ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/openmdDocs/openmdDoc.tex
Revision: 4104
Committed: Mon Apr 28 21:06:04 2014 UTC (10 years, 4 months ago) by gezelter
Content type: application/x-tex
File size: 219581 byte(s)
Log Message:
latest edits

File Contents

# Content
1 \documentclass[]{book}
2 \usepackage{amssymb}
3 \usepackage{amsmath}
4 \usepackage{times}
5 \usepackage{listings}
6 \usepackage{graphicx}
7 \usepackage{setspace}
8 \usepackage{tabularx}
9 \usepackage{longtable}
10 \pagestyle{plain}
11 \pagenumbering{arabic}
12 \usepackage{floatrow}
13 \oddsidemargin 0.0cm
14 \evensidemargin 0.0cm
15 \topmargin -21pt
16 \headsep 10pt
17 \textheight 9.0in
18 \textwidth 6.5in
19 \brokenpenalty=10000
20 \renewcommand{\baselinestretch}{1.2}
21 \usepackage[square, comma, sort&compress]{natbib}
22 \bibpunct{[}{]}{,}{n}{}{;}
23
24 \DeclareFloatFont{tiny}{\scriptsize}% "scriptsize" is defined by floatrow, "tiny" not
25 \floatsetup[table]{font=tiny}
26
27
28 %\renewcommand\citemid{\ } % no comma in optional reference note
29 \lstset{language=C,frame=TB,basicstyle=\footnotesize\ttfamily, %
30 xleftmargin=0.25in, xrightmargin=0.25in,captionpos=b, %
31 abovecaptionskip=0.5cm, belowcaptionskip=0.5cm, escapeinside={~}{~}}
32 \renewcommand{\lstlistingname}{Scheme}
33
34 \begin{document}
35
36 \newcolumntype{A}{p{1.5in}}
37 \newcolumntype{B}{p{0.75in}}
38 \newcolumntype{C}{p{1.5in}}
39 \newcolumntype{D}{p{2in}}
40
41 \newcolumntype{E}{p{0.5in}}
42 \newcolumntype{F}{p{2.25in}}
43 \newcolumntype{G}{p{3in}}
44
45 \newcolumntype{H}{p{0.75in}}
46 \newcolumntype{I}{p{5in}}
47
48 \newcolumntype{J}{p{1.5in}}
49 \newcolumntype{K}{p{1.2in}}
50 \newcolumntype{L}{p{1.5in}}
51 \newcolumntype{M}{p{1.55in}}
52
53
54 \title{{\sc OpenMD-2.2}: Molecular Dynamics in the Open}
55
56 \author{Joseph Michalka, James Marr, Kelsey Stocker, Madan Lamichhane,
57 Patrick Louden, \\
58 Teng Lin, Charles F. Vardeman II, Christopher J. Fennell, Shenyu
59 Kuang, Xiuquan Sun, \\
60 Chunlei Li, Kyle Daily, Yang Zheng, Matthew A. Meineke, and \\
61 J. Daniel Gezelter \\
62 Department of Chemistry and Biochemistry\\
63 University of Notre Dame\\
64 Notre Dame, Indiana 46556}
65
66 \maketitle
67
68 \section*{Preface}
69 {\sc OpenMD} is an open source molecular dynamics engine which is capable of
70 efficiently simulating liquids, proteins, nanoparticles, interfaces,
71 and other complex systems using atom types with orientational degrees
72 of freedom (e.g. ``sticky'' atoms, point dipoles, and coarse-grained
73 assemblies). Proteins, zeolites, lipids, transition metals (bulk, flat
74 interfaces, and nanoparticles) have all been simulated using force
75 fields included with the code. {\sc OpenMD} works on parallel computers
76 using the Message Passing Interface (MPI), and comes with a number of
77 analysis and utility programs that are easy to use and modify. An
78 OpenMD simulation is specified using a very simple meta-data language
79 that is easy to learn.
80
81 \tableofcontents
82 \listoffigures
83 \listoftables
84
85 \mainmatter
86
87 \chapter{\label{sec:intro}Introduction}
88
89 There are a number of excellent molecular dynamics packages available
90 to the chemical physics
91 community.\cite{Brooks83,MacKerell98,pearlman:1995,Gromacs,Gromacs3,DL_POLY,Tinker,Paradyn,namd,macromodel}
92 All of these packages are stable, polished programs which solve many
93 problems of interest. Most are now capable of performing molecular
94 dynamics simulations on parallel computers. Some have source code
95 which is freely available to the entire scientific community. Few,
96 however, are capable of efficiently integrating the equations of
97 motion for atom types with orientational degrees of freedom
98 (e.g. point dipoles, and ``sticky'' atoms). And only one of the
99 programs referenced can handle transition metal force fields like the
100 Embedded Atom Method ({\sc eam}). The direction our research program
101 has taken us now involves the use of atoms with orientational degrees
102 of freedom as well as transition metals. Since these simulation
103 methods may be of some use to other researchers, we have decided to
104 release our program (and all related source code) to the scientific
105 community.
106
107 This document communicates the algorithmic details of our program,
108 {\sc OpenMD}. We have structured this document to first discuss the
109 underlying concepts in this simulation package (Sec.
110 \ref{section:IOfiles}). The empirical energy functions implemented
111 are discussed in Sec.~\ref{section:empiricalEnergy}.
112 Sec.~\ref{section:mechanics} describes the various Molecular Dynamics
113 algorithms {\sc OpenMD} implements in the integration of Hamilton's
114 equations of motion. Program design considerations for parallel
115 computing are presented in Sec.~\ref{section:parallelization}.
116 Concluding remarks are presented in Sec.~\ref{section:conclusion}.
117
118 \chapter{\label{section:IOfiles}Concepts \& Files}
119
120 A simulation in {\sc OpenMD} is built using a few fundamental
121 conceptual building blocks most of which are chemically intuitive.
122 The basic unit of a simulation is an {\tt atom}. The parameters
123 describing an {\tt atom} have been generalized to make it as flexible
124 as possible; this means that in addition to translational degrees of
125 freedom, {\tt Atoms} may also have {\it orientational} degrees of freedom.
126
127 The fundamental (static) properties of {\tt atoms} are defined by the
128 {\tt forceField} chosen for the simulation. The atomic properties
129 specified by a {\tt forceField} might include (but are not limited to)
130 charge, $\sigma$ and $\epsilon$ values for Lennard-Jones interactions,
131 the strength of the dipole moment ($\mu$), the mass, and the moments
132 of inertia. Other more complicated properties of atoms might also be
133 specified by the {\tt forceField}.
134
135 {\tt Atoms} can be grouped together in many ways. A {\tt rigidBody}
136 contains atoms that exert no forces on one another and which move as a
137 single rigid unit. A {\tt cutoffGroup} may contain atoms which
138 function together as a (rigid {\it or} non-rigid) unit for potential
139 energy calculations,
140 \begin{equation}
141 V_{ab} = s(r_{ab}) \sum_{i \in a} \sum_{j \in b} V_{ij}(r_{ij})
142 \end{equation}
143 Here, $a$ and $b$ are two {\tt cutoffGroups} containing multiple atoms
144 ($a = \left\{i\right\}$ and $b = \left\{j\right\}$). $s(r_{ab})$ is a
145 generalized switching function which insures that the atoms in the two
146 {\tt cutoffGroups} are treated identically as the two groups enter or
147 leave an interaction region.
148
149 {\tt Atoms} may also be grouped in more traditional ways into {\tt
150 bonds}, {\tt bends}, {\tt torsions}, and {\tt inversions}. These
151 groupings allow the correct choice of interaction parameters for
152 short-range interactions to be chosen from the definitions in the {\tt
153 forceField}.
154
155 All of these groups of {\tt atoms} are brought together in the {\tt
156 molecule}, which is the fundamental structure for setting up and {\sc
157 OpenMD} simulation. {\tt Molecules} contain lists of {\tt atoms}
158 followed by listings of the other atomic groupings ({\tt bonds}, {\tt
159 bends}, {\tt torsions}, {\tt rigidBodies}, and {\tt cutoffGroups})
160 which relate the atoms to one another. Since a {\tt rigidBody} is a
161 collection of atoms that are propagated in fixed relationships to one
162 another, {\sc OpenMD} uses an internal structure called a {\tt
163 StuntDouble} to store information about those objects that can change
164 position {\it independently} during a simulation. That is, an atom
165 that is part of a rigid body is not itself a StuntDouble. In this
166 case, the rigid body is the StuntDouble. However, an atom that is
167 free to move independently {\it is} its own StuntDouble.
168
169 Simulations often involve heterogeneous collections of molecules. To
170 specify a mixture of {\tt molecule} types, {\sc OpenMD} uses {\tt
171 components}. Even simulations containing only one type of molecule
172 must specify a single {\tt component}.
173
174 Starting a simulation requires two types of information: {\it
175 meta-data}, which describes the types of objects present in the
176 simulation, and {\it configuration} information, which describes the
177 initial state of these objects. An {\sc OpenMD} file is a single
178 combined file format that describes both of these kinds of data. An
179 {\sc OpenMD} file contains one {\tt $<$MetaData$>$} block and {\it at least
180 one} {\tt $<$Snapshot$>$} block.
181
182 The language for the {\tt $<$MetaData$>$} block is a C-based syntax that
183 is parsed at the beginning of the simulation. Configuration
184 information is specified for all {\tt integrableObjects} in a {\tt
185 $<$Snapshot$>$} block. Both the {\tt $<$MetaData$>$} and {\tt $<$Snapshot$>$}
186 formats are described in the following sections.
187
188 \begin{lstlisting}[float,caption={[The structure of an {\sc OpenMD} file]
189 The basic structure of an {\sc OpenMD} file contains HTML-like tags to
190 define simulation meta-data and subsequent instantaneous configuration
191 information. A well-formed {\sc OpenMD} file must contain one $<$MetaData$>$
192 block and {\it at least one} $<$Snapshot$>$ block. Each
193 $<$Snapshot$>$ is further divided into $<$FrameData$>$ and
194 $<$StuntDoubles$>$ sections.},
195 label=sch:mdFormat]
196 <OpenMD>
197 <MetaData>
198 // see section ~\ref{sec:miscConcepts}~ for details on the formatting
199 // of information contained inside the <MetaData> tags
200 </MetaData>
201 <Snapshot> // An instantaneous configuration
202 <FrameData>
203 // FrameData contains information on the time
204 // stamp, the size of the simulation box, and
205 // the current state of extended system
206 // ensemble variables.
207 </FrameData>
208 <StuntDoubles>
209 // StuntDouble information comprises the
210 // positions, velocities, orientations, and
211 // angular velocities of anything that is
212 // capable of independent motion during
213 // the simulation.
214 </StuntDoubles>
215 </Snapshot>
216 <Snapshot> // Multiple <Snapshot> sections can be
217 </Snapshot> // present in a well-formed OpenMD file
218 <Snapshot> // Further information on <Snapshot> blocks
219 </Snapshot> // can be found in section ~\ref{section:coordFiles}~.
220 </OpenMD>
221 \end{lstlisting}
222
223
224 \section{OpenMD Files and $<$MetaData$>$ blocks}
225
226 {\sc OpenMD} uses a HTML-like syntax to separate {\tt $<$MetaData$>$} and
227 {\tt $<$Snapshot$>$} blocks. A C-based syntax is used to parse the {\tt
228 $<$MetaData$>$} blocks at run time. These blocks allow the user to
229 completely describe the system they wish to simulate, as well as
230 tailor {\sc OpenMD}'s behavior during the simulation. {\sc OpenMD}
231 files are typically denoted with the extension {\tt .md} (which can
232 stand for Meta-Data or Molecular Dynamics or Molecule Definition
233 depending on the user's mood). An overview of an {\sc OpenMD} file is
234 shown in Scheme~\ref{sch:mdFormat} and example file is shown in
235 Scheme~\ref{sch:mdExample}.
236
237 \begin{lstlisting}[float,caption={[An example of a complete OpenMD
238 file] An example showing a complete OpenMD file.},
239 label={sch:mdExample}]
240 <OpenMD>
241 <MetaData>
242 molecule{
243 name = "Ar";
244 atom[0]{
245 type="Ar";
246 position( 0.0, 0.0, 0.0 );
247 }
248 }
249
250 component{
251 type = "Ar";
252 nMol = 3;
253 }
254
255 forceField = "LJ";
256 ensemble = "NVE"; // specify the simulation ensemble
257 dt = 1.0; // the time step for integration
258 runTime = 1e3; // the total simulation run time
259 sampleTime = 100; // trajectory file frequency
260 statusTime = 50; // statistics file frequency
261 </MetaData>
262 <Snapshot>
263 <FrameData>
264 Time: 0
265 Hmat: {{ 28.569, 0, 0 }, { 0, 28.569, 0 }, { 0, 0, 28.569 }}
266 Thermostat: 0 , 0
267 Barostat: {{ 0, 0, 0 }, { 0, 0, 0 }, { 0, 0, 0 }}
268 </FrameData>
269 <StuntDoubles>
270 0 pv 17.5 13.3 12.8 1.181e-03 -1.630e-03 -1.369e-03
271 1 pv -12.8 -14.9 -8.4 -4.440e-04 -2.048e-03 1.130e-03
272 2 pv -10.0 -15.2 -6.5 2.239e-03 -6.310e-03 1.810e-03
273 </StuntDoubles>
274 </Snapshot>
275 </OpenMD>
276 \end{lstlisting}
277
278 Within the {\tt $<$MetaData$>$} block it is necessary to provide a
279 complete description of the molecule before it is actually placed in
280 the simulation. {\sc OpenMD}'s meta-data syntax was originally
281 developed with this goal in mind, and allows for the use of {\it
282 include files} to specify all atoms in a molecular prototype, as well
283 as any bonds, bends, or torsions. Include files allow the user to
284 describe a molecular prototype once, then simply include it into each
285 simulation containing that molecule. Returning to the example in
286 Scheme~\ref{sch:mdExample}, the include file's contents would be
287 Scheme~\ref{sch:mdIncludeExample}, and the new {\sc OpenMD} file would
288 become Scheme~\ref{sch:mdExPrime}.
