ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/mattDisertation/oopse.tex
Revision: 1087
Committed: Fri Mar 5 22:16:34 2004 UTC (20 years, 3 months ago) by mmeineke
Content type: application/x-tex
File size: 100135 byte(s)
Log Message:
semi-final draft

File Contents

# User Rev Content
1 mmeineke 1045 \chapter{\label{chapt:oopse}OOPSE: AN OPEN SOURCE OBJECT-ORIENTED PARALLEL SIMULATION ENGINE FOR MOLECULAR DYNAMICS}
2 mmeineke 1044
3    
4    
5 mmeineke 1045 %% \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 mmeineke 1044
14 mmeineke 1045 \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 mmeineke 1044
18 mmeineke 1051 \section{\label{oopseSec:foreword}Foreword}
19 mmeineke 1044
20 mmeineke 1051 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 mmeineke 1068 Gezelter. Although my contributions to {\sc oopse} are major,
26     consideration of my work apart from the others would not give a
27 mmeineke 1051 complete description to the package's capabilities. As such, all
28     contributions to {\sc oopse} to date are presented in this chapter.
29 mmeineke 1044
30 mmeineke 1068 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 mmeineke 1054 (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 mmeineke 1068 system integrators (Sec.~\ref{oopseSec:noseHooverThermo}). Christopher
39 mmeineke 1054 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 mmeineke 1068 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 mmeineke 1044
57 mmeineke 1051 \section{\label{sec:intro}Introduction}
58 mmeineke 1044
59 mmeineke 1051 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 mmeineke 1054 more convenient for a researcher to turn to pre-existing simulation
65 mmeineke 1051 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 mmeineke 1054 to perhaps explore experimental analogues to their models.
70 mmeineke 1044
71 mmeineke 1051 Despite their utility, problems with these packages arise when
72     researchers try to develop techniques or energetic models that the
73 mmeineke 1068 code was not originally designed to simulate. Examples of uncommonly
74 mmeineke 1051 implemented techniques and energetics include; dipole-dipole
75 mmeineke 1054 interactions, rigid body dynamics, and metallic embedded
76 mmeineke 1051 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 mmeineke 1044
82 mmeineke 1068 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 mmeineke 1044
93 mmeineke 1054 We have structured this chapter to first discuss the empirical energy
94 mmeineke 1051 functions that {\sc oopse } implements in
95 mmeineke 1054 Sec.~\ref{oopseSec:empiricalEnergy}. Following that is a discussion of
96 mmeineke 1051 the various input and output files associated with the package
97 mmeineke 1068 (Sec.~\ref{oopseSec:IOfiles}). Sec.~\ref{oopseSec:mechanics}
98 mmeineke 1051 elucidates the various Molecular Dynamics algorithms {\sc oopse}
99 mmeineke 1054 implements in the integration of the Newtonian equations of
100 mmeineke 1051 motion. Basic analysis of the trajectories obtained from the
101     simulation is discussed in Sec.~\ref{oopseSec:props}. Program design
102 mmeineke 1068 considerations are presented in Sec.~\ref{oopseSec:design}. And
103     lastly, Sec.~\ref{oopseSec:conclusion} concludes the chapter.
104 mmeineke 1044
105 mmeineke 1051 \section{\label{oopseSec:empiricalEnergy}The Empirical Energy Functions}
106    
107     \subsection{\label{oopseSec:atomsMolecules}Atoms, Molecules and Rigid Bodies}
108    
109 mmeineke 1044 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 mmeineke 1054 dipoles). Charges, permanent dipoles, and Lennard-Jones parameters for
116 mmeineke 1068 a given atom type are set in the force field parameter files.
117 mmeineke 1044
118 mmeineke 1054 \begin{lstlisting}[float,caption={[Specifier for molecules and atoms] A sample specification of an Ar molecule},label=sch:AtmMole]
119 mmeineke 1044 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 mmeineke 1045
130 mmeineke 1054 Atoms can be collected into secondary structures such as rigid bodies
131 mmeineke 1044 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 mmeineke 1054 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 mmeineke 1068 \ref{sch:AtmMole} shows how one creates a molecule in a ``model'' or
137 mmeineke 1054 \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 mmeineke 1044
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 mmeineke 1068 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 mmeineke 1044
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 mmeineke 1068 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 mmeineke 1044 \begin{equation}
162     \boldsymbol{\tau}_i=
163 mmeineke 1068 \sum_{a}\biggl[(\mathbf{r}_{ia}-\mathbf{r}_i)\times \mathbf{f}_{ia}
164     + \boldsymbol{\tau}_{ia}\biggr]
165 mmeineke 1044 \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 mmeineke 1068 each rigid body. In order to move between the space fixed and body
174 mmeineke 1044 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 mmeineke 1068 entire nine parameter rotation matrix. Further discussion
189 mmeineke 1054 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 mmeineke 1044 \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 mmeineke 1045 \begin{lstlisting}[float,caption={[Defining rigid bodies]A sample definition of a rigid body},label={sch:rigidBody}]
196 mmeineke 1044 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 mmeineke 1054 \subsection{\label{sec:LJPot}The Lennard Jones Force Field}
220 mmeineke 1044
221     The most basic force field implemented in {\sc oopse} is the
222 mmeineke 1054 Lennard-Jones force field, which mimics the van der Waals interaction at
223 mmeineke 1044 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 mmeineke 1054 \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 mmeineke 1044
240 mmeineke 1045 \begin{lstlisting}[float,caption={[Invocation of the Lennard-Jones force field] A sample system using the Lennard-Jones force field.},label={sch:LJFF}]
241 mmeineke 1044
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 mmeineke 1068 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 mmeineke 1044 $2.5\sigma_{ii}$, 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 mmeineke 1068 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 mmeineke 1044
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 mmeineke 1051 \subsection{\label{oopseSec:DUFF}Dipolar Unified-Atom Force Field}
284 mmeineke 1044
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 mmeineke 1054 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 mmeineke 1044
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 mmeineke 1068 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 mmeineke 1044
301     As an example, lipid head-groups in {\sc duff} are represented as
302 mmeineke 1068 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 mmeineke 1044 \emph{et al.}\cite{liu96:new_model}
311    
312     \begin{figure}
313 mmeineke 1045 \centering
314     \includegraphics[width=\linewidth]{lipidModel.eps}
315 mmeineke 1083 \caption[A representation of a lipid model in {\sc duff}]{A representation of the lipid model. $\phi$ is the torsion angle, $\theta$ %
316 mmeineke 1044 is the bend angle, $\mu$ is the dipole moment of the head group, and n
317     is the chain length.}
318 mmeineke 1045 \label{oopseFig:lipidModel}
319 mmeineke 1044 \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 mmeineke 1068 of pseudoatoms to a minimum; the parameters for a unified atom such as
330 mmeineke 1044 $\text{CH}_2$ 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 mmeineke 1068 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 mmeineke 1044
341 mmeineke 1045 \begin{lstlisting}[float,caption={[Invocation of {\sc duff}]Sample \texttt{.bass} file showing a simulation utilizing {\sc duff}},label={sch:DUFF}]
342 mmeineke 1044
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 mmeineke 1051 \subsection{\label{oopseSec:energyFunctions}{\sc duff} Energy Functions}
364 mmeineke 1044
365     The total potential energy function in {\sc duff} is
366     \begin{equation}
367     V = \sum^{N}_{I=1} V^{I}_{\text{Internal}}
368 mmeineke 1054 + \sum^{N-1}_{I=1} \sum_{J>I} V^{IJ}_{\text{Cross}}
369 mmeineke 1044 \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 mmeineke 1054 (see Fig.~\ref{oopseFig:lipidModel}), $\theta_0$ is the equilibrium
396 mmeineke 1044 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 mmeineke 1068 Where:
409 mmeineke 1044 \begin{equation}
410 mmeineke 1068 \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 mmeineke 1044 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 mmeineke 1054 (Sec.~\ref{oopseSec:SSD}). Note that not all atom types include all
451 mmeineke 1044 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 mmeineke 1068 3(\boldsymbol{\hat{u}}_i \cdot \hat{\mathbf{r}}_{ij}) %
460     (\boldsymbol{\hat{u}}_j \cdot \hat{\mathbf{r}}_{ij}) \biggr]
461 mmeineke 1044 \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 mmeineke 1054 $i$, $\boldsymbol{\hat{u}}_i$ 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 mmeineke 1044
472 mmeineke 1068 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 mmeineke 1044
494 mmeineke 1068 \subsection{\label{oopseSec:SSD}The {\sc duff} Water Models: SSD/E and SSD/RF}
495 mmeineke 1044
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 mmeineke 1068 Since SSD/E is a single-point {\it dipolar} model, the force
555 mmeineke 1044 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 mmeineke 1068 properties. Indeed, SSD/E maintains reasonable agreement with the Head-Gordon
562 mmeineke 1044 diffraction data for the structural features of liquid
563 mmeineke 1068 water.\cite{hura00,liu96:new_model} Additionally, the dynamical properties
564     exhibited by SSD/E agree with experiment better than those of more
565 mmeineke 1044 computationally expensive models (like TIP3P and
566     SPC/E).\cite{chandra99:ssd_md} The combination of speed and accurate depiction
567 mmeineke 1068 of solvent properties makes SSD/E a very attractive model for the
568 mmeineke 1044 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 mmeineke 1068 \texttt{.bass} file as illustrated in the scheme below. A table of the
579 mmeineke 1044 parameter values and the drawbacks and benefits of the different
580 mmeineke 1054 density corrected SSD models can be found in
581     reference~\cite{Gezelter04}.
582 mmeineke 1044
583 mmeineke 1045 \begin{lstlisting}[float,caption={[A simulation of {\sc ssd} water]An example file showing a simulation including {\sc ssd} water.},label={sch:ssd}]
584 mmeineke 1044
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 mmeineke 1051 \subsection{\label{oopseSec:eam}Embedded Atom Method}
611 mmeineke 1044
612 mmeineke 1054 There are Molecular Dynamics packages which have the
613 mmeineke 1044 capacity to simulate metallic systems, including some that have
614     parallel computational abilities\cite{plimpton93}. Potentials that
615     describe bonding transition metal
616 mmeineke 1068 systems\cite{Finnis84,Ercolessi88,Chen90,Qi99,Ercolessi02} have an
617 mmeineke 1044 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 mmeineke 1068 the system. A mostly-repulsive pairwise part of the potential
621 mmeineke 1044 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 mmeineke 1068 good review of {\sc eam} and other metallic potential formulations was written
626 mmeineke 1044 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 mmeineke 1068 \end{eqnarray}
634 mmeineke 1083 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 mmeineke 1044 surrounding atom $i$ for both the density $\rho$ and pairwise $\phi$
645 mmeineke 1083 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 mmeineke 1044
649 mmeineke 1051 \subsection{\label{oopseSec:pbc}Periodic Boundary Conditions}
650 mmeineke 1044
651     \newcommand{\roundme}{\operatorname{round}}
652    
653 mmeineke 1068 \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 mmeineke 1083 use a $3 \times 3$ matrix, $\mathsf{H}$, to describe the shape and
663     size of the simulation box. $\mathsf{H}$ is defined:
664 mmeineke 1044 \begin{equation}
665 mmeineke 1083 \mathsf{H} = ( \mathbf{h}_x, \mathbf{h}_y, \mathbf{h}_z )
666 mmeineke 1044 \end{equation}
667 mmeineke 1068 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 mmeineke 1083 the box can be changed to allow volume fluctuations when constraining
670 mmeineke 1068 the pressure.
671 mmeineke 1044
672 mmeineke 1068 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 mmeineke 1083 \mathbf{s} &= \mathsf{H}^{-1} \mathbf{r} \\
676     \mathbf{r} &= \mathsf{H} \mathbf{s}
677 mmeineke 1068 \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 mmeineke 1044 \begin{equation}
684     s_{i}^{\prime}=s_{i}-\roundme(s_{i})
685     \end{equation}
686 mmeineke 1068 Where $s_i$ is the $i$th element of $\mathbf{s}$, and
687     $\roundme(s_i)$is given by
688 mmeineke 1044 \begin{equation}
689 mmeineke 1068 \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 mmeineke 1044 \end{equation}
695 mmeineke 1068 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 mmeineke 1044
701     Finally, we obtain the minimum image coordinates $\mathbf{r}^{\prime}$ by
702 mmeineke 1068 transforming back to real space,
703 mmeineke 1044 \begin{equation}
704 mmeineke 1083 \mathbf{r}^{\prime}=\mathsf{H}^{-1}\mathbf{s}^{\prime}%
705 mmeineke 1044 \end{equation}
706 mmeineke 1068 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 mmeineke 1083 the inter-atomic forces.
