \theta)$: Eq.~\ref{eq:gofrCosTheta}
921 \item $g_{\text{AB}}(r, \cos \omega)$: Eq.~\ref{eq:gofrCosOmega}
922 \item $g_{\text{AB}}(x, y, z)$: Eq.~\ref{eq:gofrXYZ}
923 \item $\langle \cos \omega \rangle_{\text{AB}}(r)$:
924 Eq.~\ref{eq:cosOmegaOfR}
925 \end{itemize}
926
927 The first pair correlation, $g_{\text{AB}}(r)$, is defined as follows:
928 \begin{equation}
929 g_{\text{AB}}(r) = \frac{V}{N_{\text{A}}N_{\text{B}}}\langle %%
930 \sum_{i \in \text{A}} \sum_{j \in \text{B}} %%
931 \delta( r - |\mathbf{r}_{ij}|) \rangle \label{eq:gofr}
932 \end{equation}
933 Where $\mathbf{r}_{ij}$ is the vector
934 \begin{equation*}
935 \mathbf{r}_{ij} = \mathbf{r}_j - \mathbf{r}_i \notag
936 \end{equation*}
937 and $\frac{V}{N_{\text{A}}N_{\text{B}}}$ normalizes the average over
938 the expected pair density at a given $r$.
939
940 The next two pair correlations, $g_{\text{AB}}(r, \cos \theta)$ and
941 $g_{\text{AB}}(r, \cos \omega)$, are similar in that they are both two
942 dimensional histograms. Both use $r$ for the primary axis then a
943 $\cos$ for the secondary axis ($\cos \theta$ for
944 Eq.~\ref{eq:gofrCosTheta} and $\cos \omega$ for
945 Eq.~\ref{eq:gofrCosOmega}). This allows for the investigator to
946 correlate alignment on directional entities. $g_{\text{AB}}(r, \cos
947 \theta)$ is defined as follows:
948 \begin{equation}
949 g_{\text{AB}}(r, \cos \theta) = \frac{V}{N_{\text{A}}N_{\text{B}}}\langle
950 \sum_{i \in \text{A}} \sum_{j \in \text{B}}
951 \delta( \cos \theta - \cos \theta_{ij})
952 \delta( r - |\mathbf{r}_{ij}|) \rangle
953 \label{eq:gofrCosTheta}
954 \end{equation}
955 Where
956 \begin{equation*}
957 \cos \theta_{ij} = \mathbf{\hat{i}} \cdot \mathbf{\hat{r}}_{ij}
958 \end{equation*}
959 Here $\mathbf{\hat{i}}$ is the unit directional vector of species $i$
960 and $\mathbf{\hat{r}}_{ij}$ is the unit vector associated with vector
961 $\mathbf{r}_{ij}$.
962
963 The second two dimensional histogram is of the form:
964 \begin{equation}
965 g_{\text{AB}}(r, \cos \omega) =
966 \frac{V}{N_{\text{A}}N_{\text{B}}}\langle
967 \sum_{i \in \text{A}} \sum_{j \in \text{B}}
968 \delta( \cos \omega - \cos \omega_{ij})
969 \delta( r - |\mathbf{r}_{ij}|) \rangle \label{eq:gofrCosOmega}
970 \end{equation}
971 Here
972 \begin{equation*}
973 \cos \omega_{ij} = \mathbf{\hat{i}} \cdot \mathbf{\hat{j}}
974 \end{equation*}
975 Again, $\mathbf{\hat{i}}$ and $\mathbf{\hat{j}}$ are the unit
976 directional vectors of species $i$ and $j$.