289
290 \begin{lstlisting}[float,caption={An example molecule definition in an
291 include file.},label={sch:mdIncludeExample}]
292 molecule{
293 name = "Ar";
294 atom[0]{
295 type="Ar";
296 position( 0.0, 0.0, 0.0 );
297 }
298 }
299 \end{lstlisting}
300
301 \begin{lstlisting}[float,caption={Revised OpenMD input file
302 example.},label={sch:mdExPrime}]
303 <OpenMD>
304 <MetaData>
305 #include "argon.md"
306
307 component{
308 type = "Ar";
309 nMol = 3;
310 }
311
312 forceField = "LJ";
313 ensemble = "NVE";
314 dt = 1.0;
315 runTime = 1e3;
316 sampleTime = 100;
317 statusTime = 50;
318 </MetaData>
319 </MetaData>
320 <Snapshot>
321 <FrameData>
322 Time: 0
323 Hmat: {{ 28.569, 0, 0 }, { 0, 28.569, 0 }, { 0, 0, 28.569 }}
324 Thermostat: 0 , 0
325 Barostat: {{ 0, 0, 0 }, { 0, 0, 0 }, { 0, 0, 0 }}
326 </FrameData>
327 <StuntDoubles>
328 0 pv 17.5 13.3 12.8 1.181e-03 -1.630e-03 -1.369e-03
329 1 pv -12.8 -14.9 -8.4 -4.440e-04 -2.048e-03 1.130e-03
330 2 pv -10.0 -15.2 -6.5 2.239e-03 -6.310e-03 1.810e-03
331 </StuntDoubles>
332 </Snapshot>
333 </OpenMD>
334 \end{lstlisting}
335
336 \section{\label{section:atomsMolecules}Atoms, Molecules, and other
337 ways of grouping atoms}
338
339 As mentioned above, the fundamental unit for an {\sc OpenMD} simulation
340 is the {\tt atom}. Atoms can be collected into secondary structures
341 such as {\tt rigidBodies}, {\tt cutoffGroups}, or {\tt molecules}. The
342 {\tt molecule} is a way for {\sc OpenMD} to keep track of the atoms in
343 a simulation in logical manner. Molecular units store the identities
344 of all the atoms and rigid bodies associated with themselves, and they
345 are responsible for the evaluation of their own internal interactions
346 (\emph{i.e.}~bonds, bends, and torsions). Scheme
347 \ref{sch:mdIncludeExample} shows how one creates a molecule in an
348 included meta-data file. The positions of the atoms given in the
349 declaration are relative to the origin of the molecule, and the origin
350 is used when creating a system containing the molecule.
351
352 One of the features that sets {\sc OpenMD} apart from most of the
353 current molecular simulation packages is the ability to handle rigid
354 body dynamics. Rigid bodies are non-spherical particles or collections
355 of particles (e.g. $\mbox{C}_{60}$) that have a constant internal
356 potential and move collectively.\cite{Goldstein01} They are not
357 included in most simulation packages because of the algorithmic
358 complexity involved in propagating orientational degrees of freedom.
359 Integrators which propagate orientational motion with an acceptable
360 level of energy conservation for molecular dynamics are relatively
361 new inventions.
362
363 Moving a rigid body involves determination of both the force and
364 torque applied by the surroundings, which directly affect the
365 translational and rotational motion in turn. In order to accumulate
366 the total force on a rigid body, the external forces and torques must
367 first be calculated for all the internal particles. The total force on
368 the rigid body is simply the sum of these external forces.
369 Accumulation of the total torque on the rigid body is more complex
370 than the force because the torque is applied to the center of mass of
371 the rigid body. The space-fixed torque on rigid body $i$ is
372 \begin{equation}
373 \boldsymbol{\tau}_i=
374 \sum_{a}\biggl[(\mathbf{r}_{ia}-\mathbf{r}_i)\times \mathbf{f}_{ia}
375 + \boldsymbol{\tau}_{ia}\biggr],
376 \label{eq:torqueAccumulate}
377 \end{equation}
378 where $\boldsymbol{\tau}_i$ and $\mathbf{r}_i$ are the torque on and
379 position of the center of mass respectively, while $\mathbf{f}_{ia}$,
380 $\mathbf{r}_{ia}$, and $\boldsymbol{\tau}_{ia}$ are the force on,
381 position of, and torque on the component particles of the rigid body.
382
383 The summation of the total torque is done in the body fixed axis of
384 each rigid body. In order to move between the space fixed and body
385 fixed coordinate axes, parameters describing the orientation must be
386 maintained for each rigid body. At a minimum, the rotation matrix
387 ($\mathsf{A}$) can be described by the three Euler angles ($\phi,
388 \theta,$ and $\psi$), where the elements of $\mathsf{A}$ are composed of
389 trigonometric operations involving $\phi, \theta,$ and
390 $\psi$.\cite{Goldstein01} In order to avoid numerical instabilities
391 inherent in using the Euler angles, the four parameter ``quaternion''
392 scheme is often used. The elements of $\mathsf{A}$ can be expressed as
393 arithmetic operations involving the four quaternions ($q_w, q_x, q_y,$
394 and $q_z$).\cite{Allen87} Use of quaternions also leads to
395 performance enhancements, particularly for very small
396 systems.\cite{Evans77}
397
398 Rather than use one of the previously stated methods, {\sc OpenMD}
399 utilizes a relatively new scheme that propagates the entire nine
400 parameter rotation matrix. Further discussion on this choice can be
401 found in Sec.~\ref{section:integrate}. An example definition of a
402 rigid body can be seen in Scheme
403 \ref{sch:rigidBody}.
404
405 \begin{lstlisting}[float,caption={[Defining rigid bodies]A sample
406 definition of a molecule containing a rigid body and a cutoff
407 group},label={sch:rigidBody}]
408 molecule{
409 name = "TIP3P";
410 atom[0]{
411 type = "O_TIP3P";
412 position( 0.0, 0.0, -0.06556 );
413 }
414 atom[1]{
415 type = "H_TIP3P";
416 position( 0.0, 0.75695, 0.52032 );
417 }
418 atom[2]{
419 type = "H_TIP3P";
420 position( 0.0, -0.75695, 0.52032 );
421 }
422
423 rigidBody[0]{
424 members(0, 1, 2);
425 }
426
427 cutoffGroup{
428 members(0, 1, 2);
429 }
430 }
431 \end{lstlisting}
432
433 \section{\label{sec:miscConcepts}Creating a $<$MetaData$>$ block}
434
435 The actual creation of a {\tt $<$MetaData$>$} block requires several key
436 components. The first part of the file needs to be the declaration of
437 all of the molecule prototypes used in the simulation. This is
438 typically done through included prototype files. Only the molecules
439 actually present in the simulation need to be declared; however, {\sc
440 OpenMD} allows for the declaration of more molecules than are
441 needed. This gives the user the ability to build up a library of
442 commonly used molecules into a single include file.
443
444 Once all prototypes are declared, the ordering of the rest of the
445 block is less stringent. The molecular composition of the simulation
446 is specified with {\tt component} statements. Each different type of
447 molecule present in the simulation is considered a separate
448 component (an example is shown in
449 Sch.~\ref{sch:mdExPrime}). The component blocks tell {\sc OpenMD} the
450 number of molecules that will be in the simulation, and the order in
451 which the components blocks are declared sets the ordering of the real
452 atoms in the {\tt $<$Snapshot$>$} block as well as in the output files. The
453 remainder of the script then sets the various simulation parameters
454 for the system of interest.
455
456 The required set of parameters that must be present in all simulations
457 is given in Table~\ref{table:reqParams}. Since the user can use {\sc
458 OpenMD} to perform energy minimizations as well as molecular dynamics
459 simulations, one of the {\tt minimizer} or {\tt ensemble} keywords
460 must be present. The {\tt ensemble} keyword is responsible for
461 selecting the integration method used for the calculation of the
462 equations of motion. An in depth discussion of the various methods
463 available in {\sc OpenMD} can be found in
464 Sec.~\ref{section:mechanics}. The {\tt minimizer} keyword selects
465 which minimization method to use, and more details on the choices of
466 minimizer parameters can be found in
467 Sec.~\ref{section:minimizer}. The {\tt forceField} statement is
468 important for the selection of which forces will be used in the course
469 of the simulation. {\sc OpenMD} supports several force fields, as
470 outlined in Sec.~\ref{section:empiricalEnergy}. The force fields are
471 interchangeable between simulations, with the only requirement being
472 that all atoms needed by the simulation are defined within the
473 selected force field.
474
475 For molecular dynamics simulations, the time step between force
476 evaluations is set with the {\tt dt} parameter, and {\tt runTime} will
477 set the time length of the simulation. Note, that {\tt runTime} is an
478 absolute time, meaning if the simulation is started at t = 10.0~ns
479 with a {\tt runTime} of 25.0~ns, the simulation will only run for an
480 additional 15.0~ns.
481
482 For energy minimizations, it is not necessary to specify {\tt dt} or
483 {\tt runTime}.
484
485 To set the initial positions and velocities of all the integrable
486 objects in the simulation, {\sc OpenMD} will use the last good {\tt
487 $<$Snapshot$>$} block that was found in the startup file that it was
488 called with. If the {\tt useInitalTime} flag is set to {\tt true},
489 the time stamp from this snapshot will also set the initial time stamp
490 for the simulation. Additional parameters are summarized in
491 Table~\ref{table:genParams}.
492
493 It is important to note the fundamental units in all files which are
494 read and written by {\sc OpenMD}. Energies are in $\mbox{kcal
495 mol}^{-1}$, distances are in $\mbox{\AA}$, times are in $\mbox{fs}$,
496 translational velocities are in $\mbox{\AA~fs}^{-1}$, and masses are
497 in $\mbox{amu}$. Orientational degrees of freedom are described using
498 quaternions (unitless, but $q_w^2 + q_x^2 + q_y^2 + q_z^2 = 1$),
499 body-fixed angular momenta ($\mbox{amu \AA}^{2} \mbox{radians
500 fs}^{-1}$), and body-fixed moments of inertia ($\mbox{amu \AA}^{2}$).