709 mmeineke 1044
710    
711 mmeineke 1051 \section{\label{oopseSec:IOfiles}Input and Output Files}
712 mmeineke 1044
713     \subsection{{\sc bass} and Model Files}
714    
715 mmeineke 1071 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 mmeineke 1044 \texttt{.bass}, an example file is shown in
722 mmeineke 1071 Scheme~\ref{sch:bassExample}.
723 mmeineke 1044
724 mmeineke 1071 \begin{lstlisting}[float,caption={[An example of a complete {\sc bass} file] An example showing a complete {\sc bass} file.},label={sch:bassExample}]
725 mmeineke 1044
726 mmeineke 1071 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 mmeineke 1083 ensemble = "NVE"; // specify the simulation ensemble
745 mmeineke 1071 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 mmeineke 1054 Within the \texttt{.bass} file it is necessary to provide a complete
753 mmeineke 1044 description of the molecule before it is actually placed in the
754 mmeineke 1071 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 mmeineke 1044 descriptions can become lengthy for complex molecules, and it would be
758 mmeineke 1071 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 mmeineke 1044
768 mmeineke 1071 \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 mmeineke 1051 \subsection{\label{oopseSec:coordFiles}Coordinate Files}
803 mmeineke 1044
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 mmeineke 1071 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 mmeineke 1044
814 mmeineke 1083 \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 mmeineke 1071
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 mmeineke 1054 \texttt{.eor} file stores the final configuration of the system for a
838 mmeineke 1044 given simulation. The file is updated at the same time as the
839 mmeineke 1071 \texttt{.dump} file, however, it only contains the most recent
840 mmeineke 1044 frame. In this way, an \texttt{.eor} file may be used as the
841 mmeineke 1071 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 mmeineke 1044
845 mmeineke 1054 \subsection{\label{oopseSec:initCoords}Generation of Initial Coordinates}
846 mmeineke 1044
847 mmeineke 1071 As was stated in Sec.~\ref{oopseSec:coordFiles}, an initialization
848     file is needed to provide the starting coordinates for a
849 mmeineke 1083 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 mmeineke 1071 arguments and parameters in the \texttt{.bass} file that would
853     otherwise be ignored by the simulation.
854 mmeineke 1044
855     \subsection{The Statistics File}
856    
857 mmeineke 1071 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 mmeineke 1044
868 mmeineke 1051 \section{\label{oopseSec:mechanics}Mechanics}
869 mmeineke 1044
870    
871 mmeineke 1087 \section{\label{sec:mechanics}Mechanics}
872 mmeineke 1044
873 mmeineke 1087 \subsection{\label{oopseSec:integrate}Integrating the Equations of Motion: the
874     DLM method}
875    
876     The default method for integrating the equations of motion in {\sc
877     oopse} is a velocity-Verlet version of the symplectic splitting method
878     proposed by Dullweber, Leimkuhler and McLachlan
879     (DLM).\cite{Dullweber1997} When there are no directional atoms or
880     rigid bodies present in the simulation, this integrator becomes the
881     standard velocity-Verlet integrator which is known to sample the
882     microcanonical (NVE) ensemble.\cite{}
883    
884     Previous integration methods for orientational motion have problems
885     that are avoided in the DLM method. Direct propagation of the Euler
886     angles has a known $1/\sin\theta$ divergence in the equations of
887     motion for $\phi$ and $\psi$,\cite{allen87:csl} leading to
888     numerical instabilities any time one of the directional atoms or rigid
889     bodies has an orientation near $\theta=0$ or $\theta=\pi$. More
890     modern quaternion-based integration methods have relatively poor
891     energy conservation. While quaternions work well for orientational
892     motion in other ensembles, the microcanonical ensemble has a
893     constant energy requirement that is quite sensitive to errors in the
894     equations of motion. An earlier implementation of {\sc oopse}
895     utilized quaternions for propagation of rotational motion; however, a
896     detailed investigation showed that they resulted in a steady drift in
897     the total energy, something that has been observed by
898     Laird {\it et al.}\cite{Laird97}
899    
900 mmeineke 1044 The key difference in the integration method proposed by Dullweber
901 mmeineke 1087 \emph{et al.} is that the entire $3 \times 3$ rotation matrix is
902     propagated from one time step to the next. In the past, this would not
903     have been feasible, since the rotation matrix for a single body has
904     nine elements compared with the more memory-efficient methods (using
905     three Euler angles or 4 quaternions). Computer memory has become much
906     less costly in recent years, and this can be translated into
907     substantial benefits in energy conservation.
908 mmeineke 1044
909 mmeineke 1087 The basic equations of motion being integrated are derived from the
910     Hamiltonian for conservative systems containing rigid bodies,
911     \begin{equation}
912     H = \sum_{i} \left( \frac{1}{2} m_i {\bf v}_i^T \cdot {\bf v}_i +
913     \frac{1}{2} {\bf j}_i^T \cdot \overleftrightarrow{\mathsf{I}}_i^{-1} \cdot
914     {\bf j}_i \right) +
915     V\left(\left\{{\bf r}\right\}, \left\{\mathsf{A}\right\}\right)
916     \end{equation}
917     where ${\bf r}_i$ and ${\bf v}_i$ are the cartesian position vector
918     and velocity of the center of mass of particle $i$, and ${\bf j}_i$
919     and $\overleftrightarrow{\mathsf{I}}_i$ are the body-fixed angular
920     momentum and moment of inertia tensor, respectively. $\mathsf{A}_i$
921     is the $3 \times 3$ rotation matrix describing the instantaneous
922     orientation of the particle. $V$ is the potential energy function
923     which may depend on both the positions $\left\{{\bf r}\right\}$ and
924     orientations $\left\{\mathsf{A}\right\}$ of all particles. The
925     equations of motion for the particle centers of mass are derived from
926     Hamilton's equations and are quite simple,
927     \begin{eqnarray}
928     \dot{{\bf r}} & = & {\bf v} \\
929     \dot{{\bf v}} & = & \frac{{\bf f}}{m}
930     \end{eqnarray}
931     where ${\bf f}$ is the instantaneous force on the center of mass
932     of the particle,
933     \begin{equation}
934     {\bf f} = - \frac{\partial}{\partial
935     {\bf r}} V(\left\{{\bf r}(t)\right\}, \left\{\mathsf{A}(t)\right\}).
936     \end{equation}
937 mmeineke 1044
938 mmeineke 1087 The equations of motion for the orientational degrees of freedom are
939     \begin{eqnarray}
940     \dot{\mathsf{A}} & = & \mathsf{A} \cdot
941     \mbox{ skew}\left(\overleftrightarrow{\mathsf{I}}^{-1} \cdot {\bf j}\right) \\
942     \dot{{\bf j}} & = & {\bf j} \times \left( \overleftrightarrow{\mathsf{I}}^{-1}
943     \cdot {\bf j} \right) - \mbox{ rot}\left(\mathsf{A}^{T} \cdot \frac{\partial
944     V}{\partial \mathsf{A}} \right)
945     \end{eqnarray}
946     In these equations of motion, the $\mbox{skew}$ matrix of a vector
947     ${\bf v} = \left( v_1, v_2, v_3 \right)$ is defined:
948     \begin{equation}
949     \mbox{skew}\left( {\bf v} \right) := \left(
950     \begin{array}{ccc}
951     0 & v_3 & - v_2 \\
952     -v_3 & 0 & v_1 \\
953     v_2 & -v_1 & 0
954     \end{array}
955     \right)
956     \end{equation}
957     The $\mbox{rot}$ notation refers to the mapping of the $3 \times 3$
958     rotation matrix to a vector of orientations by first computing the
959     skew-symmetric part $\left(\mathsf{A} - \mathsf{A}^{T}\right)$ and
960     then associating this with a length 3 vector by inverting the
961     $\mbox{skew}$ function above:
962     \begin{equation}
963     \mbox{rot}\left(\mathsf{A}\right) := \mbox{ skew}^{-1}\left(\mathsf{A}
964     - \mathsf{A}^{T} \right)
965     \end{equation}
966     Written this way, the $\mbox{rot}$ operation creates a set of
967     conjugate angle coordinates to the body-fixed angular momenta
968     represented by ${\bf j}$. This equation of motion for angular momenta
969     is equivalent to the more familiar body-fixed forms,
970     \begin{eqnarray}
971     \dot{j_{x}} & = & \tau^b_x(t) +
972     \left(\overleftrightarrow{\mathsf{I}}_{yy} - \overleftrightarrow{\mathsf{I}}_{zz} \right) j_y j_z \\
973     \dot{j_{y}} & = & \tau^b_y(t) +
974     \left(\overleftrightarrow{\mathsf{I}}_{zz} - \overleftrightarrow{\mathsf{I}}_{xx} \right) j_z j_x \\
975     \dot{j_{z}} & = & \tau^b_z(t) +
976     \left(\overleftrightarrow{\mathsf{I}}_{xx} - \overleftrightarrow{\mathsf{I}}_{yy} \right) j_x j_y
977     \end{eqnarray}
978     which utilize the body-fixed torques, ${\bf \tau}^b$. Torques are
979     most easily derived in the space-fixed frame,
980     \begin{equation}
981     {\bf \tau}^b(t) = \mathsf{A}(t) \cdot {\bf \tau}^s(t)
982     \end{equation}
983     where the torques are either derived from the forces on the
984     constituent atoms of the rigid body, or for directional atoms,
985     directly from derivatives of the potential energy,
986     \begin{equation}
987     {\bf \tau}^s(t) = - \hat{\bf u}(t) \times \left( \frac{\partial}
988     {\partial \hat{\bf u}} V\left(\left\{ {\bf r}(t) \right\}, \left\{
989     \mathsf{A}(t) \right\}\right) \right).
990     \end{equation}
991     Here $\hat{\bf u}$ is a unit vector pointing along the principal axis
992     of the particle in the space-fixed frame.
993    
994     The DLM method uses a Trotter factorization of the orientational
995     propagator. This has three effects:
996     \begin{enumerate}
997     \item the integrator is area-preserving in phase space (i.e. it is
998     {\it symplectic}),
999     \item the integrator is time-{\it reversible}, making it suitable for Hybrid
1000     Monte Carlo applications, and
1001     \item the error for a single time step is of order $O\left(h^3\right)$
1002     for timesteps of length $h$.
1003     \end{enumerate}
1004    
1005     The integration of the equations of motion is carried out in a
1006     velocity-Verlet style 2-part algorithm:
1007    
1008     {\tt moveA:}
1009     \begin{eqnarray}
1010     {\bf v}\left(t + \delta t / 2\right) & \leftarrow & {\bf
1011     v}(t) + \frac{\delta t}{2} \left( {\bf f}(t) / m \right) \\
1012     {\bf r}(t + \delta t) & \leftarrow & {\bf r}(t) + \delta t {\bf
1013     v}\left(t + \delta t / 2 \right) \\
1014     {\bf j}\left(t + \delta t / 2 \right) & \leftarrow & {\bf
1015     j}(t) + \frac{\delta t}{2} {\bf \tau}^b(t) \\
1016     \mathsf{A}(t + \delta t) & \leftarrow & \mathrm{rot}\left( \delta t
1017     {\bf j}(t + \delta t / 2) \cdot \overleftrightarrow{\mathsf{I}}^{-1}
1018     \right)
1019     \end{eqnarray}
1020    
1021     In this context, the $\mathrm{rot}$ function is the reversible product
1022     of the three body-fixed rotations,
1023     \begin{equation}
1024     \mathrm{rot}({\bf a}) = \mathsf{G}_x(a_x / 2) \cdot
1025     \mathsf{G}_y(a_y / 2) \cdot \mathsf{G}_z(a_z) \cdot \mathsf{G}_y(a_y /
1026     2) \cdot \mathsf{G}_x(a_x /2)
1027     \end{equation}
1028     where each rotational propagator, $\mathsf{G}_\alpha(\theta)$, rotates
1029     both the rotation matrix ($\mathsf{A}$) and the body-fixed angular
1030     momentum (${\bf j}$) by an angle $\theta$ around body-fixed axis
1031     $\alpha$,
1032     \begin{equation}
1033     \mathsf{G}_\alpha( \theta ) = \left\{
1034     \begin{array}{lcl}
1035     \mathsf{A}(t) & \leftarrow & \mathsf{A}(0) \cdot \mathsf{R}_\alpha(\theta)^T \\
1036     {\bf j}(t) & \leftarrow & \mathsf{R}_\alpha(\theta) \cdot {\bf j}(0)
1037     \end{array}
1038     \right.