977
978 The static analysis code is also cable of calculating a three
979 dimensional pair correlation of the form:
980 \begin{equation}\label{eq:gofrXYZ}
981 g_{\text{AB}}(x, y, z) =
982 \frac{V}{N_{\text{A}}N_{\text{B}}}\langle
983 \sum_{i \in \text{A}} \sum_{j \in \text{B}}
984 \delta( x - x_{ij})
985 \delta( y - y_{ij})
986 \delta( z - z_{ij}) \rangle
987 \end{equation}
988 Where $x_{ij}$, $y_{ij}$, and $z_{ij}$ are the $x$, $y$, and $z$
989 components respectively of vector $\mathbf{r}_{ij}$.
990
991 The final pair correlation is similar to
992 Eq.~\ref{eq:gofrCosOmega}. $\langle \cos \omega
993 \rangle_{\text{AB}}(r)$ is calculated in the following way:
994 \begin{equation}\label{eq:cosOmegaOfR}
995 \langle \cos \omega \rangle_{\text{AB}}(r) =
996 \langle \sum_{i \in \text{A}} \sum_{j \in \text{B}}
997 (\cos \omega_{ij}) \delta( r - |\mathbf{r}_{ij}|) \rangle
998 \end{equation}
999 Here $\cos \omega_{ij}$ is defined in the same way as in
1000 Eq.~\ref{eq:gofrCosOmega}. This equation is a single dimensional pair
1001 correlation that gives the average correlation of two directional
1002 entities as a function of their distance from each other.
1003
1004 All static properties are calculated on a frame by frame basis. The
1005 trajectory is read a single frame at a time, and the appropriate
1006 calculations are done on each frame. Once one frame is finished, the
1007 next frame is read in, and a running average of the property being
1008 calculated is accumulated in each frame. The program allows for the
1009 user to specify more than one property be calculated in single run,
1010 preventing the need to read a file multiple times.
1011
1012 \subsection{\label{dynamicProps}Dynamic Property Analysis}
1013
1014 The dynamic properties of a trajectory are calculated with the program
1015 \texttt{dynamicProps}. The program will calculate the following properties:
1016 \begin{gather}
1017 \langle | \mathbf{r}(t) - \mathbf{r}(0) |^2 \rangle \label{eq:rms}\\
1018 \langle \mathbf{v}(t) \cdot \mathbf{v}(0) \rangle \label{eq:velCorr} \\
1019 \langle \mathbf{j}(t) \cdot \mathbf{j}(0) \rangle \label{eq:angularVelCorr}
1020 \end{gather}
1021
1022 Eq.~\ref{eq:rms} is the root mean square displacement
1023 function. Eq.~\ref{eq:velCorr} and Eq.~\ref{eq:angularVelCorr} are the
1024 velocity and angular velocity correlation functions respectively. The
1025 latter is only applicable to directional species in the simulation.
1026
1027 The \texttt{dynamicProps} program handles he file in a manner different from
1028 \texttt{staticProps}. As the properties calculated by this program are time
1029 dependent, multiple frames must be read in simultaneously by the
1030 program. For small trajectories this is no problem, and the entire
1031 trajectory is read into memory. However, for long trajectories of
1032 large systems, the files can be quite large. In order to accommodate
1033 large files, \texttt{dynamicProps} adopts a scheme whereby two blocks of memory
1034 are allocated to read in several frames each.
1035
1036 In this two block scheme, the correlation functions are first
1037 calculated within each memory block, then the cross correlations
1038 between the frames contained within the two blocks are
1039 calculated. Once completed, the memory blocks are incremented, and the
1040 process is repeated. A diagram illustrating the process is shown in
1041 Fig.~\ref{fig:dynamicPropsMemory}. As was the case with \texttt{staticProps},
1042 multiple properties may be calculated in a single run to avoid
1043 multiple reads on the same file.
1044
1045 \begin{figure}
1046 \epsfxsize=6in
1047 \epsfbox{dynamicPropsMem.eps}
1048 \caption{This diagram illustrates the dynamic memory allocation 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.}
1049 \label{fig:dynamicPropsMemory}
1050 \end{figure}
1051
1052 \section{\label{sec:ProgramDesign}Program Design}
1053
1054 \subsection{\label{sec:architecture} OOPSE Architecture}
1055
1056 The core of OOPSE is divided into two main object libraries: {\texttt
1057 libBASS} and {\texttt libmdtools}. {\texttt libBASS} is the library
1058 developed around the parseing engine and {\texttt libmdtools} is the
1059 software library developed around the simulation engine.