501
502 \begin{longtable}[c]{ABCD}
503 \caption{Meta-data Keywords: Required Parameters}
504 \\
505 {\bf keyword} & {\bf units} & {\bf use} & {\bf remarks} \\ \hline
506 \endhead
507 \hline
508 \endfoot
509 {\tt forceField} & string & Sets the base name for the force field file &
510 OpenMD appends a {\tt .frc} to the end of this to look for a force
511 field file.\\
512 {\tt component} & & Defines the molecular components of the system &
513 Every {\tt $<$MetaData$>$} block must have a component statement. \\
514 {\tt minimizer} & string & Chooses a minimizer & Possible minimizers
515 are SD and CG. Either {\tt ensemble} or {\tt minimizer} must be specified. \\
516 {\tt ensemble} & string & Sets the ensemble. & Possible ensembles are
517 NVE, NVT, NPTi, NPAT, NPTf, NPTxyz, LD and LangevinHull. Either {\tt ensemble}
518 or {\tt minimizer} must be specified. \\
519 {\tt dt} & fs & Sets the time step. & Selection of {\tt dt} should be
520 small enough to sample the fastest motion of the simulation. ({\tt
521 dt} is required for molecular dynamics simulations)\\
522 {\tt runTime} & fs & Sets the time at which the simulation should
523 end. & This is an absolute time, and will end the simulation when the
524 current time meets or exceeds the {\tt runTime}. ({\tt runTime} is
525 required for molecular dynamics simulations)
526 \label{table:reqParams}
527 \end{longtable}
528
529 \begin{longtable}[c]{ABCD}
530 \caption{Meta-data Keywords: Optional Parameters}
531 \\
532 {\bf keyword} & {\bf units} & {\bf use} & {\bf remarks} \\ \hline
533 \endhead
534 \hline
535 \endfoot
536 {\tt forceFieldVariant} & string & Sets the name of the variant of the
537 force field. & {\sc eam} has three variants: {\tt u3}, {\tt u6}, and
538 {\tt VC}. \\
539 {\tt forceFieldFileName} & string & Overrides the default force field
540 file name & Each force field has a default file name, and this
541 parameter can override the default file name for the chosen force
542 field. \\
543 {\tt usePeriodicBoundaryConditions} & & & \\
544 & logical & Turns periodic boundary conditions on/off. & Default is true. \\
545 {\tt orthoBoxTolerance} & double & & decides how orthogonal the periodic
546 box must be before we can use cheaper box calculations \\
547 {\tt cutoffRadius} & $\mbox{\AA}$ & Manually sets the cutoff radius &
548 the default value is set by the {\tt cutoffPolicy} \\
549 {\tt cutoffPolicy} & string & one of mix, max, or
550 traditional & the traditional cutoff policy is to set the cutoff
551 radius for all atoms in the system to the same value (governed by the
552 largest atom). mix and max are pair-dependent cutoff
553 methods. \\
554 {\tt skinThickness} & \AA & thickness of the skin for the Verlet
555 neighbor lists & defaults to 1 \AA \\
556 {\tt switchingRadius} & $\mbox{\AA}$ & Manually sets the inner radius
557 for the switching function. & Defaults to 85~\% of the {\tt
558 cutoffRadius}. \\
559 {\tt switchingFunctionType} & & & \\
560 & string & cubic or
561 fifth\_order\_polynomial & Default is cubic. \\
562 {\tt useInitialTime} & logical & Sets whether the initial time is
563 taken from the last $<$Snapshot$>$ in the startup file. & Useful when recovering a simulation from a crashed processor. Default is false. \\
564 {\tt useInitialExtendedSystemState} & & & \\
565 & logical & keep the extended
566 system variables? & Should the extended
567 variables (the thermostat and barostat) be kept from the {\tt $<$Snapshot$>$} block? \\
568 {\tt sampleTime} & fs & Sets the frequency at which the {\tt .dump} file is written. & The default is equal to the {\tt runTime}. \\
569 {\tt resetTime} & fs & Sets the frequency at which the extended system
570 variables are reset to zero & The default is to never reset these
571 variables. \\
572 {\tt statusTime} & fs & Sets the frequency at which the {\tt .stat} file is written. & The default is equal to the {\tt sampleTime}. \\
573 {\tt finalConfig} & string & Sets the name of the final output file. & Useful when stringing simulations together. Defaults to the root name of the initial meta-data file but with an {\tt .eor} extension. \\
574 {\tt compressDumpFile} & logical & & should the {\tt .dump} file be
575 compressed on the fly? \\
576 {\tt statFileFormat} & string & columns to print in the {\tt .stat}
577 file where each column is separated by a pipe ($\mid$) symbol. & (The
578 default is the first eight of these columns in order.) \\
579 & & \multicolumn{2}{p{3.5in}}{Allowed
580 column names are: {\sc time, total\_energy, potential\_energy, kinetic\_energy,
581 temperature, pressure, volume, conserved\_quantity, hullvolume, gyrvolume,
582 translational\_kinetic, rotational\_kinetic, long\_range\_potential,
583 short\_range\_potential, vanderwaals\_potential,
584 electrostatic\_potential, metallic\_potential,
585 hydrogen\_bonding\_potential, bond\_potential, bend\_potential,
586 dihedral\_potential, inversion\_potential, raw\_potential, restraint\_potential,
587 pressure\_tensor, system\_dipole, heatflux, electronic\_temperature}} \\
588 {\tt printPressureTensor} & logical & sets whether {\sc OpenMD} will print
589 out the pressure tensor & can be useful for calculations of the bulk
590 modulus \\
591 {\tt electrostaticSummationMethod} & & & \\
592 & string & shifted\_force,
593 shifted\_potential, shifted\_force, or reaction\_field &
594 default is shifted\_force. \\
595 {\tt electrostaticScreeningMethod} & & & \\
596 & string & undamped or damped & default is damped \\
597 {\tt dielectric} & unitless & Sets the dielectric constant for
598 reaction field. & If {\tt electrostaticSummationMethod} is set to {\tt
599 reaction\_field}, then {\tt dielectric} must be set. \\
600 {\tt dampingAlpha} & $\mbox{\AA}^{-1}$ & governs strength of
601 electrostatic damping & defaults to 0.2 $\mbox{\AA}^{-1}$. \\
602 {\tt tempSet} & logical & resample velocities from a Maxwell-Boltzmann
603 distribution set to {\tt targetTemp} & default is false. \\
604 {\tt thermalTime} & fs & how often to perform a {\tt tempSet} &
605 default is never \\
606 {\tt targetTemp} & K & sets the target temperature & no default value \\
607 {\tt tauThermostat} & fs & time constant for Nos\'{e}-Hoover
608 thermostat & times from 1000-10,000 fs are reasonable \\
609 {\tt targetPressure} & atm & sets the target pressure & no default value\\
610 {\tt surfaceTension} & & sets the target surface tension in the x-y
611 plane & no default value \\
612 {\tt tauBarostat} & fs & time constant for the
613 Nos\'{e}-Hoover-Andersen barostat & times from 10,000 to 100,000 fs
614 are reasonable \\
615 {\tt seed } & integer & Sets the seed value for the random number generator. & The seed needs to be at least 9 digits long. The default is to take the seed from the CPU clock. \\
616 \label{table:genParams}
617 \end{longtable}
618
619
620 \section{\label{section:coordFiles}$<$Snapshot$>$ Blocks}
621
622 The standard format for storage of a system's coordinates is the {\tt
623 $<$Snapshot$>$} block , the exact details of which can be seen in
624 Scheme~\ref{sch:dumpFormat}. As all bonding and molecular information
625 is stored in the {\tt $<$MetaData$>$} blocks, the {\tt $<$Snapshot$>$} blocks
626 contain only the coordinates of the objects which move independently
627 during the simulation. It is important to note that {\it not all
628 atoms} are capable of independent motion. Atoms which are part of
629 rigid bodies are not ``integrable objects'' in the equations of
630 motion; the rigid bodies themselves are the integrable objects.
631 Therefore, the coordinate file contains coordinates of all the {\tt
632 integrableObjects} in the system. For systems without rigid bodies,
633 this is simply the coordinates of all the atoms.
634
635 It is important to note that although the simulation propagates the
636 complete rotation matrix, directional entities are written out using
637 quaternions to save space in the output files.
638
639 \begin{lstlisting}[float,caption={[The format of the {\tt $<$Snapshot$>$} block]
640 An example of the format of the {\tt $<$Snapshot$>$} block. There is an
641 initial sub-block called {\tt $<$FrameData$>$} which contains the time
642 stamp, the three column vectors of $\mathsf{H}$, and optional extra
643 information for the extended sytem ensembles. The lines in the {\tt
644 $<$StuntDoubles$>$} sub-block provide information about the instantaneous
645 configuration of each integrable object. For each integrable object,
646 the global index is followed by a short string describing what
647 additional information is present on the line. Atoms with only
648 position and velocity information use the ``pv'' string which must
649 then be followed by the position and velocity vectors for that atom.
650 Directional atoms and Rigid Bodies typically use the ``pvqj'' string
651 which is followed by position, velocity, quaternions, and
652 lastly, body fixed angular momentum for that integrable object.},
653 label=sch:dumpFormat]
654 <Snapshot>
655 <FrameData>
656 Time: 0
657 Hmat: {{ Hxx, Hyx, Hzx }, { Hxy, Hyy, Hzy }, { Hxz, Hyz, Hzz }}
658 Thermostat: 0 , 0
659 Barostat: {{ 0, 0, 0 }, { 0, 0, 0 }, { 0, 0, 0 }}
660 </FrameData>
661 <StuntDoubles>
662 0 pv x y z vx vy vz
663 1 pv x y z vx vy vz
664 2 pvqj x y z vx vy vz qw qx qy qz jx jy jz
665 3 pvqj x y z vx vy vz qw qx qy qz jx jy jz
666 </StuntDoubles>
667 </Snapshot>
668 \end{lstlisting}
669
670 There are three {\sc OpenMD} files that are written using the combined
671 format. They are: the initial startup file (\texttt{.md}), the
672 simulation trajectory file (\texttt{.dump}), and the final coordinates
673 or ``end-of-run'' for the simulation (\texttt{.eor}). The initial
674 startup file is necessary for {\sc OpenMD} to start the simulation with
675 the proper coordinates, and this file must be generated by the user
676 before the simulation run. The trajectory (or ``dump'') file is
677 updated during simulation and is used to store snapshots of the
678 coordinates at regular intervals. The first frame is a duplication of
679 the initial configuration (the last good {\tt $<$Snapshot$>$} in the
680 startup file), and each subsequent frame is appended to the dump file
681 at an interval specified in the meta-data file with the
682 \texttt{sampleTime} flag. The final coordinate file is the
683 ``end-of-run'' file. The \texttt{.eor} file stores the final
684 configuration of the system for a given simulation. The file is
685 updated at the same time as the \texttt{.dump} file, but it only
686 contains the most recent frame. In this way, an \texttt{.eor} file may
687 be used to initialize a second simulation should it be necessary to
688 recover from a crash or power outage. The coordinate files generated
689 by {\sc OpenMD} (both \texttt{.dump} and \texttt{.eor}) all contain the
690 same {\tt $<$MetaData$>$} block as the startup file, so they may be
691 used to start up a new simulation if desired.
692
693 \section{\label{section:initCoords}Generation of Initial Coordinates}
694
695 As was stated in Sec.~\ref{section:coordFiles}, a meaningful {\tt
696 $<$Snapshot$>$} block is necessary for specifying for the starting
697 coordinates for a simulation. Since each simulation is different,
698 system creation is left to the end user; however, we have included a
699 few sample programs which make some specialized structures. The {\tt
700 $<$Snapshot$>$} block must index the integrable objects in the correct
701 order. The ordering of the integrable objects relies on the ordering
702 of molecules within the {\tt $<$MetaData$>$} block. {\sc OpenMD}
703 expects the order to comply with the following guidelines:
704 \begin{enumerate}
705 \item All of the molecules of the first declared component are given
706 before proceeding to the molecules of the second component, and so on
707 for all subsequently declared components.
708 \item The ordering of the atoms for each molecule follows the order
709 declared in the molecule's declaration within the model file.
710 \item Only atoms which are not members of a {\tt rigidBody} are
711 included.
712 \item Rigid Body coordinates for a molecule are listed immediately
713 after the the other atoms in a molecule. Some molecules may be
714 entirely rigid, in which case, only the rigid body coordinates are
715 given.
716 \end{enumerate}
717 An example is given in the {\sc OpenMD} file in Scheme~\ref{sch:initEx1}.
718
719 \begin{lstlisting}[float,caption={Example declaration of the
720 $\text{I}_2$ molecule and the HCl molecule in $<$MetaData$>$ and
721 $<$Snapshot$>$ blocks. Note that even though $\text{I}_2$ is
722 declared before HCl, the $<$Snapshot$>$ block follows the order {\it in
723 which the components were included}.}, label=sch:initEx1]
724 <OpenMD>
725 <MetaData>
726 molecule{
727 name = "I2";
728 atom[0]{ type = "I"; }
729 atom[1]{ type = "I"; }
730 bond{ members( 0, 1); }
731 }
732 molecule{
733 name = "HCl"
734 atom[0]{ type = "H";}
735 atom[1]{ type = "Cl";}
736 bond{ members( 0, 1); }
737 }
738 component{
739 type = "HCl";
740 nMol = 4;
741 }
742 component{
743 type = "I2";
744 nMol = 1;
745 }
746 </MetaData>
747 <Snapshot>
748 <FrameData>
749 Time: 0
750 Hmat: {{ 10.0, 0.0, 0.0 }, { 0.0, 10.0, 0.0 }, { 0.0, 0.0, 10.0 }}
751 </FrameData>
752 <StuntDoubles>
753 0 pv x y z vx vy vz // H from first HCl molecule
754 1 pv x y z vx vy vz // Cl from first HCl molecule
755 2 pv x y z vx vy vz // H from second HCl molecule
756 3 pv x y z vx vy vz // Cl from second HCl molecule
757 4 pv x y z vx vy vz // H from third HCl molecule
758 5 pv x y z vx vy vz // Cl from third HCl molecule
759 6 pv x y z vx vy vz // H from fourth HCl molecule
760 7 pv x y z vx vy vz // Cl from fourth HCl molecule
761 8 pv x y z vx vy vz // First I from I2 molecule
762 9 pv x y z vx vy vz // Second I from I2 molecule
763 </StuntDoubles>
764 </Snapshot>
765 </OpenMD>
766 \end{lstlisting}
767
768 \section{The Statistics File}
769
770 The last output file generated by {\sc OpenMD} is the statistics
771 file. This file records such statistical quantities as the
772 instantaneous temperature (in $K$), volume (in $\mbox{\AA}^{3}$),
773 pressure (in $\mbox{atm}$), etc. It is written out with the frequency
774 specified in the meta-data file with the
775 \texttt{statusTime} keyword. The file allows the user to observe the
776 system variables as a function of simulation time while the simulation
777 is in progress. One useful function the statistics file serves is to
778 monitor the conserved quantity of a given simulation ensemble,
779 allowing the user to gauge the stability of the integrator. The
780 statistics file is denoted with the \texttt{.stat} file extension.
781
782 \chapter{\label{chapter:forceFields}Force Fields}
783
784 Like many molecular simulation packages, {\sc OpenMD} splits the
785 potential energy into the short-ranged (bonded) portion and a
786 long-range (non-bonded) potential,
787 \begin{equation}
788 V = V_{\mathrm{short-range}} + V_{\mathrm{long-range}}.
789 \end{equation}
790 The short-ranged portion includes the bonds, bends, torsions, and
791 inversions which have been defined in the meta-data file for the
792 molecules. The functional forms and parameters for these interactions
793 are defined by the force field which is selected in the MetaData
794 section.
795
796 \section{\label{section:divisionOfLabor}Separation into Internal and
797 Cross interactions}
798
799 The classical potential energy function for a system composed of $N$
800 molecules is traditionally written
801 \begin{equation}
802 V = \sum^{N}_{I=1} V^{I}_{\text{Internal}}
803 + \sum^{N-1}_{I=1} \sum_{J>I} V^{IJ}_{\text{Cross}},
804 \label{eq:totalPotential}
805 \end{equation}
806 where $V^{I}_{\text{Internal}}$ contains all of the terms internal to
807 molecule $I$ (e.g. bonding, bending, torsional, and inversion terms)
808 and $V^{IJ}_{\text{Cross}}$ contains all intermolecular interactions
809 between molecules $I$ and $J$. For large molecules, the internal
810 potential may also include some non-bonded terms like electrostatic or
811 van der Waals interactions.
812
813 The types of atoms being simulated, as well as the specific functional
814 forms and parameters of the intra-molecular functions and the
815 long-range potentials are defined by the force field. In the following
816 sections we discuss the stucture of an OpenMD force field file and the
817 specification of blocks that may be present within these files.
818
819 \section{\label{section:frcFile}Force Field Files}
820
821 Force field files have a number of ``Blocks'' to delineate different
822 types of information. The blocks contain AtomType data, which provide
823 properties belonging to a single AtomType, as well as interaction
824 information which provides information about bonded or non-bonded
825 interactions that cannot be deduced from AtomType information alone.