1039     \end{equation}
1040     $\mathsf{R}_\alpha$ is a quadratic approximation to
1041     the single-axis rotation matrix. For example, in the small-angle
1042     limit, the rotation matrix around the body-fixed x-axis can be
1043     approximated as
1044     \begin{equation}
1045     \mathsf{R}_x(\theta) \approx \left(
1046     \begin{array}{ccc}
1047     1 & 0 & 0 \\
1048     0 & \frac{1-\theta^2 / 4}{1 + \theta^2 / 4} & -\frac{\theta}{1+
1049     \theta^2 / 4} \\
1050     0 & \frac{\theta}{1+
1051     \theta^2 / 4} & \frac{1-\theta^2 / 4}{1 + \theta^2 / 4}
1052     \end{array}
1053     \right).
1054     \end{equation}
1055     All other rotations follow in a straightforward manner.
1056    
1057     After the first part of the propagation, the forces and body-fixed
1058     torques are calculated at the new positions and orientations
1059    
1060     {\tt doForces:}
1061     \begin{eqnarray}
1062     {\bf f}(t + \delta t) & \leftarrow & - \left(\frac{\partial V}{\partial {\bf
1063     r}}\right)_{{\bf r}(t + \delta t)} \\
1064     {\bf \tau}^{s}(t + \delta t) & \leftarrow & {\bf u}(t + \delta t)
1065     \times \frac{\partial V}{\partial {\bf u}} \\
1066     {\bf \tau}^{b}(t + \delta t) & \leftarrow & \mathsf{A}(t + \delta t)
1067     \cdot {\bf \tau}^s(t + \delta t)
1068     \end{eqnarray}
1069    
1070     {\sc oopse} automatically updates ${\bf u}$ when the rotation matrix
1071     $\mathsf{A}$ is calculated in {\tt moveA}. Once the forces and
1072     torques have been obtained at the new time step, the velocities can be
1073     advanced to the same time value.
1074    
1075     {\tt moveB:}
1076     \begin{eqnarray}
1077     {\bf v}\left(t + \delta t \right) & \leftarrow & {\bf
1078     v}\left(t + \delta t / 2 \right) + \frac{\delta t}{2} \left(
1079     {\bf f}(t + \delta t) / m \right) \\
1080     {\bf j}\left(t + \delta t \right) & \leftarrow & {\bf
1081     j}\left(t + \delta t / 2 \right) + \frac{\delta t}{2} {\bf
1082     \tau}^b(t + \delta t)
1083     \end{eqnarray}
1084    
1085     The matrix rotations used in the DLM method end up being more costly
1086     computationally than the simpler arithmetic quaternion
1087     propagation. With the same time step, a 1000-molecule water simulation
1088     shows an average 7\% increase in computation time using the DLM method
1089     in place of quaternions. This cost is more than justified when
1090     comparing the energy conservation of the two methods as illustrated in
1091     figure \ref{timestep}.
1092    
1093 mmeineke 1044 \begin{figure}
1094 mmeineke 1045 \centering
1095     \includegraphics[width=\linewidth]{timeStep.eps}
1096 mmeineke 1087 \caption[Energy conservation for quaternion versus DLM dynamics]{Energy conservation using quaternion based integration versus
1097     the method proposed by Dullweber \emph{et al.} with increasing time
1098     step. For each time step, the dotted line is total energy using the
1099     DLM integrator, and the solid line comes from the quaternion
1100     integrator. The larger time step plots are shifted up from the true
1101     energy baseline for clarity.}
1102 mmeineke 1044 \label{timestep}
1103     \end{figure}
1104    
1105 mmeineke 1087 In figure \ref{timestep}, the resulting energy drift at various time
1106     steps for both the DLM and quaternion integration schemes is
1107     compared. All of the 1000 molecule water simulations started with the
1108 mmeineke 1044 same configuration, and the only difference was the method for
1109     handling rotational motion. At time steps of 0.1 and 0.5 fs, both
1110 mmeineke 1087 methods for propagating molecule rotation conserve energy fairly well,
1111 mmeineke 1044 with the quaternion method showing a slight energy drift over time in
1112     the 0.5 fs time step simulation. At time steps of 1 and 2 fs, the
1113 mmeineke 1087 energy conservation benefits of the DLM method are clearly
1114 mmeineke 1044 demonstrated. Thus, while maintaining the same degree of energy
1115     conservation, one can take considerably longer time steps, leading to
1116     an overall reduction in computation time.
1117    
1118 mmeineke 1087 There is only one specific keyword relevant to the default integrator,
1119     and that is the time step for integrating the equations of motion.
1120 mmeineke 1044
1121 mmeineke 1087 \begin{center}
1122     \begin{tabular}{llll}
1123     {\bf variable} & {\bf {\tt .bass} keyword} & {\bf units} & {\bf
1124     default value} \\
1125     $\delta t$ & {\tt dt = 2.0;} & fs & none
1126     \end{tabular}
1127     \end{center}
1128 mmeineke 1044
1129     \subsection{\label{sec:extended}Extended Systems for other Ensembles}
1130    
1131 mmeineke 1087 {\sc oopse} implements a number of extended system integrators for
1132     sampling from other ensembles relevant to chemical physics. The
1133     integrator can selected with the {\tt ensemble} keyword in the
1134     {\tt .bass} file:
1135 mmeineke 1044
1136 mmeineke 1087 \begin{center}
1137     \begin{tabular}{lll}
1138     {\bf Integrator} & {\bf Ensemble} & {\bf {\tt .bass} line} \\
1139     NVE & microcanonical & {\tt ensemble = ``NVE''; } \\
1140     NVT & canonical & {\tt ensemble = ``NVT''; } \\
1141     NPTi & isobaric-isothermal (with isotropic volume changes) & {\tt
1142     ensemble = ``NPTi'';} \\
1143     NPTf & isobaric-isothermal (with changes to box shape) & {\tt
1144     ensemble = ``NPTf'';} \\
1145     NPTxyz & approximate isobaric-isothermal & {\tt ensemble =
1146     ``NPTxyz'';} \\
1147     & (with separate barostats on each box dimension) &
1148     \end{tabular}
1149     \end{center}
1150 mmeineke 1044
1151 mmeineke 1087 The relatively well-known Nos\'e-Hoover thermostat is implemented in
1152     {\sc oopse}'s NVT integrator. This method couples an extra degree of
1153     freedom (the thermostat) to the kinetic energy of the system, and has
1154     been shown to sample the canonical distribution in the system degrees
1155     of freedom while conserving a quantity that is, to within a constant,
1156     the Helmholtz free energy.
1157 mmeineke 1044
1158 mmeineke 1087 NPT algorithms attempt to maintain constant pressure in the system by
1159     coupling the volume of the system to a barostat. {\sc oopse} contains
1160     three different constant pressure algorithms. The first two, NPTi and
1161     NPTf have been shown to conserve a quantity that is, to within a
1162     constant, the Gibbs free energy. The Melchionna modification to the
1163     Hoover barostat is implemented in both NPTi and NPTf. NPTi allows
1164     only isotropic changes in the simulation box, while box {\it shape}
1165     variations are allowed in NPTf. The NPTxyz integrator has {\it not}
1166     been shown to sample from the isobaric-isothermal ensemble. It is
1167     useful, however, in that it maintains orthogonality for the axes of
1168     the simulation box while attempting to equalize pressure along the
1169     three perpendicular directions in the box.
1170 mmeineke 1044
1171 mmeineke 1087 Each of the extended system integrators requires additional keywords
1172     to set target values for the thermodynamic state variables that are
1173     being held constant. Keywords are also required to set the
1174     characteristic decay times for the dynamics of the extended
1175     variables.
1176    
1177     \begin{tabular}{llll}
1178     {\bf variable} & {\bf {\tt .bass} keyword} & {\bf units} & {\bf
1179     default value} \\
1180     $T_{\mathrm{target}}$ & {\tt targetTemperature = 300;} & K & none \\
1181     $P_{\mathrm{target}}$ & {\tt targetPressure = 1;} & atm & none \\
1182     $\tau_T$ & {\tt tauThermostat = 1e3;} & fs & none \\
1183     $\tau_B$ & {\tt tauBarostat = 5e3;} & fs & none \\
1184     & {\tt resetTime = 200;} & fs & none \\
1185     & {\tt useInitialExtendedSystemState = ``true'';} & logical &
1186     false
1187     \end{tabular}
1188    
1189     Two additional keywords can be used to either clear the extended
1190     system variables periodically ({\tt resetTime}), or to maintain the
1191     state of the extended system variables between simulations ({\tt
1192     useInitialExtendedSystemState}). More details on these variables
1193     and their use in the integrators follows below.
1194    
1195     \subsubsection{\label{oopseSec:noseHooverThermo}Nos\'{e}-Hoover Thermostatting}
1196    
1197     The Nos\'e-Hoover equations of motion are given by\cite{Hoover85}
1198 mmeineke 1044 \begin{eqnarray}
1199     \dot{{\bf r}} & = & {\bf v} \\
1200 mmeineke 1087 \dot{{\bf v}} & = & \frac{{\bf f}}{m} - \chi {\bf v} \\
1201     \dot{\mathsf{A}} & = & \mathsf{A} \cdot
1202     \mbox{ skew}\left(\overleftrightarrow{\mathsf{I}}^{-1} \cdot {\bf j}\right) \\
1203     \dot{{\bf j}} & = & {\bf j} \times \left( \overleftrightarrow{\mathsf{I}}^{-1}
1204     \cdot {\bf j} \right) - \mbox{ rot}\left(\mathsf{A}^{T} \cdot \frac{\partial
1205     V}{\partial \mathsf{A}} \right) - \chi {\bf j}
1206 mmeineke 1044 \label{eq:nosehoovereom}
1207     \end{eqnarray}
1208    
1209     $\chi$ is an ``extra'' variable included in the extended system, and
1210     it is propagated using the first order equation of motion
1211     \begin{equation}
1212 mmeineke 1087 \dot{\chi} = \frac{1}{\tau_{T}^2} \left( \frac{T}{T_{\mathrm{target}}} - 1 \right).
1213 mmeineke 1044 \label{eq:nosehooverext}
1214     \end{equation}
1215    
1216 mmeineke 1087 The instantaneous temperature $T$ is proportional to the total kinetic
1217     energy (both translational and orientational) and is given by
1218     \begin{equation}
1219     T = \frac{2 K}{f k_B}
1220     \end{equation}
1221     Here, $f$ is the total number of degrees of freedom in the system,
1222     \begin{equation}
1223     f = 3 N + 3 N_{\mathrm{orient}} - N_{\mathrm{constraints}}
1224     \end{equation}
1225     and $K$ is the total kinetic energy,
1226     \begin{equation}
1227     K = \sum_{i=1}^{N} \frac{1}{2} m_i {\bf v}_i^T \cdot {\bf v}_i +
1228     \sum_{i=1}^{N_{\mathrm{orient}}} \frac{1}{2} {\bf j}_i^T \cdot
1229     \overleftrightarrow{\mathsf{I}}_i^{-1} \cdot {\bf j}_i
1230     \end{equation}
1231 mmeineke 1044
1232 mmeineke 1087 In eq.(\ref{eq:nosehooverext}), $\tau_T$ is the time constant for
1233     relaxation of the temperature to the target value. To set values for
1234     $\tau_T$ or $T_{\mathrm{target}}$ in a simulation, one would use the
1235     {\tt tauThermostat} and {\tt targetTemperature} keywords in the {\tt
1236     .bass} file. The units for {\tt tauThermostat} are fs, and the units
1237     for the {\tt targetTemperature} are degrees K. The integration of
1238     the equations of motion is carried out in a velocity-Verlet style 2
1239     part algorithm:
1240    
1241     {\tt moveA:}
1242     \begin{eqnarray}
1243     T(t) & \leftarrow & \left\{{\bf v}(t)\right\}, \left\{{\bf j}(t)\right\} \\
1244     {\bf v}\left(t + \delta t / 2\right) & \leftarrow & {\bf
1245     v}(t) + \frac{\delta t}{2} \left( \frac{{\bf f}(t)}{m} - {\bf v}(t)
1246     \chi(t)\right) \\
1247     {\bf r}(t + \delta t) & \leftarrow & {\bf r}(t) + \delta t {\bf
1248     v}\left(t + \delta t / 2 \right) \\
1249     {\bf j}\left(t + \delta t / 2 \right) & \leftarrow & {\bf
1250     j}(t) + \frac{\delta t}{2} \left( {\bf \tau}^b(t) - {\bf j}(t)
1251     \chi(t) \right) \\
1252     \mathsf{A}(t + \delta t) & \leftarrow & \mathrm{rot}\left(\delta t *
1253     {\bf j}(t + \delta t / 2) \overleftrightarrow{\mathsf{I}}^{-1} \right) \\
1254     \chi\left(t + \delta t / 2 \right) & \leftarrow & \chi(t) +
1255     \frac{\delta t}{2 \tau_T^2} \left( \frac{T(t)}{T_{\mathrm{target}}} - 1
1256     \right)
1257     \end{eqnarray}
1258    
1259     Here $\mathrm{rot}(\delta t * {\bf j}
1260     \overleftrightarrow{\mathsf{I}}^{-1})$ is the same symplectic Trotter
1261     factorization of the three rotation operations that was discussed in
1262     the section on the DLM integrator. Note that this operation modifies
1263     both the rotation matrix $\mathsf{A}$ and the angular momentum ${\bf
1264     j}$. {\tt moveA} propagates velocities by a half time step, and
1265     positional degrees of freedom by a full time step. The new positions
1266     (and orientations) are then used to calculate a new set of forces and
1267     torques in exactly the same way they are calculated in the {\tt
1268     doForces} portion of the DLM integrator.