1060
1061
1062
1063 \subsection{\label{sec:programLang} Programming Languages }
1064
1065 \subsection{\label{sec:parallelization} Parallelization of OOPSE}
1066
1067 Although processor power is doubling roughly every 18 months according
1068 to the famous Moore's Law\cite{moore}, it is still unreasonable to
1069 simulate systems of more then a 1000 atoms on a single processor. To
1070 facilitate study of larger system sizes or smaller systems on long
1071 time scales in a reasonable period of time, parallel methods were
1072 developed allowing multiple CPU's to share the simulation
1073 workload. Three general categories of parallel decomposition method's
1074 have been developed including atomic, spatial and force decomposition
1075 methods.
1076
1077 Algorithmically simplest of the three method's is atomic decomposition
1078 where N particles in a simulation are split among P processors for the
1079 duration of the simulation. Computational cost scales as an optimal
1080 $O(N/P)$ for atomic decomposition. Unfortunately all processors must
1081 communicate positions and forces with all other processors leading
1082 communication to scale as an unfavorable $O(N)$ independent of the
1083 number of processors. This communication bottleneck led to the
1084 development of spatial and force decomposition methods in which
1085 communication among processors scales much more favorably. Spatial or
1086 domain decomposition divides the physical spatial domain into 3D boxes
1087 in which each processor is responsible for calculation of forces and
1088 positions of particles located in its box. Particles are reassigned to
1089 different processors as they move through simulation space. To
1090 calculate forces on a given particle, a processor must know the
1091 positions of particles within some cutoff radius located on nearby
1092 processors instead of the positions of particles on all
1093 processors. Both communication between processors and computation
1094 scale as $O(N/P)$ in the spatial method. However, spatial
1095 decomposition adds algorithmic complexity to the simulation code and
1096 is not very efficient for small N since the overall communication
1097 scales as the surface to volume ratio $(N/P)^{2/3}$ in three
1098 dimensions.
1099
1100 Force decomposition assigns particles to processors based on a block
1101 decomposition of the force matrix. Processors are split into a
1102 optimally square grid forming row and column processor groups. Forces
1103 are calculated on particles in a given row by particles located in
1104 that processors column assignment. Force decomposition is less complex
1105 to implement then the spatial method but still scales computationally
1106 as $O(N/P)$ and scales as $(N/\sqrt{p})$ in communication
1107 cost. Plimpton also found that force decompositions scales more
1108 favorably then spatial decomposition up to 10,000 atoms and favorably
1109 competes with spatial methods for up to 100,000 atoms.
1110
1111 \subsection{\label{sec:memory}Memory Allocation in Analysis}
1112
1113 \subsection{\label{sec:documentation}Documentation}
1114
1115 \subsection{\label{openSource}Open Source and Distribution License}
1116
1117
1118 \section{\label{sec:conclusion}Conclusion}
1119
1120 \begin{itemize}
1121
1122 \item Restate capabilities
1123
1124 \item recap major structure / design choices
1125
1126 \begin{itemize}
1127
1128 \item parallel
1129 \item symplectic integration
1130 \item languages
1131
1132 \end{itemize}
1133
1134 \item How well does it meet the primary goal
1135
1136 \end{itemize}
1137 \section{Acknowledgments}
1138 The authors would like to thank espresso for fueling this work, and
1139 would also like to send a special acknowledgement to single malt
1140 scotch for its wonderful calming effects and its ability to make the
1141 troubles of the world float away.
1142 \bibliographystyle{achemso}
1143
1144 \bibliography{oopse}
1145
1146 \end{document}