826 A simple example of a forceField file is shown in scheme
827 \ref{sch:frcExample}.
828
829 \begin{lstlisting}[float,caption={[An example of a complete OpenMD
830 force field file for straight-chain united-atom alkanes.] An example
831 showing a complete OpenMD force field for straight-chain united-atom
832 alkanes.}, label={sch:frcExample}]
833 begin Options
834 Name = "alkane"
835 end Options
836
837 begin BaseAtomTypes
838 //name mass
839 C 12.0107
840 end BaseAtomTypes
841
842 begin AtomTypes
843 //name base mass
844 CH4 C 16.05
845 CH3 C 15.04
846 CH2 C 14.03
847 end AtomTypes
848
849 begin LennardJonesAtomTypes
850 //name epsilon sigma
851 CH4 0.2941 3.73
852 CH3 0.1947 3.75
853 CH2 0.09140 3.95
854 end LennardJonesAtomTypes
855
856 begin BondTypes
857 //AT1 AT2 Type r0 k
858 CH3 CH3 Harmonic 1.526 260
859 CH3 CH2 Harmonic 1.526 260
860 CH2 CH2 Harmonic 1.526 260
861 end BondTypes
862
863 begin BendTypes
864 //AT1 AT2 AT3 Type theta0 k
865 CH3 CH2 CH3 Harmonic 114.0 124.19
866 CH3 CH2 CH2 Harmonic 114.0 124.19
867 CH2 CH2 CH2 Harmonic 114.0 124.19
868 end BendTypes
869
870 begin TorsionTypes
871 //AT1 AT2 AT3 AT4 Type
872 CH3 CH2 CH2 CH3 Trappe 0.0 0.70544 -0.13549 1.5723
873 CH3 CH2 CH2 CH2 Trappe 0.0 0.70544 -0.13549 1.5723
874 CH2 CH2 CH2 CH2 Trappe 0.0 0.70544 -0.13549 1.5723
875 end TorsionTypes
876 \end{lstlisting}
877
878 \section{\label{section:ffOptions}The Options block}
879
880 The Options block defines properties governing how the force field
881 interactions are carried out. This block is delineated with the text
882 tags {\tt begin Options} and {\tt end Options}. Most options don't
883 need to be set as they come with fairly sensible default values, but
884 the various keywords and their possible values are given in Scheme
885 \ref{sch:optionsBlock}.
886
887 \begin{lstlisting}[caption={[A force field Options block showing default values
888 for many force field options.] A force field Options block showing default values
889 for many force field options. Most of these options do not need to be
890 specified if the default values are working.},
891 label={sch:optionsBlock}]
892 begin Options
893 Name = "alkane" // any string
894 vdWtype = "Lennard-Jones"
895 DistanceMixingRule = "arithmetic" // can also be "geometric" or "cubic"
896 DistanceType = "sigma" // can also be "Rmin"
897 EnergyMixingRule = "geometric" // can also be "arithmetic" or "hhg"
898 EnergyUnitScaling = 1.0
899 MetallicEnergyUnitScaling = 1.0
900 DistanceUnitScaling = 1.0
901 AngleUnitScaling = 1.0
902 TorsionAngleConvention = "180_is_trans" // can also be "0_is_trans"
903 vdW-12-scale = 0.0
904 vdW-13-scale = 0.0
905 vdW-14-scale = 0.0
906 electrostatic-12-scale = 0.0
907 electrostatic-13-scale = 0.0
908 electrostatic-14-scale = 0.0
909 GayBerneMu = 2.0
910 GayBerneNu = 1.0
911 EAMMixingMethod = "Johnson" // can also be "Daw"
912 end Options
913 \end{lstlisting}
914
915 \section{\label{section:ffBase}The BaseAtomTypes block}
916
917 An AtomType the primary data structure that OpenMD uses to store
918 static data about an atom. Things that belong to AtomType are
919 universal properties (i.e. does this atom have a fixed charge? What
920 is its mass?) Dynamic properties of an atom are not intended to be
921 properties of an atom type. A BaseAtomType can be used to build
922 extended sets of related atom types that all fall back to one
923 particular type. For example, one might want a series of atomTypes
924 that inherit from more basic types:
925 \begin{displaymath}
926 \mathtt{ALA-CA} \rightarrow \mathtt{CT} \rightarrow \mathtt{CSP3} \rightarrow \mathtt{C}
927 \end{displaymath}
928 where for each step to the right, the atomType falls back to more
929 primitive data. That is, the mass could be a property of the {\tt C}
930 type, while Lennard-Jones parameters could be properties of the {\tt
931 CSP3} type. {\tt CT} could have charge information and its own set
932 of Lennard-Jones parameter that override the CSP3 parameters. And the
933 {\tt ALA-CA} type might have specific torsion or charge information
934 that override the lower level types. A BaseAtomType contains only
935 information a primitive name and the mass of this atom type.
936 BaseAtomTypes can also be useful in creating files that can be easily
937 viewed in visualization programs. The {\tt Dump2XYZ} utility has the
938 ability to print out the names of the base atom types for displaying
939 simulations in Jmol or VMD.
940
941 \begin{lstlisting}[caption={[A simple example of a BaseAtomTypes
942 block.] A simple example of a BaseAtomTypes block.},
943 label={sch:baseAtomTypesBlock}]
944 begin BaseAtomTypes
945 //Name mass (amu)
946 H 1.0079
947 O 15.9994
948 Si 28.0855
949 Al 26.981538
950 Mg 24.3050
951 Ca 40.078
952 Fe 55.845
953 Li 6.941
954 Na 22.98977
955 K 39.0983
956 Cs 132.90545
957 Ca 40.078
958 Ba 137.327
959 Cl 35.453
960 end BaseAtomTypes
961 \end{lstlisting}
962
963 \section{\label{section:ffAtom}The AtomTypes block}
964
965 AtomTypes inherit most properties from BaseAtomTypes, but can override
966 their lower-level properties as well. Scheme \ref{sch:atomTypesBlock}
967 shows an example where multiple types of oxygen atoms can inherit mass
968 from the oxygen base type.
969
970 \begin{lstlisting}[caption={[An example of a AtomTypes block.] A
971 simple example of an AtomTypes block which
972 shows how multiple types can inherit from the same base type.},
973 label={sch:atomTypesBlock}]
974 begin AtomTypes
975 //Name baseatomtype
976 h* H
977 ho H
978 o* O
979 oh O
980 ob O
981 obos O
982 obts O
983 obss O
984 ohs O
985 st Si
986 ao Al
987 at Al
988 mgo Mg
989 mgh Mg
990 cao Ca
991 cah Ca
992 feo Fe
993 lio Li
994 end AtomTypes
995 \end{lstlisting}
996
997 \section{\label{section:ffDirectionalAtom}The DirectionalAtomTypes
998 block}
999 DirectionalAtoms have orientational degrees of freedom as well as
1000 translation, so moving these atoms requires information about the
1001 moments of inertias in the same way that translational motion requires
1002 mass. For DirectionalAtoms, OpenMD treats the mass distribution with
1003 higher priority than electrostatic distributions; the moment of
1004 inertia tensor, $\overleftrightarrow{\mathsf I}$, should be
1005 diagonalized to obtain body-fixed axes, and the three diagonal moments
1006 should correspond to rotational motion \textit{around} each of these
1007 body-fixed axes. Charge distributions may then result in dipole
1008 vectors that are oriented along a linear combination of the body-axes,
1009 and in quadrupole tensors that are not necessarily diagonal in the
1010 body frame.
1011
1012 \begin{lstlisting}[caption={[An example of a DirectionalAtomTypes block.] A
1013 simple example of a DirectionalAtomTypes block.},
1014 label={sch:datomTypesBlock}]
1015 begin DirectionalAtomTypes
1016 //Name I_xx I_yy I_zz (All moments in (amu*Ang^2)
1017 SSD 1.7696 0.6145 1.1550
1018 GBC6H6 88.781 88.781 177.561
1019 GBCH3OH 4.056 20.258 20.999
1020 GBH2O 1.777 0.581 1.196
1021 CO2 43.06 43.06 0.0 // single-site model for CO2
1022 end DirectionalAtomTypes
1023
1024 \end{lstlisting}
1025
1026 For a DirectionalAtom that represents a linear object, it is
1027 appropriate for one of the moments of inertia to be zero. In this
1028 case, OpenMD identifies that DirectionalAtom as having only 5 degrees
1029 of freedom (three translations and two rotations), and will alter
1030 calculation of temperatures to reflect this.
1031
1032 \section{\label{section::ffAtomProperties}AtomType properties}
1033 \subsection{\label{section:ffLJ}The LennardJonesAtomTypes block}
1034 One of the most basic interatomic interactions implemented in {\sc
1035 OpenMD} is the Lennard-Jones potential, which mimics the van der
1036 Waals interaction at long distances and uses an empirical repulsion at
1037 short distances. The Lennard-Jones potential is given by:
1038 \begin{equation}
1039 V_{\text{LJ}}(r_{ij}) =
1040 4\epsilon_{ij} \biggl[
1041 \biggl(\frac{\sigma_{ij}}{r_{ij}}\biggr)^{12}
1042 - \biggl(\frac{\sigma_{ij}}{r_{ij}}\biggr)^{6}
1043 \biggr],
1044 \label{eq:lennardJonesPot}
1045 \end{equation}
1046 where $r_{ij}$ is the distance between particles $i$ and $j$,
1047 $\sigma_{ij}$ scales the length of the interaction, and
1048 $\epsilon_{ij}$ scales the well depth of the potential.
1049
1050 Interactions between dissimilar particles requires the generation of
1051 cross term parameters for $\sigma$ and $\epsilon$. These parameters
1052 are usually determined using the Lorentz-Berthelot mixing
1053 rules:\cite{Allen87}
1054 \begin{equation}
1055 \sigma_{ij} = \frac{1}{2}[\sigma_{ii} + \sigma_{jj}],
1056 \label{eq:sigmaMix}
1057 \end{equation}
1058 and
1059 \begin{equation}
1060 \epsilon_{ij} = \sqrt{\epsilon_{ii} \epsilon_{jj}}.
1061 \label{eq:epsilonMix}
1062 \end{equation}
1063
1064 The values of $\sigma_{ii}$ and $\epsilon_{ii}$ are properties of atom
1065 type $i$, and must be specified in a section of the force field file
1066 called the {\tt LennardJonesAtomTypes} block (see listing
1067 \ref{sch:LJatomTypesBlock}). Separate Lennard-Jones interactions
1068 which are not determined by the mixing rules can also be specified in
1069 the {\tt NonbondedInteractionTypes} block (see section
1070 \ref{section:ffNBinteraction}).
1071
1072 \begin{lstlisting}[caption={[An example of a LennardJonesAtomTypes block.] A
1073 simple example of a LennardJonesAtomTypee block. Units for
1074 $\epsilon$ are kcal / mol and for $\sigma$ are \AA\ .},
1075 label={sch:LJatomTypesBlock}]
1076 begin LennardJonesAtomTypes
1077 //Name epsilon sigma
1078 O_TIP4P 0.1550 3.15365
1079 O_TIP4P-Ew 0.16275 3.16435
1080 O_TIP5P 0.16 3.12
1081 O_TIP5P-E 0.178 3.097
1082 O_SPCE 0.15532 3.16549
1083 O_SPC 0.15532 3.16549
1084 CH4 0.279 3.73
1085 CH3 0.185 3.75
1086 CH2 0.0866 3.95
1087 CH 0.0189 4.68
1088 end LennardJonesAtomTypes
1089 \end{lstlisting}
1090
1091 \subsection{\label{section:ffCharge}The ChargeAtomTypes block}
1092
1093 In molecular simulations, proper accumulation of the electrostatic
1094 interactions is essential and is one of the most
1095 computationally-demanding tasks. Most common molecular mechanics
1096 force fields represent atomic sites with full or partial charges
1097 protected by Lennard-Jones (short range) interactions. Partial charge
1098 values, $q_i$ are empirical representations of the distribution of
1099 electronic charge in a molecule. This means that nearly every pair
1100 interaction involves a calculation of charge-charge forces. Coupled
1101 with relatively long-ranged $r^{-1}$ decay, the monopole interactions
1102 quickly become the most expensive part of molecular simulations. The
1103 interactions between point charges can be handled via a number of
1104 different algorithms, but Coulomb's law is the fundamental physical
1105 principle governing these interactions,
1106 \begin{equation}
1107 V_{\text{charge}}(r_{ij}) = \sum_{ij}\frac{q_iq_je^2}{4 \pi \epsilon_0
1108 r_{ij}},
1109 \end{equation}
1110 where $q$ represents the charge on particle $i$ or $j$, and $e$ is the
1111 charge of an electron in Coulombs. $\epsilon_0$ is the permittivity
1112 of free space.