1269    
1270     Once the forces and torques have been obtained at the new time step,
1271     the temperature, velocities, and the extended system variable can be
1272     advanced to the same time value.
1273    
1274     {\tt moveB:}
1275     \begin{eqnarray}
1276     T(t + \delta t) & \leftarrow & \left\{{\bf v}(t + \delta t)\right\},
1277     \left\{{\bf j}(t + \delta t)\right\} \\
1278     \chi\left(t + \delta t \right) & \leftarrow & \chi\left(t + \delta t /
1279     2 \right) + \frac{\delta t}{2 \tau_T^2} \left( \frac{T(t+\delta
1280     t)}{T_{\mathrm{target}}} - 1 \right) \\
1281     {\bf v}\left(t + \delta t \right) & \leftarrow & {\bf
1282     v}\left(t + \delta t / 2 \right) + \frac{\delta t}{2} \left(
1283     \frac{{\bf f}(t + \delta t)}{m} - {\bf v}(t + \delta t)
1284     \chi(t \delta t)\right) \\
1285     {\bf j}\left(t + \delta t \right) & \leftarrow & {\bf
1286     j}\left(t + \delta t / 2 \right) + \frac{\delta t}{2} \left( {\bf
1287     \tau}^b(t + \delta t) - {\bf j}(t + \delta t)
1288     \chi(t + \delta t) \right)
1289     \end{eqnarray}
1290    
1291     Since ${\bf v}(t + \delta t)$ and ${\bf j}(t + \delta t)$ are required
1292     to caclculate $T(t + \delta t)$ as well as $\chi(t + \delta t)$, they
1293     indirectly depend on their own values at time $t + \delta t$. {\tt
1294     moveB} is therefore done in an iterative fashion until $\chi(t +
1295     \delta t)$ becomes self-consistent. The relative tolerance for the
1296     self-consistency check defaults to a value of $\mbox{10}^{-6}$, but
1297     {\sc oopse} will terminate the iteration after 4 loops even if the
1298     consistency check has not been satisfied.
1299    
1300     The Nos\'e-Hoover algorithm is known to conserve a Hamiltonian for the
1301     extended system that is, to within a constant, identical to the
1302     Helmholtz free energy,
1303     \begin{equation}
1304     H_{\mathrm{NVT}} = V + K + f k_B T_{\mathrm{target}} \left(
1305     \frac{\tau_{T}^2 \chi^2(t)}{2} + \int_{0}^{t} \chi(t^\prime) dt^\prime
1306     \right)
1307     \end{equation}
1308     Poor choices of $\delta t$ or $\tau_T$ can result in non-conservation
1309     of $H_{\mathrm{NVT}}$, so the conserved quantity is maintained in the
1310     last column of the {\tt .stat} file to allow checks on the quality of
1311     the integration.
1312    
1313     Bond constraints are applied at the end of both the {\tt moveA} and
1314     {\tt moveB} portions of the algorithm. Details on the constraint
1315     algorithms are given in section \ref{oopseSec:rattle}.
1316    
1317     \subsubsection{\label{sec:NPTi}Constant-pressure integration with
1318     isotropic box deformations (NPTi)}
1319    
1320     To carry out isobaric-isothermal ensemble calculations {\sc oopse}
1321     implements the Melchionna modifications to the Nos\'e-Hoover-Andersen
1322     equations of motion,\cite{melchionna93}
1323    
1324     \begin{eqnarray}
1325     \dot{{\bf r}} & = & {\bf v} + \eta \left( {\bf r} - {\bf R}_0 \right) \\
1326     \dot{{\bf v}} & = & \frac{{\bf f}}{m} - (\eta + \chi) {\bf v} \\
1327     \dot{\mathsf{A}} & = & \mathsf{A} \cdot
1328     \mbox{ skew}\left(\overleftrightarrow{I}^{-1} \cdot {\bf j}\right) \\
1329     \dot{{\bf j}} & = & {\bf j} \times \left( \overleftrightarrow{I}^{-1}
1330     \cdot {\bf j} \right) - \mbox{ rot}\left(\mathsf{A}^{T} \cdot \frac{\partial
1331     V}{\partial \mathsf{A}} \right) - \chi {\bf j} \\
1332     \dot{\chi} & = & \frac{1}{\tau_{T}^2} \left(
1333     \frac{T}{T_{\mathrm{target}}} - 1 \right) \\
1334     \dot{\eta} & = & \frac{1}{\tau_{B}^2 f k_B T_{\mathrm{target}}} V \left( P -
1335     P_{\mathrm{target}} \right) \\
1336     \dot{\mathcal{V}} & = & 3 \mathcal{V} \eta
1337     \label{eq:melchionna1}
1338     \end{eqnarray}
1339    
1340     $\chi$ and $\eta$ are the ``extra'' degrees of freedom in the extended
1341     system. $\chi$ is a thermostat, and it has the same function as it
1342     does in the Nos\'e-Hoover NVT integrator. $\eta$ is a barostat which
1343     controls changes to the volume of the simulation box. ${\bf R}_0$ is
1344     the location of the center of mass for the entire system, and
1345     $\mathcal{V}$ is the volume of the simulation box. At any time, the
1346     volume can be calculated from the determinant of the matrix which
1347     describes the box shape:
1348     \begin{equation}
1349     \mathcal{V} = \det(\mathsf{H})
1350     \end{equation}
1351    
1352     The NPTi integrator requires an instantaneous pressure. This quantity
1353     is calculated via the pressure tensor,
1354     \begin{equation}
1355     \overleftrightarrow{\mathsf{P}}(t) = \frac{1}{\mathcal{V}(t)} \left(
1356     \sum_{i=1}^{N} m_i {\bf v}_i(t) \otimes {\bf v}_i(t) \right) +
1357     \overleftrightarrow{\mathsf{W}}(t)
1358     \end{equation}
1359     The kinetic contribution to the pressure tensor utilizes the {\it
1360     outer} product of the velocities denoted by the $\otimes$ symbol. The
1361     stress tensor is calculated from another outer product of the
1362     inter-atomic separation vectors (${\bf r}_{ij} = {\bf r}_j - {\bf
1363     r}_i$) with the forces between the same two atoms,
1364     \begin{equation}
1365     \overleftrightarrow{\mathsf{W}}(t) = \sum_{i} \sum_{j>i} {\bf r}_{ij}(t)
1366     \otimes {\bf f}_{ij}(t)
1367     \end{equation}
1368     The instantaneous pressure is then simply obtained from the trace of
1369     the Pressure tensor,
1370     \begin{equation}
1371     P(t) = \frac{1}{3} \mathrm{Tr} \left( \overleftrightarrow{\mathsf{P}}(t)
1372     \right)
1373     \end{equation}
1374    
1375     In eq.(\ref{eq:melchionna1}), $\tau_B$ is the time constant for
1376     relaxation of the pressure to the target value. To set values for
1377     $\tau_B$ or $P_{\mathrm{target}}$ in a simulation, one would use the
1378     {\tt tauBarostat} and {\tt targetPressure} keywords in the {\tt .bass}
1379     file. The units for {\tt tauBarostat} are fs, and the units for the
1380     {\tt targetPressure} are atmospheres. Like in the NVT integrator, the
1381     integration of the equations of motion is carried out in a
1382     velocity-Verlet style 2 part algorithm:
1383    
1384     {\tt moveA:}
1385     \begin{eqnarray}
1386     T(t) & \leftarrow & \left\{{\bf v}(t)\right\}, \left\{{\bf j}(t)\right\} \\
1387     P(t) & \leftarrow & \left\{{\bf r}(t)\right\}, \left\{{\bf v}(t)\right\}, \left\{{\bf f}(t)\right\} \\
1388     {\bf v}\left(t + \delta t / 2\right) & \leftarrow & {\bf
1389     v}(t) + \frac{\delta t}{2} \left( \frac{{\bf f}(t)}{m} - {\bf v}(t)
1390     \left(\chi(t) + \eta(t) \right) \right) \\
1391     {\bf j}\left(t + \delta t / 2 \right) & \leftarrow & {\bf
1392     j}(t) + \frac{\delta t}{2} \left( {\bf \tau}^b(t) - {\bf j}(t)
1393     \chi(t) \right) \\
1394     \mathsf{A}(t + \delta t) & \leftarrow & \mathrm{rot}\left(\delta t *
1395     {\bf j}(t + \delta t / 2) \overleftrightarrow{\mathsf{I}}^{-1} \right) \\
1396     \chi\left(t + \delta t / 2 \right) & \leftarrow & \chi(t) +
1397     \frac{\delta t}{2 \tau_T^2} \left( \frac{T(t)}{T_{\mathrm{target}}} - 1
1398     \right) \\
1399     \eta(t + \delta t / 2) & \leftarrow & \eta(t) + \frac{\delta t \mathcal{V}(t)}{2 N k_B
1400     T(t) \tau_B^2} \left( P(t) - P_{\mathrm{target}} \right) \\
1401     {\bf r}(t + \delta t) & \leftarrow & {\bf r}(t) + \delta t \left\{ {\bf
1402     v}\left(t + \delta t / 2 \right) + \eta(t + \delta t / 2)\left[ {\bf
1403     r}(t + \delta t) - {\bf R}_0 \right] \right\} \\
1404     \mathsf{H}(t + \delta t) & \leftarrow & e^{-\delta t \eta(t + \delta t
1405     / 2)} \mathsf{H}(t)
1406     \end{eqnarray}
1407    
1408     Most of these equations are identical to their counterparts in the NVT
1409     integrator, but the propagation of positions to time $t + \delta t$
1410     depends on the positions at the same time. {\sc oopse} carries out
1411     this step iteratively (with a limit of 5 passes through the iterative
1412     loop). Also, the simulation box $\mathsf{H}$ is scaled uniformly for
1413     one full time step by an exponential factor that depends on the value
1414     of $\eta$ at time $t +
1415     \delta t / 2$. Reshaping the box uniformly also scales the volume of
1416     the box by
1417     \begin{equation}
1418     \mathcal{V}(t + \delta t) \leftarrow e^{ - 3 \delta t \eta(t + \delta t /2)}
1419     \mathcal{V}(t)
1420     \end{equation}
1421    
1422     The {\tt doForces} step for the NPTi integrator is exactly the same as
1423     in both the DLM and NVT integrators. Once the forces and torques have
1424     been obtained at the new time step, the velocities can be advanced to
1425     the same time value.
1426    
1427     {\tt moveB:}
1428     \begin{eqnarray}
1429     T(t + \delta t) & \leftarrow & \left\{{\bf v}(t + \delta t)\right\},
1430     \left\{{\bf j}(t + \delta t)\right\} \\
1431     P(t + \delta t) & \leftarrow & \left\{{\bf r}(t + \delta t)\right\},
1432     \left\{{\bf v}(t + \delta t)\right\}, \left\{{\bf f}(t + \delta t)\right\} \\
1433     \chi\left(t + \delta t \right) & \leftarrow & \chi\left(t + \delta t /
1434     2 \right) + \frac{\delta t}{2 \tau_T^2} \left( \frac{T(t+\delta
1435     t)}{T_{\mathrm{target}}} - 1 \right) \\
1436     \eta(t + \delta t) & \leftarrow & \eta(t + \delta t / 2) +
1437     \frac{\delta t \mathcal{V}(t + \delta t)}{2 N k_B T(t + \delta t) \tau_B^2}
1438     \left( P(t + \delta t) - P_{\mathrm{target}}
1439     \right) \\
1440     {\bf v}\left(t + \delta t \right) & \leftarrow & {\bf
1441     v}\left(t + \delta t / 2 \right) + \frac{\delta t}{2} \left(
1442     \frac{{\bf f}(t + \delta t)}{m} - {\bf v}(t + \delta t)
1443     (\chi(t + \delta t) + \eta(t + \delta t)) \right) \\
1444     {\bf j}\left(t + \delta t \right) & \leftarrow & {\bf
1445     j}\left(t + \delta t / 2 \right) + \frac{\delta t}{2} \left( {\bf
1446     \tau}^b(t + \delta t) - {\bf j}(t + \delta t)
1447     \chi(t + \delta t) \right)
1448     \end{eqnarray}
1449    
1450     Once again, since ${\bf v}(t + \delta t)$ and ${\bf j}(t + \delta t)$
1451     are required to caclculate $T(t + \delta t)$, $P(t + \delta t)$, $\chi(t +
1452     \delta t)$, and $\eta(t + \delta t)$, they indirectly depend on their
1453     own values at time $t + \delta t$. {\tt moveB} is therefore done in
1454     an iterative fashion until $\chi(t + \delta t)$ and $\eta(t + \delta
1455     t)$ become self-consistent. The relative tolerance for the
1456     self-consistency check defaults to a value of $\mbox{10}^{-6}$, but
1457     {\sc oopse} will terminate the iteration after 4 loops even if the
1458     consistency check has not been satisfied.
1459    
1460     The Melchionna modification of the Nos\'e-Hoover-Andersen algorithm is
1461     known to conserve a Hamiltonian for the extended system that is, to
1462     within a constant, identical to the Gibbs free energy,
1463     \begin{equation}
1464     H_{\mathrm{NPTi}} = V + K + f k_B T_{\mathrm{target}} \left(
1465     \frac{\tau_{T}^2 \chi^2(t)}{2} + \int_{0}^{t} \chi(t^\prime) dt^\prime
1466     \right) + P_{\mathrm{target}} \mathcal{V}(t).
1467     \end{equation}
1468     Poor choices of $\delta t$, $\tau_T$, or $\tau_B$ can result in
1469     non-conservation of $H_{\mathrm{NPTi}}$, so the conserved quantity is
1470     maintained in the last column of the {\tt .stat} file to allow checks
1471     on the quality of the integration. It is also known that this
1472     algorithm samples the equilibrium distribution for the enthalpy
1473     (including contributions for the thermostat and barostat),
1474     \begin{equation}
1475     H_{\mathrm{NPTi}} = V + K + \frac{f k_B T_{\mathrm{target}}}{2} \left(
1476     \chi^2 \tau_T^2 + \eta^2 \tau_B^2 \right) + P_{\mathrm{target}}
1477     \mathcal{V}(t).
1478     \end{equation}
1479    
1480     Bond constraints are applied at the end of both the {\tt moveA} and
1481     {\tt moveB} portions of the algorithm. Details on the constraint
1482     algorithms are given in section \ref{oopseSec:rattle}.
1483    
1484     \subsubsection{\label{sec:NPTf}Constant-pressure integration with a
1485     flexible box (NPTf)}
1486    
1487     There is a relatively simple generalization of the
1488     Nos\'e-Hoover-Andersen method to include changes in the simulation box
1489     {\it shape} as well as in the volume of the box. This method utilizes
1490     the full $3 \times 3$ pressure tensor and introduces a tensor of
1491     extended variables ($\overleftrightarrow{\eta}$) to control changes to
1492     the box shape. The equations of motion for this method are
1493     \begin{eqnarray}
1494     \dot{{\bf r}} & = & {\bf v} + \overleftrightarrow{\eta} \cdot \left( {\bf r} - {\bf R}_0 \right) \\
1495     \dot{{\bf v}} & = & \frac{{\bf f}}{m} - (\overleftrightarrow{\eta} +
1496     \chi \mathsf{1}) {\bf v} \\
1497     \dot{\mathsf{A}} & = & \mathsf{A} \cdot
1498     \mbox{ skew}\left(\overleftrightarrow{I}^{-1} \cdot {\bf j}\right) \\
1499     \dot{{\bf j}} & = & {\bf j} \times \left( \overleftrightarrow{I}^{-1}
1500     \cdot {\bf j} \right) - \mbox{ rot}\left(\mathsf{A}^{T} \cdot \frac{\partial
1501     V}{\partial \mathsf{A}} \right) - \chi {\bf j} \\
1502     \dot{\chi} & = & \frac{1}{\tau_{T}^2} \left(
1503     \frac{T}{T_{\mathrm{target}}} - 1 \right) \\
1504     \dot{\overleftrightarrow{eta}} & = & \frac{1}{\tau_{B}^2 f k_B
1505     T_{\mathrm{target}}} V \left( \overleftrightarrow{\mathsf{P}} - P_{\mathrm{target}}\mathsf{1} \right) \\
1506     \dot{\mathsf{H}} & = & \overleftrightarrow{\eta} \cdot \mathsf{H}
1507     \label{eq:melchionna2}
1508     \end{eqnarray}
1509    
1510     Here, $\mathsf{1}$ is the unit matrix and $\overleftrightarrow{\mathsf{P}}$
1511     is the pressure tensor. Again, the volume, $\mathcal{V} = \det
1512     \mathsf{H}$.
1513    
1514     The propagation of the equations of motion is nearly identical to the
1515     NPTi integration:
1516    
1517     {\tt moveA:}
1518     \begin{eqnarray}
1519     T(t) & \leftarrow & \left\{{\bf v}(t)\right\}, \left\{{\bf j}(t)\right\} \\
1520     \overleftrightarrow{\mathsf{P}}(t) & \leftarrow & \left\{{\bf r}(t)\right\}, \left\{{\bf v}(t)\right\}, \left\{{\bf f}(t)\right\} \\
1521     {\bf v}\left(t + \delta t / 2\right) & \leftarrow & {\bf
1522     v}(t) + \frac{\delta t}{2} \left( \frac{{\bf f}(t)}{m} -
1523     \left(\chi(t)\mathsf{1} + \overleftrightarrow{\eta}(t) \right) \cdot
1524     {\bf v}(t) \right) \\
1525     {\bf j}\left(t + \delta t / 2 \right) & \leftarrow & {\bf
1526     j}(t) + \frac{\delta t}{2} \left( {\bf \tau}^b(t) - {\bf j}(t)
1527     \chi(t) \right) \\
1528     \mathsf{A}(t + \delta t) & \leftarrow & \mathrm{rot}\left(\delta t *
1529     {\bf j}(t + \delta t / 2) \overleftrightarrow{\mathsf{I}}^{-1} \right) \\
1530     \chi\left(t + \delta t / 2 \right) & \leftarrow & \chi(t) +
1531     \frac{\delta t}{2 \tau_T^2} \left( \frac{T(t)}{T_{\mathrm{target}}} - 1
1532     \right) \\
1533     \overleftrightarrow{\eta}(t + \delta t / 2) & \leftarrow & \overleftrightarrow{\eta}(t) + \frac{\delta t \mathcal{V}(t)}{2 N k_B
1534     T(t) \tau_B^2} \left( \overleftrightarrow{\mathsf{P}}(t) - P_{\mathrm{target}}\mathsf{1} \right) \\
1535     {\bf r}(t + \delta t) & \leftarrow & {\bf r}(t) + \delta t \left\{ {\bf
1536     v}\left(t + \delta t / 2 \right) + \overleftrightarrow{\eta}(t +
1537     \delta t / 2) \cdot \left[ {\bf
1538     r}(t + \delta t) - {\bf R}_0 \right] \right\} \\
1539     \mathsf{H}(t + \delta t) & \leftarrow & \mathsf{H}(t) \cdot e^{-\delta t
1540     \overleftrightarrow{\eta}(t + \delta t / 2)}
1541     \end{eqnarray}
1542     {\sc oopse} uses a power series expansion truncated at second order
1543     for the exponential operation which scales the simulation box.
1544    
1545     The {\tt moveB} portion of the algorithm is largely unchanged from the
1546     NPTi integrator:
1547    
1548     {\tt moveB:}
1549     \begin{eqnarray}
1550     T(t + \delta t) & \leftarrow & \left\{{\bf v}(t + \delta t)\right\},
1551     \left\{{\bf j}(t + \delta t)\right\} \\
1552     \overleftrightarrow{\mathsf{P}}(t + \delta t) & \leftarrow & \left\{{\bf r}(t + \delta t)\right\},
1553     \left\{{\bf v}(t + \delta t)\right\}, \left\{{\bf f}(t + \delta t)\right\} \\
1554     \chi\left(t + \delta t \right) & \leftarrow & \chi\left(t + \delta t /
1555     2 \right) + \frac{\delta t}{2 \tau_T^2} \left( \frac{T(t+\delta
1556     t)}{T_{\mathrm{target}}} - 1 \right) \\
1557     \overleftrightarrow{\eta}(t + \delta t) & \leftarrow & \overleftrightarrow{\eta}(t + \delta t / 2) +
1558     \frac{\delta t \mathcal{V}(t + \delta t)}{2 N k_B T(t + \delta t) \tau_B^2}
1559     \left( \overleftrightarrow{P}(t + \delta t) - P_{\mathrm{target}}\mathsf{1}
1560     \right) \\
1561     {\bf v}\left(t + \delta t \right) & \leftarrow & {\bf
1562     v}\left(t + \delta t / 2 \right) + \frac{\delta t}{2} \left(
1563     \frac{{\bf f}(t + \delta t)}{m} -
1564     (\chi(t + \delta t)\mathsf{1} + \overleftrightarrow{\eta}(t + \delta
1565     t)) \right) \cdot {\bf v}(t + \delta t) \\
1566     {\bf j}\left(t + \delta t \right) & \leftarrow & {\bf
1567     j}\left(t + \delta t / 2 \right) + \frac{\delta t}{2} \left( {\bf
1568     \tau}^b(t + \delta t) - {\bf j}(t + \delta t)
1569     \chi(t + \delta t) \right)
1570     \end{eqnarray}
1571    
1572     The iterative schemes for both {\tt moveA} and {\tt moveB} are
1573     identical to those described for the NPTi integrator.
1574    
1575     The NPTf integrator is known to conserve the following Hamiltonian:
1576     \begin{equation}
1577     H_{\mathrm{NPTf}} = V + K + f k_B T_{\mathrm{target}} \left(
1578     \frac{\tau_{T}^2 \chi^2(t)}{2} + \int_{0}^{t} \chi(t^\prime) dt^\prime
1579     \right) + P_{\mathrm{target}} \mathcal{V}(t) + \frac{f k_B
1580     T_{\mathrm{target}}}{2}
1581     \mathrm{Tr}\left[\overleftrightarrow{\eta}(t)\right]^2 \tau_B^2.
1582     \end{equation}
1583    
1584     This integrator must be used with care, particularly in liquid
1585     simulations. Liquids have very small restoring forces in the
1586     off-diagonal directions, and the simulation box can very quickly form
1587     elongated and sheared geometries which become smaller than the
1588     electrostatic or Lennard-Jones cutoff radii. It finds most use in
1589     simulating crystals or liquid crystals which assume non-orthorhombic
1590     geometries.
1591    
1592     \subsubsection{\label{nptxyz}Constant pressure in 3 axes (NPTxyz)}
1593    
1594     There is one additional extended system integrator which is somewhat
1595     simpler than the NPTf method described above. In this case, the three
1596     axes have independent barostats which each attempt to preserve the
1597     target pressure along the box walls perpendicular to that particular
1598     axis. The lengths of the box axes are allowed to fluctuate
1599     independently, but the angle between the box axes does not change.
1600     The equations of motion are identical to those described above, but
1601     only the {\it diagonal} elements of $\overleftrightarrow{\eta}$ are
1602     computed. The off-diagonal elements are set to zero (even when the
1603     pressure tensor has non-zero off-diagonal elements).