1113
1114 \begin{lstlisting}[caption={[An example of a ChargeAtomTypes block.] A
1115 simple example of a ChargeAtomTypes block. Units for
1116 charge are in multiples of electron charge.},
1117 label={sch:ChargeAtomTypesBlock}]
1118 begin ChargeAtomTypes
1119 // Name charge
1120 O_TIP3P -0.834
1121 O_SPCE -0.8476
1122 H_TIP3P 0.417
1123 H_TIP4P 0.520
1124 H_SPCE 0.4238
1125 EP_TIP4P -1.040
1126 Na+ 1.0
1127 Cl- -1.0
1128 end ChargeAtomTypes
1129 \end{lstlisting}
1130
1131 \subsection{\label{section:ffMultipole}The MultipoleAtomTypes
1132 block}
1133 For complex charge distributions that are centered on single sites, it
1134 is convenient to write the total electrostatic potential in terms of
1135 multipole moments,
1136 \begin{equation}
1137 U_{\bf{ab}}(r)=\hat{M}_{\bf a} \hat{M}_{\bf b} \frac{1}{r} \label{kernel}.
1138 \end{equation}
1139 where the multipole operator on site $\bf a$,
1140 \begin{equation}
1141 \hat{M}_{\bf a} = C_{\bf a} - D_{{\bf a}\alpha} \frac{\partial}{\partial r_{\alpha}}
1142 + Q_{{\bf a}\alpha\beta}
1143 \frac{\partial^2}{\partial r_{\alpha} \partial r_{\beta}} + \dots
1144 \end{equation}
1145 Here, the point charge, dipole, and quadrupole for site $\bf a$ are
1146 given by $C_{\bf a}$, $D_{{\bf a}\alpha}$, and $Q_{{\bf
1147 a}\alpha\beta}$, respectively. These are the {\it primitive}
1148 multipoles. If the site is representing a distribution of charges,
1149 these can be expressed as,
1150 \begin{align}
1151 C_{\bf a} =&\sum_{k \, \text{in \bf a}} q_k , \label{eq:charge} \\
1152 D_{{\bf a}\alpha} =&\sum_{k \, \text{in \bf a}} q_k r_{k\alpha}, \label{eq:dipole}\\
1153 Q_{{\bf a}\alpha\beta} =& \frac{1}{2} \sum_{k \, \text{in \bf a}} q_k
1154 r_{k\alpha} r_{k\beta} . \label{eq:quadrupole}
1155 \end{align}
1156 Note that the definition of the primitive quadrupole here differs from
1157 the standard traceless form, and contains an additional Taylor-series
1158 based factor of $1/2$.
1159
1160 The details of the multipolar interactions will be given later, but
1161 many readers are familiar with the dipole-dipole potential:
1162 \begin{equation}
1163 V_{\text{dipole}}(\mathbf{r}_{ij},\boldsymbol{\Omega}_{i},
1164 \boldsymbol{\Omega}_{j}) = \frac{|{\bf D}_i||{\bf D}_j|}{4\pi\epsilon_{0}r_{ij}^{3}} \biggl[
1165 \boldsymbol{\hat{u}}_{i} \cdot \boldsymbol{\hat{u}}_{j}
1166 -
1167 3(\boldsymbol{\hat{u}}_i \cdot \hat{\mathbf{r}}_{ij}) %
1168 (\boldsymbol{\hat{u}}_j \cdot \hat{\mathbf{r}}_{ij}) \biggr].
1169 \label{eq:dipolePot}
1170 \end{equation}
1171 Here $\mathbf{r}_{ij}$ is the vector starting at atom $i$ pointing
1172 towards $j$, and $\boldsymbol{\Omega}_i$ and $\boldsymbol{\Omega}_j$
1173 are the orientational degrees of freedom for atoms $i$ and $j$
1174 respectively. The magnitude of the dipole moment of atom $i$ is $|{\bf
1175 D}_i|$, $\boldsymbol{\hat{u}}_i$ is the standard unit orientation
1176 vector of $\boldsymbol{\Omega}_i$, and $\boldsymbol{\hat{r}}_{ij}$ is
1177 the unit vector pointing along $\mathbf{r}_{ij}$
1178 ($\boldsymbol{\hat{r}}_{ij}=\mathbf{r}_{ij}/|\mathbf{r}_{ij}|$).
1179
1180
1181 \begin{lstlisting}[caption={[An example of a MultipoleAtomTypes block.] A
1182 simple example of a MultipoleAtomTypes block. Dipoles are given in
1183 units of Debyes, and Quadrupole moments are given in units of Debye
1184 \AA~(or $10^{-26} \mathrm{~esu~cm}^2$)},
1185 label={sch:MultipoleAtomTypesBlock}]
1186 begin MultipoleAtomTypes
1187 // Euler angles are given in zxz convention in units of degrees.
1188 //
1189 // point dipoles:
1190 // name d phi theta psi dipole_moment
1191 DIP d 0.0 0.0 0.0 1.91 // dipole points along z-body axis
1192 //
1193 // point quadrupoles:
1194 // name q phi theta psi Qxx Qyy Qzz
1195 CO2 q 0.0 0.0 0.0 0.0 0.0 -0.430592 //quadrupole tensor has zz element
1196 //
1197 // Atoms with both dipole and quadrupole moments:
1198 // name dq phi theta psi dipole_moment Qxx Qyy Qzz
1199 SSD dq 0.0 0.0 0.0 2.35 -1.682 1.762 -0.08
1200 end MultipoleAtomTypes
1201 \end{lstlisting}
1202
1203 Specifying a MultipoleAtomType requires declaring how the
1204 electrostatic frame for the site is rotated relative to the body-fixed
1205 axes for that atom. The Euler angles $(\phi, \theta, \psi)$ for this
1206 rotation must be given, and then the dipole, quadrupole, or all of
1207 these moments are specified in the electrostatic frame. In OpenMD,
1208 the Euler angles are specified in the $zxz$ convention and are entered
1209 in units of degrees. Dipole moments are entered in units of Debye,
1210 and Quadrupole moments in units of Debye \AA.
1211
1212 \subsection{\label{section:ffGB}The FluctuatingChargeAtomTypes block}
1213 %\subsubsection{\label{section:ffPol}The PolarizableAtomTypes block}
1214
1215 \subsection{\label{section:ffGB}The GayBerneAtomTypes block}
1216
1217 The Gay-Berne potential has been widely used in the liquid crystal
1218 community to describe anisotropic phase
1219 behavior.~\cite{Gay:1981yu,Berne:1972pb,Kushick:1976xy,Luckhurst:1990fy,Cleaver:1996rt}
1220 The form of the Gay-Berne potential implemented in OpenMD was
1221 generalized by Cleaver {\it et al.} and is appropriate for dissimilar
1222 uniaxial ellipsoids.\cite{Cleaver:1996rt} The potential is constructed
1223 in the familiar form of the Lennard-Jones function using
1224 orientation-dependent $\sigma$ and $\epsilon$ parameters,
1225 \begin{equation*}
1226 V_{ij}({{\bf \hat u}_i}, {{\bf \hat u}_j}, {{\bf \hat
1227 r}_{ij}}) = 4\epsilon ({{\bf \hat u}_i}, {{\bf \hat u}_j},
1228 {{\bf \hat r}_{ij}})\left[\left(\frac{\sigma_0}{r_{ij}-\sigma({{\bf \hat u
1229 }_i},
1230 {{\bf \hat u}_j}, {{\bf \hat r}_{ij}})+\sigma_0}\right)^{12}
1231 -\left(\frac{\sigma_0}{r_{ij}-\sigma({{\bf \hat u}_i}, {{\bf \hat u}_j},
1232 {{\bf \hat r}_{ij}})+\sigma_0}\right)^6\right]
1233 \label{eq:gb}
1234 \end{equation*}
1235
1236 The range $(\sigma({\bf \hat{u}}_{i},{\bf \hat{u}}_{j},{\bf
1237 \hat{r}}_{ij}))$, and strength $(\epsilon({\bf \hat{u}}_{i},{\bf
1238 \hat{u}}_{j},{\bf \hat{r}}_{ij}))$ parameters
1239 are dependent on the relative orientations of the two ellipsoids (${\bf
1240 \hat{u}}_{i},{\bf \hat{u}}_{j}$) as well as the direction of the
1241 inter-ellipsoid separation (${\bf \hat{r}}_{ij}$). The shape and
1242 attractiveness of each ellipsoid is governed by a relatively small set
1243 of parameters:
1244 \begin{itemize}
1245 \item $d$: range parameter for the side-by-side (S) and cross (X) configurations
1246 \item $l$: range parameter for the end-to-end (E) configuration
1247 \item $\epsilon_X$: well-depth parameter for the cross (X) configuration
1248 \item $\epsilon_S$: well-depth parameter for the side-by-side (S) configuration
1249 \item $\epsilon_E$: well depth parameter for the end-to-end (E) configuration
1250 \item $dw$: The ``softness'' of the potential
1251 \end{itemize}
1252 Additionally, there are two universal paramters to govern the overall
1253 importance of the purely orientational ($\nu$) and the mixed
1254 orientational / translational ($\mu$) parts of strength of the
1255 interactions. These parameters have default or ``canonical'' values,
1256 but may be changed as a force field option:
1257 \begin{itemize}
1258 \item $\nu$: purely orientational part : defaults to 1
1259 \item $\mu$: mixed orientational / translational part : defaults to
1260 2
1261 \end{itemize}
1262 Further details of the potential are given
1263 elsewhere,\cite{Luckhurst:1990fy,Golubkov06,SunX._jp0762020} and an
1264 excellent overview of the computational methods that can be used to
1265 efficiently compute forces and torques for this potential can be found
1266 in Ref. \citealp{Golubkov06}
1267
1268 \begin{lstlisting}[caption={[An example of a GayBerneAtomTypes block.] A
1269 simple example of a GayBerneAtomTypes block. Distances ($d$ and $l$)
1270 are given in \AA\ and energies ($\epsilon_X, \epsilon_S, \epsilon_E$)
1271 are in units of kcal/mol. $dw$ is unitless.},
1272 label={sch:GayBerneAtomTypes}]
1273 begin GayBerneAtomTypes
1274 //Name d l eps_X eps_S eps_E dw
1275 GBlinear 2.8104 9.993 0.774729 0.774729 0.116839 1.0
1276 GBC6H6 4.65 2.03 0.540 0.540 1.9818 0.6
1277 GBCH3OH 2.55 3.18 0.542 0.542 0.55826 1.0
1278 end GayBerneAtomTypes
1279 \end{lstlisting}
1280
1281 \subsection{\label{section:ffSticky}The StickyAtomTypes block}
1282
1283 One of the solvents that can be simulated by {\sc OpenMD} is the
1284 extended Soft Sticky Dipole (SSD/E) water model.\cite{fennell04} The
1285 original SSD was developed by Ichiye \emph{et
1286 al.}\cite{liu96:new_model} as a modified form of the hard-sphere
1287 water model proposed by Bratko, Blum, and
1288 Luzar.\cite{Bratko85,Bratko95} It consists of a single point dipole
1289 with a Lennard-Jones core and a sticky potential that directs the
1290 particles to assume the proper hydrogen bond orientation in the first
1291 solvation shell. Thus, the interaction between two SSD water molecules
1292 \emph{i} and \emph{j} is given by the potential
1293 \begin{equation}
1294 V_{ij} =
1295 V_{ij}^{LJ} (r_{ij})\ + V_{ij}^{dp}
1296 (\mathbf{r}_{ij},\boldsymbol{\Omega}_i,\boldsymbol{\Omega}_j)\ +
1297 V_{ij}^{sp}
1298 (\mathbf{r}_{ij},\boldsymbol{\Omega}_i,\boldsymbol{\Omega}_j),
1299 \label{eq:ssdPot}
1300 \end{equation}
1301 where the $\mathbf{r}_{ij}$ is the position vector between molecules
1302 \emph{i} and \emph{j} with magnitude equal to the distance $r_{ij}$, and
1303 $\boldsymbol{\Omega}_i$ and $\boldsymbol{\Omega}_j$ represent the
1304 orientations of the respective molecules. The Lennard-Jones and dipole
1305 parts of the potential are given by equations \ref{eq:lennardJonesPot}
1306 and \ref{eq:dipolePot} respectively. The sticky part is described by
1307 the following,
1308 \begin{equation}
1309 u_{ij}^{sp}(\mathbf{r}_{ij},\boldsymbol{\Omega}_i,\boldsymbol{\Omega}_j)=
1310 \frac{\nu_0}{2}[s(r_{ij})w(\mathbf{r}_{ij},
1311 \boldsymbol{\Omega}_i,\boldsymbol{\Omega}_j) +
1312 s^\prime(r_{ij})w^\prime(\mathbf{r}_{ij},
1313 \boldsymbol{\Omega}_i,\boldsymbol{\Omega}_j)]\ ,
1314 \label{eq:stickyPot}
1315 \end{equation}
1316 where $\nu_0$ is a strength parameter for the sticky potential, and
1317 $s$ and $s^\prime$ are cubic switching functions which turn off the
1318 sticky interaction beyond the first solvation shell. The $w$ function
1319 can be thought of as an attractive potential with tetrahedral
1320 geometry:
1321 \begin{equation}
1322 w({\bf r}_{ij},{\bf \Omega}_i,{\bf \Omega}_j)=
1323 \sin\theta_{ij}\sin2\theta_{ij}\cos2\phi_{ij},
1324 \label{eq:stickyW}
1325 \end{equation}
1326 while the $w^\prime$ function counters the normal aligned and
1327 anti-aligned structures favored by point dipoles:
1328 \begin{equation}
1329 w^\prime({\bf r}_{ij},{\bf \Omega}_i,{\bf \Omega}_j)=
1330 (\cos\theta_{ij}-0.6)^2(\cos\theta_{ij}+0.8)^2-w^0,
1331 \label{eq:stickyWprime}
1332 \end{equation}
1333 It should be noted that $w$ is proportional to the sum of the $Y_3^2$
1334 and $Y_3^{-2}$ spherical harmonics (a linear combination which
1335 enhances the tetrahedral geometry for hydrogen bonded structures),
1336 while $w^\prime$ is a purely empirical function. A more detailed
1337 description of the functional parts and variables in this potential
1338 can be found in the original SSD
1339 articles.\cite{liu96:new_model,liu96:monte_carlo,chandra99:ssd_md,Ichiye03}
1340
1341 \begin{figure}
1342 \centering
1343 \includegraphics[width=\linewidth]{waterAngle.pdf}
1344 \caption[Coordinate definition for the SSD/E water model]{Coordinates
1345 for the interaction between two SSD/E water molecules. $\theta_{ij}$
1346 is the angle that $r_{ij}$ makes with the $\hat{z}$ vector in the
1347 body-fixed frame for molecule $i$. The $\hat{z}$ vector bisects the
1348 HOH angle in each water molecule. }
1349 \label{fig:ssd}
1350 \end{figure}
1351
1352 Since SSD/E is a single-point {\it dipolar} model, the force
1353 calculations are simplified significantly relative to the standard
1354 {\it charged} multi-point models. In the original Monte Carlo
1355 simulations using this model, Ichiye {\it et al.} reported that using
1356 SSD decreased computer time by a factor of 6-7 compared to other
1357 models.\cite{liu96:new_model} What is most impressive is that these
1358 savings did not come at the expense of accurate depiction of the
1359 liquid state properties. Indeed, SSD/E maintains reasonable agreement
1360 with the Head-Gordon diffraction data for the structural features of
1361 liquid water.\cite{hura00,liu96:new_model} Additionally, the dynamical
1362 properties exhibited by SSD/E agree with experiment better than those
1363 of more computationally expensive models (like TIP3P and
1364 SPC/E).\cite{chandra99:ssd_md} The combination of speed and accurate
1365 depiction of solvent properties makes SSD/E a very attractive model
1366 for the simulation of large scale biochemical simulations.
1367
1368 Recent constant pressure simulations revealed issues in the original
1369 SSD model that led to lower than expected densities at all target
1370 pressures,\cite{Ichiye03,fennell04} so variants on the sticky
1371 potential can be specified by using one of a number of substitute atom
1372 types (see listing \ref{sch:StickyAtomTypes}). A table of the
1373 parameter values and the drawbacks and benefits of the different
1374 density corrected SSD models can be found in
1375 reference~\citealp{fennell04}.
1376
1377 \begin{lstlisting}[caption={[An example of a StickyAtomTypes block.] A
1378 simple example of a StickyAtomTypes block. Distances ($r_l$, $r_u$,
1379 $r_{l}'$ and $r_{u}'$) are given in \AA\ and energies ($v_0, v_{0}'$)
1380 are in units of kcal/mol. $w_0$ is unitless.},
1381 label={sch:StickyAtomTypes}]
1382 begin StickyAtomTypes
1383 //name w0 v0 (kcal/mol) v0p rl (Ang) ru rlp rup
1384 SSD_E 0.07715 3.90 3.90 2.40 3.80 2.75 3.35
1385 SSD_RF 0.07715 3.90 3.90 2.40 3.80 2.75 3.35
1386 SSD 0.07715 3.7284 3.7284 2.75 3.35 2.75 4.0
1387 SSD1 0.07715 3.6613 3.6613 2.75 3.35 2.75 4.0
1388 end StickyAtomTypes
1389 \end{lstlisting}
1390
1391 \section{\label{section::ffMetals}Metallic Atom Types}
1392
1393 {\sc OpenMD} implements a number of related potentials that describe
1394 bonding in transition metals. These potentials have an attractive
1395 interaction which models ``Embedding'' a positively charged
1396 pseudo-atom core in the electron density due to the free valance
1397 ``sea'' of electrons created by the surrounding atoms in the system.
1398 A pairwise part of the potential (which is primarily repulsive)
1399 describes the interaction of the positively charged metal core ions
1400 with one another. These potentials have the form:
1401 \begin{equation}
1402 V = \sum_{i} F_{i}\left[\rho_{i}\right] + \sum_{i} \sum_{j \neq i}
1403 \phi_{ij}({\bf r}_{ij})
1404 \end{equation}
1405 where $F_{i} $ is an embedding functional that approximates the energy
1406 required to embed a positively-charged core ion $i$ into a linear
1407 superposition of spherically averaged atomic electron densities given
1408 by $\rho_{i}$,
1409 \begin{equation}
1410 \rho_{i} = \sum_{j \neq i} f_{j}({\bf r}_{ij}),
1411 \end{equation}
1412 Since the density at site $i$ ($\rho_i$) must be computed before the
1413 embedding functional can be evaluated, {\sc eam} and the related
1414 transition metal potentials require two loops through the atom pairs
1415 to compute the inter-atomic forces.
1416
1417 The pairwise portion of the potential, $\phi_{ij}$, is usually a
1418 repulsive interaction between atoms $i$ and $j$.
1419
1420 \subsection{\label{section:ffEAM}The EAMAtomTypes block}
1421 The Embedded Atom Method ({\sc eam}) is one of the most widely-used
1422 potentials for transition
1423 metals.~\cite{Finnis84,Ercolessi88,Chen90,Qi99,Ercolessi02,Daw84,FBD86,johnson89,Lu97}
1424 It has been widely adopted in the materials science community and a
1425 good review of {\sc eam} and other formulations of metallic potentials
1426 was given by Voter.\cite{Voter:95}
1427
1428 In the original formulation of {\sc eam}\cite{Daw84}, the pair
1429 potential, $\phi_{ij}$ was an entirely repulsive term; however later
1430 refinements to {\sc eam} allowed for more general forms for
1431 $\phi$.\cite{Daw89} The effective cutoff distance, $r_{{\text cut}}$
1432 is the distance at which the values of $f(r)$ and $\phi(r)$ drop to
1433 zero for all atoms present in the simulation. In practice, this
1434 distance is fairly small, limiting the summations in the {\sc eam}
1435 equation to the few dozen atoms surrounding atom $i$ for both the
1436 density $\rho$ and pairwise $\phi$ interactions.
1437
1438 In computing forces for alloys, OpenMD uses mixing rules outlined by
1439 Johnson~\cite{johnson89} to compute the heterogenous pair potential,
1440 \begin{equation}
1441 \label{eq:johnson}
1442 \phi_{ab}(r)=\frac{1}{2}\left(
1443 \frac{f_{b}(r)}{f_{a}(r)}\phi_{aa}(r)+
1444 \frac{f_{a}(r)}{f_{b}(r)}\phi_{bb}(r)
1445 \right).
1446 \end{equation}
1447 No mixing rule is needed for the densities, since the density at site
1448 $i$ is simply the linear sum of density contributions of all the other
1449 atoms.
1450
1451 The {\sc eam} force field illustrates an additional feature of {\sc
1452 OpenMD}. Foiles, Baskes and Daw fit {\sc eam} potentials for Cu, Ag,
1453 Au, Ni, Pd, Pt and alloys of these metals.\cite{FBD86} These fits are
1454 included in {\sc OpenMD} as the {\tt u3} variant of the {\sc eam} force
1455 field. Voter and Chen reparamaterized a set of {\sc eam} functions
1456 which do a better job of predicting melting points.\cite{Voter:87}
1457 These functions are included in {\sc OpenMD} as the {\tt VC} variant of
1458 the {\sc eam} force field. An additional set of functions (the
1459 ``Universal 6'' functions) are included in {\sc OpenMD} as the {\tt u6}
1460 variant of {\sc eam}. For example, to specify the Voter-Chen variant
1461 of the {\sc eam} force field, the user would add the {\tt
1462 forceFieldVariant = "VC";} line to the meta-data file.
1463
1464 The potential files used by the {\sc eam} force field are in the
1465 standard {\tt funcfl} format, which is the format utilized by a number
1466 of other codes (e.g. ParaDyn~\cite{Paradyn}, {\sc dynamo 86}). It
1467 should be noted that the energy units in these files are in eV, not
1468 $\mbox{kcal mol}^{-1}$ as in the rest of the {\sc OpenMD} force field
1469 files.
1470
1471 \begin{lstlisting}[caption={[An example of a EAMAtomTypes block.] A
1472 simple example of a EAMAtomTypes block. Here the only data provided is
1473 the name of a {\tt funcfl} file which contains the raw data for spline
1474 interpolations for the density, functional, and pair potential.},
1475 label={sch:EAMAtomTypes}]
1476 begin EAMAtomTypes
1477 Au Au.u3.funcfl
1478 Ag Ag.u3.funcfl
1479 Cu Cu.u3.funcfl
1480 Ni Ni.u3.funcfl
1481 Pd Pd.u3.funcfl
1482 Pt Pt.u3.funcfl
1483 end EAMAtomTypes
1484 \end{lstlisting}
1485
1486 \subsection{\label{section:ffSC}The SuttonChenAtomTypes block}
1487
1488 The Sutton-Chen ({\sc sc})~\cite{Chen90} potential has been used to
1489 study a wide range of phenomena in metals. Although it has the same
1490 basic form as the {\sc eam} potential, the Sutton-Chen model requires
1491 a simpler set of parameters,
1492 \begin{equation}
1493 \label{eq:SCP1}
1494 U_{tot}=\sum _{i}\left[ \frac{1}{2}\sum _{j\neq
1495 i}\epsilon_{ij}V^{pair}_{ij}(r_{ij})-c_{i}\epsilon_{ii}\sqrt{\rho_{i}}\right] ,
1496 \end{equation}
1497 where $V^{pair}_{ij}$ and $\rho_{i}$ are given by
1498 \begin{equation}
1499 \label{eq:SCP2}
1500 V^{pair}_{ij}(r)=\left(
1501 \frac{\alpha_{ij}}{r_{ij}}\right)^{n_{ij}} \hspace{1in} \rho_{i}=\sum_{j\neq i}\left(
1502 \frac{\alpha_{ij}}{r_{ij}}\right) ^{m_{ij}}
1503 \end{equation}
1504
1505 $V^{pair}_{ij}$ is a repulsive pairwise potential that accounts for
1506 interactions of the pseudo-atom cores. The $\sqrt{\rho_i}$ term in
1507 Eq. (\ref{eq:SCP1}) is an attractive many-body potential that models
1508 the interactions between the valence electrons and the cores of the
1509 pseudo-atoms. $\epsilon_{ij}$, $\epsilon_{ii}$, $c_i$ and
1510 $\alpha_{ij}$ are parameters used to tune the potential for different
1511 transition metals.
1512
1513 The {\sc sc} potential form has also been parameterized by Qi {\it et
1514 al.}\cite{Qi99} These parameters were obtained via empirical and {\it
1515 ab initio} calculations to match structural features of the FCC
1516 crystal. Interested readers are encouraged to consult reference
1517 \citealp{Qi99} for further details.
1518
1519 \begin{lstlisting}[caption={[An example of a SCAtomTypes block.] A
1520 simple example of a SCAtomTypes block. Distances ($\alpha$)
1521 are given in \AA\ and energies ($\epsilon$) are (by convention) given in
1522 units of eV. These units must be specified in the {\tt Options} block
1523 using the keyword {\tt MetallicEnergyUnitScaling}. Without this {\tt
1524 Options} keyword, the default units for $\epsilon$ are kcal/mol. The
1525 other parameters, $m$, $n$, and $c$ are unitless.},
1526 label={sch:SCAtomTypes}]
1527 begin SCAtomTypes
1528 // Name epsilon(eV) c m n alpha(angstroms)
1529 Ni 0.0073767 84.745 5.0 10.0 3.5157
1530 Cu 0.0057921 84.843 5.0 10.0 3.6030
1531 Rh 0.0024612 305.499 5.0 13.0 3.7984
1532 Pd 0.0032864 148.205 6.0 12.0 3.8813
1533 Ag 0.0039450 96.524 6.0 11.0 4.0691
1534 Ir 0.0037674 224.815 6.0 13.0 3.8344
1535 Pt 0.0097894 71.336 7.0 11.0 3.9163
1536 Au 0.0078052 53.581 8.0 11.0 4.0651
1537 Au2 0.0078052 53.581 8.0 11.0 4.0651
1538 end SCAtomTypes
1539 \end{lstlisting}
1540
1541 \section{\label{section::ffShortRange}Short Range Interactions}
1542 The internal structure of a molecule is usually specified in terms of
1543 a set of ``bonded'' terms in the potential energy function for
1544 molecule $I$,
1545 \begin{align*}
1546 V^{I}_{\text{Internal}} = &
1547 \sum_{r_{ij} \in I} V_{\text{bond}}(r_{ij})
1548 + \sum_{\theta_{ijk} \in I} V_{\text{bend}}(\theta_{ijk})
1549 + \sum_{\phi_{ijkl} \in I} V_{\text{torsion}}(\phi_{ijkl})
1550 + \sum_{\omega_{ijkl} \in I} V_{\text{inversion}}(\omega_{ijkl}) \\
1551 & + \sum_{i \in I} \sum_{(j>i+4) \in I}
1552 \biggl[ V_{\text{dispersion}}(r_{ij}) + V_{\text{electrostatic}}
1553 (\mathbf{r}_{ij},\boldsymbol{\Omega}_{i},\boldsymbol{\Omega}_{j})
1554 \biggr].
1555 \label{eq:internalPotential}
1556 \end{align*}
1557 Here $V_{\text{bond}}, V_{\text{bend}},
1558 V_{\text{torsion}},\mathrm{~and~} V_{\text{inversion}}$ represent the
1559 bond, bend, torsion, and inversion potentials for all
1560 topologically-connected sets of atoms within the molecule. Bonds are
1561 the primary way of specifying how the atoms are connected together to
1562 form the molecule (i.e. they define the molecular topology). The
1563 other short-range interactions may be specified explicitly in the
1564 molecule definition, or they may be deduced from bonding information.
1565 For example, bends can be implicitly deduced from two bonds which
1566 share an atom. Torsions can be deduced from two bends that share a
1567 bond. Inversion potentials are utilized primarily to enforce
1568 planarity around $sp^2$-hybridized sites, and these are specified with
1569 central atoms and satellites (or an atom with bonds to exactly three
1570 satellites). Non-bonded interactions are usually excluded for atom
1571 pairs that are involved in the same bond, bend, or torsion, but all
1572 other atom pairs are included in the standard non-bonded interactions.
1573
1574 Bond lengths, angles, and torsions (dihedrals) are ``natural''
1575 coordinates to treat molecular motion, as it is usually in these
1576 coordinates that most chemists understand the behavior of molecules.