1604    
1605     It should be noted that the NPTxyz integrator is {\it not} known to
1606     preserve any Hamiltonian of interest to the chemical physics
1607     community. The integrator is extremely useful, however, in generating
1608     initial conditions for other integration methods. It {\it is} suitable
1609     for use with liquid simulations, or in cases where there is
1610     orientational anisotropy in the system (i.e. in lipid bilayer
1611     simulations).
1612    
1613 mmeineke 1071 \subsection{\label{oopseSec:rattle}The {\sc rattle} Method for Bond
1614     Constraints}
1615 mmeineke 1044
1616 mmeineke 1071 In order to satisfy the constraints of fixed bond lengths within {\sc
1617     oopse}, we have implemented the {\sc rattle} algorithm of
1618     Andersen.\cite{andersen83} The algorithm is a velocity verlet
1619     formulation of the {\sc shake} method\cite{ryckaert77} of iteratively
1620     solving the Lagrange multipliers of constraint. The system of lagrange
1621     multipliers allows one to reformulate the equations of motion with
1622 mmeineke 1083 explicit constraint forces.\cite{fowles99:lagrange}
1623 mmeineke 1071
1624 mmeineke 1083 Consider a system described by coordinates $q_1$ and $q_2$ subject to an
1625 mmeineke 1071 equation of constraint:
1626     \begin{equation}
1627     \sigma(q_1, q_2,t) = 0
1628     \label{oopseEq:lm1}
1629     \end{equation}
1630     The Lagrange formulation of the equations of motion can be written:
1631     \begin{equation}
1632     \delta\int_{t_1}^{t_2}L\, dt =
1633     \int_{t_1}^{t_2} \sum_i \biggl [ \frac{\partial L}{\partial q_i}
1634     - \frac{d}{dt}\biggl(\frac{\partial L}{\partial \dot{q}_i}
1635     \biggr ) \biggr] \delta q_i \, dt = 0
1636     \label{oopseEq:lm2}
1637     \end{equation}
1638     Here, $\delta q_i$ is not independent for each $q$, as $q_1$ and $q_2$
1639     are linked by $\sigma$. However, $\sigma$ is fixed at any given
1640     instant of time, giving:
1641     \begin{align}
1642     \delta\sigma &= \biggl( \frac{\partial\sigma}{\partial q_1} \delta q_1 %
1643     + \frac{\partial\sigma}{\partial q_2} \delta q_2 \biggr) = 0 \\
1644     %
1645     \frac{\partial\sigma}{\partial q_1} \delta q_1 &= %
1646     - \frac{\partial\sigma}{\partial q_2} \delta q_2 \\
1647     %
1648     \delta q_2 &= - \biggl(\frac{\partial\sigma}{\partial q_1} \bigg / %
1649     \frac{\partial\sigma}{\partial q_2} \biggr) \delta q_1
1650     \end{align}
1651     Substituted back into Eq.~\ref{oopseEq:lm2},
1652     \begin{equation}
1653     \int_{t_1}^{t_2}\biggl [ \biggl(\frac{\partial L}{\partial q_1}
1654     - \frac{d}{dt}\,\frac{\partial L}{\partial \dot{q}_1}
1655     \biggr)
1656     - \biggl( \frac{\partial L}{\partial q_1}
1657     - \frac{d}{dt}\,\frac{\partial L}{\partial \dot{q}_1}
1658     \biggr) \biggl(\frac{\partial\sigma}{\partial q_1} \bigg / %
1659     \frac{\partial\sigma}{\partial q_2} \biggr)\biggr] \delta q_1 \, dt = 0
1660     \label{oopseEq:lm3}
1661     \end{equation}
1662     Leading to,
1663     \begin{equation}
1664     \frac{\biggl(\frac{\partial L}{\partial q_1}
1665     - \frac{d}{dt}\,\frac{\partial L}{\partial \dot{q}_1}
1666     \biggr)}{\frac{\partial\sigma}{\partial q_1}} =
1667     \frac{\biggl(\frac{\partial L}{\partial q_2}
1668     - \frac{d}{dt}\,\frac{\partial L}{\partial \dot{q}_2}
1669     \biggr)}{\frac{\partial\sigma}{\partial q_2}}
1670     \label{oopseEq:lm4}
1671     \end{equation}
1672     This relation can only be statisfied, if both are equal to a single
1673     function $-\lambda(t)$,
1674     \begin{align}
1675     \frac{\biggl(\frac{\partial L}{\partial q_1}
1676     - \frac{d}{dt}\,\frac{\partial L}{\partial \dot{q}_1}
1677     \biggr)}{\frac{\partial\sigma}{\partial q_1}} &= -\lambda(t) \\
1678     %
1679     \frac{\partial L}{\partial q_1}
1680     - \frac{d}{dt}\,\frac{\partial L}{\partial \dot{q}_1} &=
1681     -\lambda(t)\,\frac{\partial\sigma}{\partial q_1} \\
1682     %
1683     \frac{\partial L}{\partial q_1}
1684     - \frac{d}{dt}\,\frac{\partial L}{\partial \dot{q}_1}
1685     + \mathcal{G}_i &= 0
1686     \end{align}
1687     Where $\mathcal{G}_i$, the force of constraint on $i$, is:
1688     \begin{equation}
1689     \mathcal{G}_i = \lambda(t)\,\frac{\partial\sigma}{\partial q_1}
1690     \label{oopseEq:lm5}
1691     \end{equation}
1692    
1693     In a simulation, this would involve the solution of a set of $(m + n)$
1694     number of equations. Where $m$ is the number of constraints, and $n$
1695     is the number of constrained coordinates. In practice, this is not
1696 mmeineke 1083 done, as the matrix inversion necessary to solve the system of
1697 mmeineke 1071 equations would be very time consuming to solve. Additionally, the
1698     numerical error in the solution of the set of $\lambda$'s would be
1699     compounded by the error inherent in propagating by the Velocity Verlet
1700 mmeineke 1083 algorithm ($\Delta t^4$). The Verlet propagation error is negligible
1701     in an unconstrained system, as one is interested in the statistics of
1702 mmeineke 1071 the run, and not that the run be numerically exact to the ``true''
1703     integration. This relates back to the ergodic hypothesis that a time
1704 mmeineke 1083 integral of a valid trajectory will still give the correct ensemble
1705 mmeineke 1071 average. However, in the case of constraints, if the equations of
1706     motion leave the ``true'' trajectory, they are departing from the
1707     constrained surface. The method that is used, is to iteratively solve
1708     for $\lambda(t)$ at each time step.
1709    
1710     In {\sc rattle} the equations of motion are modified subject to the
1711     following two constraints:
1712     \begin{align}
1713     \sigma_{ij}[\mathbf{r}(t)] \equiv
1714     [ \mathbf{r}_i(t) - \mathbf{r}_j(t)]^2 - d_{ij}^2 &= 0 %
1715     \label{oopseEq:c1} \\
1716     %
1717     [\mathbf{\dot{r}}_i(t) - \mathbf{\dot{r}}_j(t)] \cdot
1718     [\mathbf{r}_i(t) - \mathbf{r}_j(t)] &= 0 \label{oopseEq:c2}
1719     \end{align}
1720     Eq.~\ref{oopseEq:c1} is the set of bond constraints, where $d_{ij}$ is
1721     the constrained distance between atom $i$ and
1722     $j$. Eq.~\ref{oopseEq:c2} constrains the velocities of $i$ and $j$ to
1723 mmeineke 1083 be perpendicular to the bond vector, so that the bond can neither grow
1724 mmeineke 1071 nor shrink. The constrained dynamics equations become:
1725     \begin{equation}
1726     m_i \mathbf{\ddot{r}}_i = \mathbf{F}_i + \mathbf{\mathcal{G}}_i
1727     \label{oopseEq:r1}
1728     \end{equation}
1729 mmeineke 1083 Where,$\mathbf{\mathcal{G}}_i$ are the forces of constraint on $i$,
1730     and are defined:
1731 mmeineke 1071 \begin{equation}
1732     \mathbf{\mathcal{G}}_i = - \sum_j \lambda_{ij}(t)\,\nabla \sigma_{ij}
1733     \label{oopseEq:r2}
1734     \end{equation}
1735    
1736     In Velocity Verlet, if $\Delta t = h$, the propagation can be written:
1737     \begin{align}
1738     \mathbf{r}_i(t+h) &=
1739     \mathbf{r}_i(t) + h\mathbf{\dot{r}}(t) +
1740     \frac{h^2}{2m_i}\,\Bigl[ \mathbf{F}_i(t) +
1741     \mathbf{\mathcal{G}}_{Ri}(t) \Bigr] \label{oopseEq:vv1} \\
1742     %
1743     \mathbf{\dot{r}}_i(t+h) &=
1744     \mathbf{\dot{r}}_i(t) + \frac{h}{2m_i}
1745     \Bigl[ \mathbf{F}_i(t) + \mathbf{\mathcal{G}}_{Ri}(t) +
1746     \mathbf{F}_i(t+h) + \mathbf{\mathcal{G}}_{Vi}(t+h) \Bigr] %
1747     \label{oopseEq:vv2}
1748     \end{align}
1749 mmeineke 1083 Where:
1750     \begin{align}
1751     \mathbf{\mathcal{G}}_{Ri}(t) &=
1752     -2 \sum_j \lambda_{Rij}(t) \mathbf{r}_{ij}(t) \\
1753     %
1754     \mathbf{\mathcal{G}}_{Vi}(t+h) &=
1755     -2 \sum_j \lambda_{Vij}(t+h) \mathbf{r}(t+h)
1756     \end{align}
1757     Next, define:
1758     \begin{align}
1759     g_{ij} &= h \lambda_{Rij}(t) \\
1760     k_{ij} &= h \lambda_{Vij}(t+h) \\
1761     \mathbf{q}_i &= \mathbf{\dot{r}}_i(t) + \frac{h}{2m_i} \mathbf{F}_i(t)
1762     - \frac{1}{m_i}\sum_j g_{ij}\mathbf{r}_{ij}(t)
1763     \end{align}
1764     Using these definitions, Eq.~\ref{oopseEq:vv1} and \ref{oopseEq:vv2}
1765     can be rewritten as,
1766     \begin{align}
1767     \mathbf{r}_i(t+h) &= \mathbf{r}_i(t) + h \mathbf{q}_i \\
1768     %
1769     \mathbf{\dot{r}}(t+h) &= \mathbf{q}_i + \frac{h}{2m_i}\mathbf{F}_i(t+h)
1770     -\frac{1}{m_i}\sum_j k_{ij} \mathbf{r}_{ij}(t+h)
1771     \end{align}
1772 mmeineke 1071
1773 mmeineke 1083 To integrate the equations of motion, the {\sc rattle} algorithm first
1774     solves for $\mathbf{r}(t+h)$. Let,
1775     \begin{equation}
1776     \mathbf{q}_i = \mathbf{\dot{r}}(t) + \frac{h}{2m_i}\mathbf{F}_i(t)
1777     \end{equation}
1778     Here $\mathbf{q}_i$ corresponds to an initial unconstrained move. Next
1779     pick a constraint $j$, and let,
1780     \begin{equation}
1781     \mathbf{s} = \mathbf{r}_i(t) + h\mathbf{q}_i(t)
1782     - \mathbf{r}_j(t) + h\mathbf{q}_j(t)
1783     \label{oopseEq:ra1}
1784     \end{equation}
1785     If
1786     \begin{equation}
1787     \Big| |\mathbf{s}|^2 - d_{ij}^2 \Big| > \text{tolerance},
1788     \end{equation}
1789     then the constraint is unsatisfied, and corrections are made to the
1790     positions. First we define a test corrected configuration as,
1791     \begin{align}
1792     \mathbf{r}_i^T(t+h) = \mathbf{r}_i(t) + h\biggl[\mathbf{q}_i -
1793     g_{ij}\,\frac{\mathbf{r}_{ij}(t)}{m_i} \biggr] \\
1794     %
1795     \mathbf{r}_j^T(t+h) = \mathbf{r}_j(t) + h\biggl[\mathbf{q}_j +
1796     g_{ij}\,\frac{\mathbf{r}_{ij}(t)}{m_j} \biggr]
1797     \end{align}
1798     And we chose $g_{ij}$ such that, $|\mathbf{r}_i^T - \mathbf{r}_j^T|^2
1799     = d_{ij}^2$. Solving the quadratic for $g_{ij}$ we obtain the
1800     approximation,
1801     \begin{equation}
1802     g_{ij} = \frac{(s^2 - d^2)}{2h[\mathbf{s}\cdot\mathbf{r}_{ij}(t)]
1803     (\frac{1}{m_i} + \frac{1}{m_j})}
1804     \end{equation}
1805     Although not an exact solution for $g_{ij}$, as this is an iterative
1806     scheme overall, the eventual solution will converge. With a trial
1807     $g_{ij}$, the new $\mathbf{q}$'s become,
1808     \begin{align}
1809     \mathbf{q}_i &= \mathbf{q}^{\text{old}}_i - g_{ij}\,
1810     \frac{\mathbf{r}_{ij}(t)}{m_i} \\
1811     %
1812     \mathbf{q}_j &= \mathbf{q}^{\text{old}}_j + g_{ij}\,
1813     \frac{\mathbf{r}_{ij}(t)}{m_j}
1814     \end{align}
1815     The whole algorithm is then repeated from Eq.~\ref{oopseEq:ra1} until
1816     all constraints are satisfied.
1817 mmeineke 1071
1818 mmeineke 1083 The second step of {\sc rattle}, is to then update the velocities. The
1819     step starts with,
1820     \begin{equation}
1821     \mathbf{\dot{r}}_i(t+h) = \mathbf{q}_i + \frac{h}{2m_i}\mathbf{F}_i(t+h)
1822     \end{equation}
1823     Next we pick a constraint $j$, and calculate the dot product $\ell$.
1824     \begin{equation}
1825     \ell = \mathbf{r}_{ij}(t+h) \cdot \mathbf{\dot{r}}_{ij}(t+h)
1826     \label{oopseEq:rv1}
1827     \end{equation}
1828     Here if constraint Eq.~\ref{oopseEq:c2} holds, $\ell$ should be
1829     zero. Therefore if $\ell$ is greater than some tolerance, then
1830     corrections are made to the $i$ and $j$ velocities.
1831     \begin{align}
1832     \mathbf{\dot{r}}_i^T &= \mathbf{\dot{r}}_i(t+h) - k_{ij}
1833     \frac{\mathbf{\dot{r}}_{ij}(t+h)}{m_i} \\
1834     %
1835     \mathbf{\dot{r}}_j^T &= \mathbf{\dot{r}}_j(t+h) + k_{ij}
1836     \frac{\mathbf{\dot{r}}_{ij}(t+h)}{m_j}
1837     \end{align}
1838     Like in the previous step, we select a value for $k_{ij}$ such that
1839     $\ell$ is zero.
1840     \begin{equation}
1841     k_{ij} = \frac{\ell}{d^2_{ij}(\frac{1}{m_i} + \frac{1}{m_j})}
1842     \end{equation}
1843     The test velocities, $\mathbf{\dot{r}}^T_i$ and
1844     $\mathbf{\dot{r}}^T_j$, then replace their respective velocities, and
1845     the algorithm is iterated from Eq.~\ref{oopseEq:rv1} until all
1846     constraints are satisfied.
1847 mmeineke 1071
1848 mmeineke 1083
1849 mmeineke 1054 \subsection{\label{oopseSec:zcons}Z-Constraint Method}
1850 mmeineke 1044
1851 mmeineke 1083 Based on the fluctuation-dissipation theorem, a force auto-correlation
1852     method was developed by Roux and Karplus to investigate the dynamics
1853     of ions inside ion channels.\cite{Roux91} The time-dependent friction
1854     coefficient can be calculated from the deviation of the instantaneous
1855     force from its mean force.
1856 mmeineke 1044 \begin{equation}
1857     \xi(z,t)=\langle\delta F(z,t)\delta F(z,0)\rangle/k_{B}T
1858     \end{equation}
1859     where%
1860     \begin{equation}
1861     \delta F(z,t)=F(z,t)-\langle F(z,t)\rangle
1862     \end{equation}
1863    
1864    
1865 mmeineke 1083 If the time-dependent friction decays rapidly, the static friction
1866     coefficient can be approximated by
1867 mmeineke 1044 \begin{equation}
1868 mmeineke 1087 \xi_{\text{static}}(z)=\int_{0}^{\infty}\langle\delta F(z,t)\delta F(z,0)\rangle dt
1869 mmeineke 1044 \end{equation}
1870 mmeineke 1087 Allowing diffusion constant to then be calculated through the
1871     Einstein relation:\cite{Marrink94}
1872 mmeineke 1044 \begin{equation}
1873 mmeineke 1087 D(z)=\frac{k_{B}T}{\xi_{\text{static}}(z)}=\frac{(k_{B}T)^{2}}{\int_{0}^{\infty
1874 mmeineke 1044 }\langle\delta F(z,t)\delta F(z,0)\rangle dt}%
1875     \end{equation}
1876    
1877 mmeineke 1083 The Z-Constraint method, which fixes the z coordinates of the
1878     molecules with respect to the center of the mass of the system, has
1879     been a method suggested to obtain the forces required for the force
1880     auto-correlation calculation.\cite{Marrink94} However, simply resetting the
1881     coordinate will move the center of the mass of the whole system. To
1882     avoid this problem, a new method was used in {\sc oopse}. Instead of
1883 mmeineke 1087 resetting the coordinate, we reset the forces of z-constrained
1884 mmeineke 1083 molecules as well as subtract the total constraint forces from the
1885 mmeineke 1087 rest of the system after the force calculation at each time step.
1886 mmeineke 1044
1887 mmeineke 1087 After the force calculation, define $G_\alpha$ as
1888     \begin{equation}
1889     G_{\alpha} = \sum_i F_{\alpha i}
1890     \label{oopseEq:zc1}
1891     \end{equation}
1892     Where $F_{\alpha i}$ is the force in the z direction of atom $i$ in
1893     z-constrained molecule $\alpha$. The forces of the z constrained
1894     molecule are then set to:
1895     \begin{equation}
1896     F_{\alpha i} = F_{\alpha i} -
1897     \frac{m_{\alpha i} G_{\alpha}}{\sum_i m_{\alpha i}}
1898     \end{equation}
1899     Here, $m_{\alpha i}$ is the mass of atom $i$ in the z-constrained
1900     molecule. Having rescaled the forces, the velocities must also be
1901     rescaled to subtract out any center of mass velocity in the z
1902     direction.
1903     \begin{equation}
1904     v_{\alpha i} = v_{\alpha i} -
1905     \frac{\sum_i m_{\alpha i} v_{\alpha i}}{\sum_i m_{\alpha i}}
1906     \end{equation}
1907     Where $v_{\alpha i}$ is the velocity of atom $i$ in the z direction.
1908     Lastly, all of the accumulated z constrained forces must be subtracted
1909     from the system to keep the system center of mass from drifting.
1910     \begin{equation}
1911     F_{\beta i} = F_{\beta i} - \frac{m_{\beta i} \sum_{\alpha} G_{\alpha}}
1912     {\sum_{\beta}\sum_i m_{\beta i}}
1913     \end{equation}
1914     Where $\beta$ are all of the unconstrained molecules in the system.
1915    
1916 mmeineke 1083 At the very beginning of the simulation, the molecules may not be at their
1917     constrained positions. To move a z-constrained molecule to its specified
1918     position, a simple harmonic potential is used
1919 mmeineke 1044 \begin{equation}
1920 mmeineke 1083 U(t)=\frac{1}{2}k_{\text{Harmonic}}(z(t)-z_{\text{cons}})^{2}%
1921 mmeineke 1044 \end{equation}
1922 mmeineke 1083 where $k_{\text{Harmonic}}$ is the harmonic force constant, $z(t)$ is the
1923     current $z$ coordinate of the center of mass of the constrained molecule, and
1924     $z_{\text{cons}}$ is the constrained position. The harmonic force operating
1925     on the z-constrained molecule at time $t$ can be calculated by
1926 mmeineke 1044 \begin{equation}
1927 mmeineke 1083 F_{z_{\text{Harmonic}}}(t)=-\frac{\partial U(t)}{\partial z(t)}=
1928     -k_{\text{Harmonic}}(z(t)-z_{\text{cons}})
1929 mmeineke 1044 \end{equation}
1930    
1931 mmeineke 1051 \section{\label{oopseSec:props}Trajectory Analysis}
1932 mmeineke 1044
1933 mmeineke 1051 \subsection{\label{oopseSec:staticProps}Static Property Analysis}
1934 mmeineke 1044
1935     The static properties of the trajectories are analyzed with the
1936 mmeineke 1083 program \texttt{staticProps}. The code is capable of calculating a
1937     number of pair correlations between species A and B. Some of which
1938     only apply to directional entities. The summary of pair correlations
1939     can be found in Table~\ref{oopseTb:gofrs}
1940 mmeineke 1044
1941 mmeineke 1083 \begin{table}
1942     \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.}
1943     \label{oopseTb:gofrs}
1944     \begin{center}
1945     \begin{tabular}{|l|c|c|}
1946     \hline
1947     Name & Equation & Directional Atom \\ \hline
1948     $g_{\text{AB}}(r)$ & Eq.~\ref{eq:gofr} & neither \\ \hline
1949     $g_{\text{AB}}(r, \cos \theta)$ & Eq.~\ref{eq:gofrCosTheta} & A \\ \hline
1950     $g_{\text{AB}}(r, \cos \omega)$ & Eq.~\ref{eq:gofrCosOmega} & both \\ \hline
1951     $g_{\text{AB}}(x, y, z)$ & Eq.~\ref{eq:gofrXYZ} & neither \\ \hline
1952     $\langle \cos \omega \rangle_{\text{AB}}(r)$ & Eq.~\ref{eq:cosOmegaOfR} &%
1953     both \\ \hline
1954     \end{tabular}
1955     \end{center}
1956     \end{table}
1957    
1958 mmeineke 1044 The first pair correlation, $g_{\text{AB}}(r)$, is defined as follows:
1959     \begin{equation}
1960     g_{\text{AB}}(r) = \frac{V}{N_{\text{A}}N_{\text{B}}}\langle %%
1961     \sum_{i \in \text{A}} \sum_{j \in \text{B}} %%
1962     \delta( r - |\mathbf{r}_{ij}|) \rangle \label{eq:gofr}
1963     \end{equation}
1964     Where $\mathbf{r}_{ij}$ is the vector
1965     \begin{equation*}
1966     \mathbf{r}_{ij} = \mathbf{r}_j - \mathbf{r}_i \notag
1967     \end{equation*}
1968     and $\frac{V}{N_{\text{A}}N_{\text{B}}}$ normalizes the average over
1969     the expected pair density at a given $r$.
1970    
1971     The next two pair correlations, $g_{\text{AB}}(r, \cos \theta)$ and
1972     $g_{\text{AB}}(r, \cos \omega)$, are similar in that they are both two
1973     dimensional histograms. Both use $r$ for the primary axis then a
1974     $\cos$ for the secondary axis ($\cos \theta$ for
1975     Eq.~\ref{eq:gofrCosTheta} and $\cos \omega$ for
1976     Eq.~\ref{eq:gofrCosOmega}). This allows for the investigator to
1977     correlate alignment on directional entities. $g_{\text{AB}}(r, \cos
1978     \theta)$ is defined as follows:
1979     \begin{equation}
1980     g_{\text{AB}}(r, \cos \theta) = \frac{V}{N_{\text{A}}N_{\text{B}}}\langle
1981     \sum_{i \in \text{A}} \sum_{j \in \text{B}}
1982     \delta( \cos \theta - \cos \theta_{ij})
1983     \delta( r - |\mathbf{r}_{ij}|) \rangle
1984     \label{eq:gofrCosTheta}
1985     \end{equation}
1986     Where
1987     \begin{equation*}
1988     \cos \theta_{ij} = \mathbf{\hat{i}} \cdot \mathbf{\hat{r}}_{ij}
1989     \end{equation*}
1990     Here $\mathbf{\hat{i}}$ is the unit directional vector of species $i$
1991     and $\mathbf{\hat{r}}_{ij}$ is the unit vector associated with vector
1992     $\mathbf{r}_{ij}$.
1993    
1994     The second two dimensional histogram is of the form:
1995     \begin{equation}
1996     g_{\text{AB}}(r, \cos \omega) =
1997     \frac{V}{N_{\text{A}}N_{\text{B}}}\langle
1998     \sum_{i \in \text{A}} \sum_{j \in \text{B}}
1999     \delta( \cos \omega - \cos \omega_{ij})
2000     \delta( r - |\mathbf{r}_{ij}|) \rangle \label{eq:gofrCosOmega}
2001     \end{equation}
2002     Here
2003     \begin{equation*}
2004     \cos \omega_{ij} = \mathbf{\hat{i}} \cdot \mathbf{\hat{j}}
2005     \end{equation*}
2006     Again, $\mathbf{\hat{i}}$ and $\mathbf{\hat{j}}$ are the unit
2007     directional vectors of species $i$ and $j$.