1577 The bond lengths and angles are often considered ``hard'' degrees of
1578 freedom. That is, we can't deform them very much without a
1579 significant energetic penalty. On the other hand, dihedral angles or
1580 torsions are ``soft'' and typically undergo significant deformation
1581 under normal conditions.
1582
1583 \subsection{\label{section:ffBond}The BondTypes block}
1584
1585 Bonds are the primary way to specify how the atoms are connected
1586 together to form the molecule. In general, bonds exert strong
1587 restoring forces to keep the molecule compact. The bond energy
1588 functions are relatively simple functions of the distance between two
1589 atomic sites,
1590 \begin{equation}
1591 b = \left| \vec{r}_{ij} \right| = \sqrt{(x_j - x_i)^2 + (y_j - y_i)^2
1592 + (z_j - z_i)^2}.
1593 \end{equation}
1594 All BondTypes must specify two AtomType names ($i$ and $j$) that
1595 describe when that bond should be applied, as well as an equilibrium
1596 bond length, $b_{ij}^0$, in units of \AA. The most common forms for
1597 bonding potentials are {\tt Harmonic} bonds,
1598 \begin{equation}
1599 V_{\text{bond}}(b) = \frac{k_{ij}}{2} \left(b -
1600 b_{ij}^0 \right)^2
1601 \end{equation}
1602 and {\tt Morse} bonds,
1603 \begin{equation}
1604 V_{\text{bond}}(b) = D_{ij} \left[ 1 - e^{-\beta_{ij} (b - b_{ij}^0)} \right]^2
1605 \end{equation}
1606
1607 \begin{figure}[h]
1608 \centering
1609 \includegraphics[width=2.5in]{bond.pdf}
1610 \caption[Bond coordinates]{The coordinate describing a
1611 a bond between atoms $i$ and $j$ is $|r_{ij}|$, the length of the
1612 $\vec{r}_{ij}$ vector. }
1613 \label{fig:bond}
1614 \end{figure}
1615
1616 OpenMD can also simulate some less common types of bond potentials,
1617 including {\tt Fixed} bonds (which are constrained to be at a
1618 specified bond length),
1619 \begin{equation}
1620 V_{\text{bond}}(b) = 0.
1621 \end{equation}
1622 {\tt Cubic} bonds include terms to model anharmonicity,
1623 \begin{equation}
1624 V_{\text{bond}}(b) = K_3 (b - b_{ij}^0)^3 + K_2 (b - b_{ij}^0)^2 + K_1 (b - b_{ij}^0) + K_0,
1625 \end{equation}
1626 and {\tt Quartic} bonds, include another term in the Taylor
1627 expansion around $b_{ij}^0$,
1628 \begin{equation}
1629 V_{\text{bond}}(b) = K_4 (b - b_{ij}^0)^4 + K_3 (b - b_{ij}^0)^3 +
1630 K_2 (b - b_{ij}^0)^2 + K_1 (b - b_{ij}^0) + K_0,
1631 \end{equation}
1632 can also be simulated. Note that the factor of $1/2$ that is included
1633 in the {\tt Harmonic} bond type force constant is {\it not} present in
1634 either the {\tt Cubic} or {\tt Quartic} bond types.
1635
1636 {\tt Polynomial} bonds which can have any number of terms,
1637 \begin{equation}
1638 V_{\text{bond}}(b) = \sum_n K_n (b - b_{ij}^0)^n.
1639 \end{equation}
1640 can also be specified by giving a sequence of integer ($n$) and force
1641 constant ($K_n$) pairs.
1642
1643 The order of terms in the BondTypes block is:
1644 \begin{itemize}
1645 \item {\tt AtomType} 1
1646 \item {\tt AtomType} 2
1647 \item {\tt BondType} (one of {\tt Harmonic}, {\tt Morse}, {\tt Fixed}, {\tt
1648 Cubic}, {\tt Quartic}, or {\tt Polynomial})
1649 \item $b_{ij}^0$, the equilibrium bond length in \AA
1650 \item any other parameters required by the {\tt BondType}
1651 \end{itemize}
1652
1653 \begin{lstlisting}[caption={[An example of a BondTypes block.] A
1654 simple example of a BondTypes block. Distances ($b_0$)
1655 are given in \AA\ and force constants are given in
1656 units so that when multiplied by the correct power of distance they
1657 return energies in kcal/mol. For example $k$ for a Harmonic bond is
1658 in units of kcal/mol/\AA$^2$.},
1659 label={sch:BondTypes}]
1660 begin BondTypes
1661 //Atom1 Atom2 Harmonic b0 k (kcal/mol/A^2)
1662 CH2 CH2 Harmonic 1.526 260
1663 //Atom1 Atom2 Morse b0 D beta (A^-1)
1664 CN NC Morse 1.157437 212.95 2.5802
1665 //Atom1 Atom2 Fixed b0
1666 CT HC Fixed 1.09
1667 //Atom1 Atom2 Cubic b0 K3 K2 K1 K0
1668 //Atom1 Atom2 Quartic b0 K4 K3 K2 K1 K0
1669 //Atom1 Atom2 Polynomial b0 n Kn [m Km]
1670 end BondTypes
1671 \end{lstlisting}
1672
1673 There are advantages and disadvantages of all of the different types
1674 of bonds, but specific simulation tasks may call for specific
1675 behaviors.
1676
1677 \subsection{\label{section:ffBend}The BendTypes block}
1678 The equilibrium geometries and energy functions for bending motions in
1679 a molecule are strongly dependent on the bonding environment of the
1680 central atomic site. For example, different types of hybridized
1681 carbon centers require different bending angles and force constants to
1682 describe the local geometry.
1683
1684 The bending potential energy functions used in most force fields are
1685 often simple functions of the angle between two bonds,
1686 \begin{equation}
1687 \theta_{ijk} = \cos^{-1} \left(\frac{\vec{r}_{ji} \cdot
1688 \vec{r}_{jk}}{\left| \vec{r}_{ji} \right| \left| \vec{r}_{jk}
1689 \right|} \right)
1690 \end{equation}
1691 Here atom $j$ is the central atom that is bonded to two partners $i$
1692 and $k$.
1693
1694 \begin{figure}[h]
1695 \centering
1696 \includegraphics[width=3.5in]{bend.pdf}
1697 \caption[Bend angle coordinates]{The coordinate describing a bend
1698 between atoms $i$, $j$, and $k$ is the angle $\theta_{ijk} =
1699 \cos^{-1} \left(\hat{r}_{ji} \cdot \hat{r}_{jk}\right)$ where $\hat{r}_{ji}$ is
1700 the unit vector between atoms $j$ and $i$. }
1701 \label{fig:bend}
1702 \end{figure}
1703
1704
1705 All BendTypes must specify three AtomType names ($i$, $j$ and $k$)
1706 that describe when that bend potential should be applied, as well as
1707 an equilibrium bending angle, $\theta_{ijk}^0$, in units of
1708 degrees. The most common forms for bending potentials are {\tt
1709 Harmonic} bends,
1710 \begin{equation}
1711 V_{\text{bend}}(\theta_{ijk}) = \frac{k_{ijk}}{2}( \theta_{ijk} - \theta_{ijk}^0
1712 )^2, \label{eq:bendPot}
1713 \end{equation}
1714 where $k_{ijk}$ is the force constant which determines the strength of
1715 the harmonic bend. {\tt UreyBradley} bends utilize an additional 1-3
1716 bond-type interaction in addition to the harmonic bending potential:
1717 \begin{equation}
1718 V_{\text{bend}}(\vec{r}_i , \vec{r}_j, \vec{r}_k)
1719 =\frac{k_{ijk}}{2}( \theta_{ijk} - \theta_{ijk}^0)^2
1720 + \frac{k_{ub}}{2}( r_{ik} - s_0 )^2. \label{eq:ubBend}
1721 \end{equation}
1722
1723 A {\tt Cosine} bend is a variant on the harmonic bend which utilizes
1724 the cosine of the angle instead of the angle itself,
1725 \begin{equation}
1726 V_{\text{bend}}(\theta_{ijk}) = \frac{k_{ijk}}{2}( \cos\theta_{ijk} -
1727 \cos \theta_{ijk}^0 )^2. \label{eq:cosBend}
1728 \end{equation}
1729
1730 OpenMD can also simulate some less common types of bend potentials,
1731 including {\tt Cubic} bends, which include terms to model anharmonicity,
1732 \begin{equation}
1733 V_{\text{bend}}(\theta_{ijk}) = K_3 (\theta_{ijk} - \theta_{ijk}^0)^3 + K_2 (\theta_{ijk} - \theta_{ijk}^0)^2 + K_1 (\theta_{ijk} - \theta_{ijk}^0) + K_0,
1734 \end{equation}
1735 and {\tt Quartic} bends, which include another term in the Taylor
1736 expansion around $\theta_{ijk}^0$,
1737 \begin{equation}
1738 V_{\text{bend}}(\theta_{ijk}) = K_4 (\theta_{ijk} - \theta_{ijk}^0)^4 + K_3 (\theta_{ijk} - \theta_{ijk}^0)^3 +
1739 K_2 (\theta_{ijk} - \theta_{ijk}^0)^2 + K_1 (\theta_{ijk} -
1740 \theta_{ijk}^0) + K_0,
1741 \end{equation}
1742 can also be simulated. Note that the factor of $1/2$ that is included
1743 in the {\tt Harmonic} bend type force constant is {\it not} present in
1744 either the {\tt Cubic} or {\tt Quartic} bend types.
1745
1746 {\tt Polynomial} bends which can have any number of terms,
1747 \begin{equation}
1748 V_{\text{bend}}(\theta_{ijk}) = \sum_n K_n (\theta_{ijk} - \theta_{ijk}^0)^n.
1749 \end{equation}
1750 can also be specified by giving a sequence of integer ($n$) and force
1751 constant ($K_n$) pairs.
1752
1753 The order of terms in the BendTypes block is:
1754 \begin{itemize}
1755 \item {\tt AtomType} 1
1756 \item {\tt AtomType} 2 (this is the central atom)
1757 \item {\tt AtomType} 3
1758 \item {\tt BendType} (one of {\tt Harmonic}, {\tt UreyBradley}, {\tt
1759 Cosine}, {\tt Cubic}, {\tt Quartic}, or {\tt Polynomial})
1760 \item $\theta_{ijk}^0$, the equilibrium bending angle in degrees.
1761 \item any other parameters required by the {\tt BendType}
1762 \end{itemize}
1763
1764 \begin{lstlisting}[caption={[An example of a BendTypes block.] A
1765 simple example of a BendTypes block. By convention, equilibrium angles
1766 ($\theta_0$) are given in degrees but force constants are given in
1767 units so that when multiplied by the correct power of angle (in
1768 radians) they return energies in kcal/mol. For example $k$ for a
1769 Harmonic bend is in units of kcal/mol/radians$^2$.},
1770 label={sch:BendTypes}]
1771 begin BendTypes
1772 //Atom1 Atom2 Atom3 Harmonic theta0(deg) Ktheta(kcal/mol/radians^2)
1773 CT CT CT Harmonic 109.5 80.000000
1774 CH2 CH CH2 Harmonic 112.0 117.68
1775 CH3 CH2 SH Harmonic 96.0 67.220
1776 //UreyBradley
1777 //Atom1 Atom2 Atom3 UreyBradley theta0 Ktheta s0 Kub
1778 //Cosine
1779 //Atom1 Atom2 Atom3 Cosine theta0 Ktheta(kcal/mol)
1780 //Cubic
1781 //Atom1 Atom2 Atom3 Cubic theta0 K3 K2 K1 K0
1782 //Quartic
1783 //Atom1 Atom2 Atom3 Quartic theta0 K4 K3 K2 K1 K0
1784 //Polynomial
1785 //Atom1 Atom2 Atom3 Polynomial theta0 n Kn [m Km]
1786 end BendTypes
1787 \end{lstlisting}
1788
1789 Note that the parameters for a particular bend type are the same for
1790 any bending triplet of the same atomic types (in the same or reversed
1791 order). Changing the AtomType in the Atom2 position will change the
1792 matched bend types in the force field.
1793
1794 \subsection{\label{section:ffTorsion}The TorsionTypes block}
1795 The torsion potential for rotation of groups around a central bond can
1796 often be represented with various cosine functions. For two
1797 tetrahedral ($sp^3$) carbons connected by a single bond, the torsion
1798 potential might be
1799 \begin{equation*}
1800 V_{\text{torsion}} = \frac{v}{2} \left[ 1 + \cos( 3 \phi ) \right]
1801 \end{equation*}
1802 where $v$ is the barrier for going from {\em staggered} $\rightarrow$
1803 {\em eclipsed} conformations, while for $sp^2$ carbons connected by a
1804 double bond, the potential might be
1805 \begin{equation*}
1806 V_{\text{torsion}} = \frac{w}{2} \left[ 1 - \cos( 2 \phi ) \right]
1807 \end{equation*}
1808 where $w$ is the barrier for going from {\em cis} $\rightarrow$ {\em
1809 trans} conformations.
1810
1811 A general torsion potentials can be represented as a cosine series of
1812 the form:
1813 \begin{equation}
1814 V_{\text{torsion}}(\phi_{ijkl}) = c_1[1 + \cos \phi_{ijkl}]
1815 + c_2[1 - \cos(2\phi_{ijkl})]
1816 + c_3[1 + \cos(3\phi_{ijkl})],
1817 \label{eq:origTorsionPot}
1818 \end{equation}
1819 where the angle $\phi_{ijkl}$ is defined
1820 \begin{equation}
1821 \cos\phi_{ijkl} = (\hat{\mathbf{r}}_{ij} \times \hat{\mathbf{r}}_{jk}) \cdot
1822 (\hat{\mathbf{r}}_{jk} \times \hat{\mathbf{r}}_{kl}).