2008    
2009     The static analysis code is also cable of calculating a three
2010     dimensional pair correlation of the form:
2011     \begin{equation}\label{eq:gofrXYZ}
2012     g_{\text{AB}}(x, y, z) =
2013     \frac{V}{N_{\text{A}}N_{\text{B}}}\langle
2014     \sum_{i \in \text{A}} \sum_{j \in \text{B}}
2015     \delta( x - x_{ij})
2016     \delta( y - y_{ij})
2017     \delta( z - z_{ij}) \rangle
2018     \end{equation}
2019     Where $x_{ij}$, $y_{ij}$, and $z_{ij}$ are the $x$, $y$, and $z$
2020     components respectively of vector $\mathbf{r}_{ij}$.
2021    
2022     The final pair correlation is similar to
2023     Eq.~\ref{eq:gofrCosOmega}. $\langle \cos \omega
2024     \rangle_{\text{AB}}(r)$ is calculated in the following way:
2025     \begin{equation}\label{eq:cosOmegaOfR}
2026     \langle \cos \omega \rangle_{\text{AB}}(r) =
2027     \langle \sum_{i \in \text{A}} \sum_{j \in \text{B}}
2028     (\cos \omega_{ij}) \delta( r - |\mathbf{r}_{ij}|) \rangle
2029     \end{equation}
2030     Here $\cos \omega_{ij}$ is defined in the same way as in
2031     Eq.~\ref{eq:gofrCosOmega}. This equation is a single dimensional pair
2032     correlation that gives the average correlation of two directional
2033     entities as a function of their distance from each other.
2034    
2035     \subsection{\label{dynamicProps}Dynamic Property Analysis}
2036    
2037     The dynamic properties of a trajectory are calculated with the program
2038 mmeineke 1083 \texttt{dynamicProps}. The program calculates the following properties:
2039 mmeineke 1044 \begin{gather}
2040     \langle | \mathbf{r}(t) - \mathbf{r}(0) |^2 \rangle \label{eq:rms}\\
2041     \langle \mathbf{v}(t) \cdot \mathbf{v}(0) \rangle \label{eq:velCorr} \\
2042     \langle \mathbf{j}(t) \cdot \mathbf{j}(0) \rangle \label{eq:angularVelCorr}
2043     \end{gather}
2044    
2045 mmeineke 1083 Eq.~\ref{eq:rms} is the root mean square displacement function. Which
2046     allows one to observe the average displacement of an atom as a
2047     function of time. The quantity is useful when calculating diffusion
2048     coefficients because of the Einstein Relation, which is valid at long
2049     times.\cite{allen87:csl}
2050     \begin{equation}
2051     2tD = \langle | \mathbf{r}(t) - \mathbf{r}(0) |^2 \rangle
2052     \label{oopseEq:einstein}
2053     \end{equation}
2054    
2055     Eq.~\ref{eq:velCorr} and \ref{eq:angularVelCorr} are the translational
2056 mmeineke 1044 velocity and angular velocity correlation functions respectively. The
2057 mmeineke 1083 latter is only applicable to directional species in the
2058     simulation. The velocity autocorrelation functions are useful when
2059     determining vibrational information about the system of interest.
2060 mmeineke 1044
2061 mmeineke 1051 \section{\label{oopseSec:design}Program Design}
2062 mmeineke 1044
2063 mmeineke 1054 \subsection{\label{sec:architecture} {\sc oopse} Architecture}
2064 mmeineke 1044
2065 mmeineke 1054 The core of OOPSE is divided into two main object libraries:
2066     \texttt{libBASS} and \texttt{libmdtools}. \texttt{libBASS} is the
2067     library developed around the parsing engine and \texttt{libmdtools}
2068     is the software library developed around the simulation engine. These
2069     two libraries are designed to encompass all the basic functions and
2070     tools that {\sc oopse} provides. Utility programs, such as the
2071     property analyzers, need only link against the software libraries to
2072     gain access to parsing, force evaluation, and input / output
2073     routines.
2074 mmeineke 1044
2075 mmeineke 1054 Contained in \texttt{libBASS} are all the routines associated with
2076     reading and parsing the \texttt{.bass} input files. Given a
2077     \texttt{.bass} file, \texttt{libBASS} will open it and any associated
2078     \texttt{.mdl} files; then create structures in memory that are
2079     templates of all the molecules specified in the input files. In
2080     addition, any simulation parameters set in the \texttt{.bass} file
2081     will be placed in a structure for later query by the controlling
2082     program.
2083 mmeineke 1044
2084 mmeineke 1054 Located in \texttt{libmdtools} are all other routines necessary to a
2085     Molecular Dynamics simulation. The library uses the main data
2086     structures returned by \texttt{libBASS} to initialize the various
2087     parts of the simulation: the atom structures and positions, the force
2088     field, the integrator, \emph{et cetera}. After initialization, the
2089     library can be used to perform a variety of tasks: integrate a
2090     Molecular Dynamics trajectory, query phase space information from a
2091     specific frame of a completed trajectory, or even recalculate force or
2092     energetic information about specific frames from a completed
2093     trajectory.
2094 mmeineke 1044
2095 mmeineke 1054 With these core libraries in place, several programs have been
2096     developed to utilize the routines provided by \texttt{libBASS} and
2097     \texttt{libmdtools}. The main program of the package is \texttt{oopse}
2098     and the corresponding parallel version \texttt{oopse\_MPI}. These two
2099 mmeineke 1083 programs will take the \texttt{.bass} file, and create (and integrate)
2100 mmeineke 1054 the simulation specified in the script. The two analysis programs
2101     \texttt{staticProps} and \texttt{dynamicProps} utilize the core
2102     libraries to initialize and read in trajectories from previously
2103     completed simulations, in addition to the ability to use functionality
2104     from \texttt{libmdtools} to recalculate forces and energies at key
2105     frames in the trajectories. Lastly, the family of system building
2106     programs (Sec.~\ref{oopseSec:initCoords}) also use the libraries to
2107     store and output the system configurations they create.
2108    
2109     \subsection{\label{oopseSec:parallelization} Parallelization of {\sc oopse}}
2110    
2111 mmeineke 1083 Although processor power is continually growing roughly following
2112     Moore's Law, it is still unreasonable to simulate systems of more then
2113     a 1000 atoms on a single processor. To facilitate study of larger
2114     system sizes or smaller systems on long time scales in a reasonable
2115     period of time, parallel methods were developed allowing multiple
2116     CPU's to share the simulation workload. Three general categories of
2117     parallel decomposition methods have been developed including atomic,
2118     spatial and force decomposition methods.
2119 mmeineke 1054
2120 mmeineke 1083 Algorithmically simplest of the three methods is atomic decomposition
2121 mmeineke 1044 where N particles in a simulation are split among P processors for the
2122     duration of the simulation. Computational cost scales as an optimal
2123     $O(N/P)$ for atomic decomposition. Unfortunately all processors must
2124 mmeineke 1083 communicate positions and forces with all other processors at every
2125     force evaluation, leading communication costs to scale as an
2126     unfavorable $O(N)$, \emph{independent of the number of processors}. This
2127     communication bottleneck led to the development of spatial and force
2128     decomposition methods in which communication among processors scales
2129     much more favorably. Spatial or domain decomposition divides the
2130     physical spatial domain into 3D boxes in which each processor is
2131     responsible for calculation of forces and positions of particles
2132     located in its box. Particles are reassigned to different processors
2133     as they move through simulation space. To calculate forces on a given
2134     particle, a processor must know the positions of particles within some
2135     cutoff radius located on nearby processors instead of the positions of
2136     particles on all processors. Both communication between processors and
2137     computation scale as $O(N/P)$ in the spatial method. However, spatial
2138 mmeineke 1044 decomposition adds algorithmic complexity to the simulation code and
2139     is not very efficient for small N since the overall communication
2140     scales as the surface to volume ratio $(N/P)^{2/3}$ in three
2141     dimensions.
2142    
2143 mmeineke 1083 The parallelization method used in {\sc oopse} is the force
2144     decomposition method. Force decomposition assigns particles to
2145     processors based on a block decomposition of the force
2146     matrix. Processors are split into an optimally square grid forming row
2147     and column processor groups. Forces are calculated on particles in a
2148     given row by particles located in that processors column
2149     assignment. Force decomposition is less complex to implement than the
2150     spatial method but still scales computationally as $O(N/P)$ and scales
2151     as $O(N/\sqrt{P})$ in communication cost. Plimpton has also found that
2152     force decompositions scale more favorably than spatial decompositions
2153     for systems up to 10,000 atoms and favorably compete with spatial
2154     methods up to 100,000 atoms.\cite{plimpton95}
2155 mmeineke 1044
2156 mmeineke 1054 \subsection{\label{oopseSec:memAlloc}Memory Issues in Trajectory Analysis}
2157 mmeineke 1044
2158 mmeineke 1054 For large simulations, the trajectory files can sometimes reach sizes
2159     in excess of several gigabytes. In order to effectively analyze that
2160 mmeineke 1083 amount of data, two memory management schemes have been devised for
2161 mmeineke 1054 \texttt{staticProps} and for \texttt{dynamicProps}. The first scheme,
2162     developed for \texttt{staticProps}, is the simplest. As each frame's
2163     statistics are calculated independent of each other, memory is
2164     allocated for each frame, then freed once correlation calculations are
2165     complete for the snapshot. To prevent multiple passes through a
2166     potentially large file, \texttt{staticProps} is capable of calculating
2167     all requested correlations per frame with only a single pair loop in
2168 mmeineke 1083 each frame and a single read of the file.
2169 mmeineke 1044
2170 mmeineke 1054 The second, more advanced memory scheme, is used by
2171     \texttt{dynamicProps}. Here, the program must have multiple frames in
2172     memory to calculate time dependent correlations. In order to prevent a
2173     situation where the program runs out of memory due to large
2174     trajectories, the user is able to specify that the trajectory be read
2175     in blocks. The number of frames in each block is specified by the
2176     user, and upon reading a block of the trajectory,
2177     \texttt{dynamicProps} will calculate all of the time correlation frame
2178 mmeineke 1083 pairs within the block. After in-block correlations are complete, a
2179 mmeineke 1054 second block of the trajectory is read, and the cross correlations are
2180     calculated between the two blocks. this second block is then freed and
2181     then incremented and the process repeated until the end of the
2182     trajectory. Once the end is reached, the first block is freed then
2183     incremented, and the again the internal time correlations are
2184     calculated. The algorithm with the second block is then repeated with
2185     the new origin block, until all frame pairs have been correlated in
2186     time. This process is illustrated in
2187     Fig.~\ref{oopseFig:dynamicPropsMemory}.
2188 mmeineke 1044
2189 mmeineke 1054 \begin{figure}
2190     \centering
2191     \includegraphics[width=\linewidth]{dynamicPropsMem.eps}
2192     \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.}
2193     \label{oopseFig:dynamicPropsMemory}
2194     \end{figure}
2195 mmeineke 1044
2196 mmeineke 1054 \section{\label{oopseSec:conclusion}Conclusion}
2197 mmeineke 1044
2198 mmeineke 1054 We have presented the design and implementation of our open source
2199 mmeineke 1083 simulation package {\sc oopse}. The package offers novel capabilities
2200     to the field of Molecular Dynamics simulation packages in the form of
2201     dipolar force fields, and symplectic integration of rigid body
2202     dynamics. It is capable of scaling across multiple processors through
2203     the use of force based decomposition using MPI. It also implements
2204     several advanced integrators allowing the end user control over
2205     temperature and pressure. In addition, it is capable of integrating
2206     constrained dynamics through both the {\sc rattle} algorithm and the
2207     z-constraint method.
2208 mmeineke 1044
2209 mmeineke 1054 These features are all brought together in a single open-source
2210 mmeineke 1083 program. Allowing researchers to not only benefit from
2211 mmeineke 1054 {\sc oopse}, but also contribute to {\sc oopse}'s development as
2212     well.Documentation and source code for {\sc oopse} can be downloaded
2213     from \texttt{http://www.openscience.org/oopse/}.
2214 mmeineke 1044