1823 \label{eq:torsPhi}
1824 \end{equation}
1825 Here, $\hat{\mathbf{r}}_{\alpha\beta}$ are the set of unit bond
1826 vectors between atoms $i$, $j$, $k$, and $l$. Note that some force
1827 fields define the zero of the $\phi_{ijkl}$ angle when atoms $i$ and
1828 $l$ are in the {\em trans} configuration, while most define the zero
1829 angle for when $i$ and $l$ are in the fully eclipsed orientation. The
1830 behavior of the torsion parser can be altered with the {\tt
1831 TorsionAngleConvention} keyword in the Options block. The default
1832 behavior is {\tt "180\_is\_trans"} while the opposite behavior can be
1833 invoked by setting this keyword to {\tt "0\_is\_trans"}.
1834
1835 \begin{figure}[h]
1836 \centering
1837 \includegraphics[width=4.5in]{torsion.pdf}
1838 \caption[Torsion or dihedral angle coordinates]{The coordinate
1839 describing a torsion between atoms $i$, $j$, $k$, and $l$ is the
1840 dihedral angle $\phi_{ijkl}$ which measures the relative rotation of
1841 the two terminal atoms around the axis defined by the middle bond. }
1842 \label{fig:torsion}
1843 \end{figure}
1844
1845 For computational efficiency, OpenMD recasts torsion potential in the
1846 method of {\sc charmm},\cite{Brooks83} in which the angle series is
1847 converted to a power series of the form:
1848 \begin{equation}
1849 V_{\text{torsion}}(\phi_{ijkl}) =
1850 k_3 \cos^3 \phi_{ijkl} + k_2 \cos^2 \phi_{ijkl} + k_1 \cos \phi_{ijkl} + k_0,
1851 \label{eq:torsionPot}
1852 \end{equation}
1853 where:
1854 \begin{align*}
1855 k_0 &= c_1 + 2 c_2 + c_3, \\
1856 k_1 &= c_1 - 3c_3, \\
1857 k_2 &= - 2 c_2, \\
1858 k_3 &= 4 c_3.
1859 \end{align*}
1860 By recasting the potential as a power series, repeated trigonometric
1861 evaluations are avoided during the calculation of the potential
1862 energy.
1863
1864 Using this framework, OpenMD implements a variety of different
1865 potential energy functions for torsions:
1866 \begin{itemize}
1867 \item {\tt Cubic}:
1868 \begin{equation*}
1869 V_{\text{torsion}}(\phi) =
1870 k_3 \cos^3 \phi + k_2 \cos^2 \phi + k_1 \cos \phi + k_0,
1871 \end{equation*}
1872 \item {\tt Quartic}:
1873 \begin{equation*}
1874 V_{\text{torsion}}(\phi) = k_4 \cos^4 \phi +
1875 k_3 \cos^3 \phi + k_2 \cos^2 \phi + k_1 \cos \phi + k_0,
1876 \end{equation*}
1877 \item {\tt Polynomial}:
1878 \begin{equation*}
1879 V_{\text{torsion}}(\phi) = \sum_n k_n \cos^n \phi ,
1880 \end{equation*}
1881 \item {\tt Charmm}:
1882 \begin{equation*}
1883 V_{\text{torsion}}(\phi) = \sum_n K_n \left( 1 + cos(n
1884 \phi - \delta_n) \right),
1885 \end{equation*}
1886 \item {\tt Opls}:
1887 \begin{equation*}
1888 V_{\text{torsion}}(\phi) = \frac{1}{2} \left(v_1 (1 + \cos \phi) \right)
1889 + v_2 (1 - \cos 2 \phi) + v_3 (1 + \cos 3 \phi),
1890 \end{equation*}
1891 \item {\tt Trappe}:\cite{Siepmann1998}
1892 \begin{equation*}
1893 V_{\text{torsion}}(\phi) = c_0 + c_1 (1 + \cos \phi) + c_2 (1 - \cos 2 \phi) +
1894 c_3 (1 + \cos 3 \phi),
1895 \end{equation*}
1896 \item {\tt Harmonic}:
1897 \begin{equation*}
1898 V_{\text{torsion}}(\phi) = \frac{d_0}{2} \left(\phi - \phi^0\right).
1899 \end{equation*}
1900 \end{itemize}
1901
1902 Most torsion types don't require specific angle information in the
1903 parameters since they are typically expressed in cosine polynomials.
1904 {\tt Charmm} and {\tt Harmonic} torsions are a bit different. {\tt
1905 Charmm} torsion types require a set of phase angles, $\delta_n$ that
1906 are expressed in degrees, and associated periodicity numbers, $n$.
1907 {\tt Harmonic} torsions have an equilibrium torsion angle, $\phi_0$
1908 that is measured in degrees, while $d_0$ has units of
1909 kcal/mol/degrees$^2$. All other torsion parameters are measured in
1910 units of kcal/mol.
1911
1912 \begin{lstlisting}[caption={[An example of a TorsionTypes block.] A
1913 simple example of a TorsionTypes block. Energy constants are given in
1914 kcal / mol, and when required by the form, $\delta$ angles are given
1915 in degrees.},
1916 label={sch:TorsionTypes}]
1917 begin TorsionTypes
1918 //Cubic
1919 //Atom1 Atom2 Atom3 Atom4 Cubic k3 k2 k1 k0
1920 CH2 CH2 CH2 CH2 Cubic 5.9602 -0.2568 -3.802 2.1586
1921 CH2 CH CH CH2 Cubic 3.3254 -0.4215 -1.686 1.1661
1922 //Trappe
1923 //Atom1 Atom2 Atom3 Atom4 Trappe c0 c1 c2 c3
1924 CH3 CH2 CH2 SH Trappe 0.10507 -0.10342 0.03668 0.60874
1925 //Charmm
1926 //Atom1 Atom2 Atom3 Atom4 Charmm Kchi n delta [Kchi n delta]
1927 CT CT CT C Charmm 0.156 3 0.00
1928 OH CT CT OH Charmm 0.144 3 0.00 1.175 2 0
1929 HC CT CM CM Charmm 1.150 1 0.00 0.38 3 180
1930 //Quartic
1931 //Atom1 Atom2 Atom3 Atom4 Quartic k4 k3 k2 k1 k0
1932 //Polynomial
1933 //Atom1 Atom2 Atom3 Atom4 Polynomial n Kn [m Km]
1934 S CH2 CH2 C Polynomial 0 2.218 1 2.905 2 -3.136 3 -0.7313 4 6.272 5 -7.528
1935 end TorsionTypes
1936 \end{lstlisting}
1937
1938 Note that the parameters for a particular torsion type are the same
1939 for any torsional quartet of the same atomic types (in the same or
1940 reversed order).
1941
1942 \subsection{\label{section:ffInversion}The InversionTypes block}
1943
1944 Inversion potentials are often added to force fields to enforce
1945 planarity around $sp^2$-hybridized carbons or to correct vibrational
1946 frequencies for umbrella-like vibrational modes for central atoms
1947 bonded to exactly three satellite groups.
1948
1949 In OpenMD's version of an inversion, the central atom is entered {\it
1950 first} in each line in the {\tt InversionTypes} block. Note that
1951 this is quite different than how other codes treat Improper torsional
1952 potentials to mimic inversion behavior. In some other widely-used
1953 simulation packages, the central atom is treated as atom 3 in a
1954 standard torsion form:
1955 \begin{itemize}
1956 \item OpenMD: I - (J - K - L) (e.g. I is $sp^2$ hybridized carbon)
1957 \item AMBER: I - J - K - L (e.g. K is $sp^2$ hybridized carbon)
1958 \end{itemize}
1959
1960 The inversion angle itself is defined as:
1961 \begin{equation}
1962 \cos\omega_{i-jkl} = \left(\hat{\mathbf{r}}_{il} \times
1963 \hat{\mathbf{r}}_{ij}\right)\cdot\left( \hat{\mathbf{r}}_{il} \times
1964 \hat{\mathbf{r}}_{ik}\right)
1965 \end{equation}
1966 Here, $\hat{\mathbf{r}}_{\alpha\beta}$ are the set of unit bond
1967 vectors between the central atoms $i$, and the satellite atoms $j$,
1968 $k$, and $l$. Note that other definitions of inversion angles are
1969 possible, so users are encouraged to be particularly careful when
1970 converting other force field files for use with OpenMD.
1971
1972 There are many common ways to create planarity or umbrella behavior in
1973 a potential energy function, and OpenMD implements a number of the
1974 more common functions:
1975 \begin{itemize}
1976 \item {\tt ImproperCosine}:
1977 \begin{equation*}
1978 V_{\text{torsion}}(\omega) = \sum_n \frac{K_n}{2} \left( 1 + cos(n
1979 \omega - \delta_n) \right),
1980 \end{equation*}
1981 \item {\tt AmberImproper}:
1982 \begin{equation*}
1983 V_{\text{torsion}}(\omega) = \frac{v}{2} (1 - \cos\left(2 \left(\omega - \omega_0\right)\right),
1984 \end{equation*}
1985 \item {\tt Harmonic}:
1986 \begin{equation*}
1987 V_{\text{torsion}}(\omega) = \frac{d}{2} \left(\omega - \omega_0\right).
1988 \end{equation*}
1989 \end{itemize}
1990 \begin{lstlisting}[caption={[An example of an InversionTypes block.] A
1991 simple example of a InversionTypes block. Angles ($\delta_n$ and
1992 $\omega_0$) angles are given in degrees, while energy parameters ($v,
1993 K_n$) are given in kcal / mol. The Harmonic Inversion type has a
1994 force constant that must be given in kcal/mol/degrees$^2$.},
1995 label={sch:InversionTypes}]
1996 begin InversionTypes
1997 //Harmonic
1998 //Atom1 Atom2 Atom3 Atom4 Harmonic d(kcal/mol/deg^2) omega0
1999 RCHar3 X X X Harmonic 1.21876e-2 180.0
2000 //AmberImproper
2001 //Atom1 Atom2 Atom3 Atom4 AmberImproper v(kcal/mol)
2002 C CT N O AmberImproper 10.500000
2003 CA CA CA CT AmberImproper 1.100000
2004 //ImproperCosine
2005 //Atom1 Atom2 Atom3 Atom4 ImproperCosine Kn n delta_n [Kn n delta_n]
2006 end InversionTypes
2007 \end{lstlisting}
2008
2009 \section{\label{section::ffLongRange}Long Range Interactions}
2010
2011 Calculating the long-range (non-bonded) potential involves a sum over
2012 all pairs of atoms (except for those atoms which are involved in a
2013 bond, bend, or torsion with each other). Many of these interactions
2014 can be inferred from the AtomTypes,
2015
2016 \subsection{\label{section:ffNBinteraction}The NonBondedInteractions
2017 block}
2018
2019 The user might want like to specify explicit or special interactions
2020 that override the default non-bodned interactions that are inferred
2021 from the AtomTypes. To do this, OpenMD implements a
2022 NonBondedInteractions block to allow the user to specify the following
2023 (pair-wise) non-bonded interactions:
2024
2025 \begin{itemize}
2026 \item {\tt LennardJones}:
2027 \begin{equation*}
2028 V_{\text{NB}}(r) = 4 \epsilon_{ij} \left(
2029 \left(\frac{\sigma_{ij}}{r} \right)^{12} -
2030 \left(\frac{\sigma_{ij}}{r} \right)^{6} \right),
2031 \end{equation*}
2032 \item {\tt ShiftedMorse}:
2033 \begin{equation*}
2034 V_{\text{NB}}(r) = D_{ij} \left( e^{-2 \beta_{ij} (r -
2035 r^0)} - 2 e^{- \beta_{ij} (r -
2036 r^0)} \right),
2037 \end{equation*}
2038 \item {\tt RepulsiveMorse}:
2039 \begin{equation*}
2040 V_{\text{NB}}(r) = D_{ij} \left( e^{-2 \beta_{ij} (r -
2041 r^0)} \right),
2042 \end{equation*}
2043 \item {\tt RepulsivePower}:
2044 \begin{equation*}
2045 V_{\text{NB}}(r) = \epsilon_{ij}
2046 \left(\frac{\sigma_{ij}}{r} \right)^{n_{ij}}.
2047 \end{equation*}
2048 \end{itemize}
2049
2050 \begin{lstlisting}[caption={[An example of a NonBondedInteractions block.] A
2051 simple example of a NonBondedInteractions block. Distances ($\sigma,
2052 r_0$) are given in \AA, while energies ($\epsilon, D0$) are in
2053 kcal/mol. The Morse potentials have an additional parameter $\beta_0$
2054 which is in units of \AA$^{-1}$.},
2055 label={sch:InversionTypes}]
2056 begin NonBondedInteractions
2057
2058 //Lennard-Jones
2059 //Atom1 Atom2 LennardJones sigma epsilon
2060 Au CH3 LennardJones 3.54 0.2146
2061 Au CH2 LennardJones 3.54 0.1749
2062 Au CH LennardJones 3.54 0.1749
2063 Au S LennardJones 2.40 8.465
2064
2065 //Shifted Morse
2066 //Atom1 Atom2 ShiftedMorse r0 D0 beta0
2067 Au O_SPCE ShiftedMorse 3.70 0.0424 0.769
2068
2069 //Repulsive Morse
2070 //Atom1 Atom2 RepulsiveMorse r0 D0 beta0
2071 Au H_SPCE RepulsiveMorse -1.00 0.00850 0.769
2072
2073 //Repulsive Power
2074 //Atom1 Atom2 RepulsivePower sigma epsilon n
2075 Au ON RepulsivePower 3.47005 0.186208 11
2076 Au NO RepulsivePower 3.53955 0.168629 11
2077 end NonBondedInteractions
2078 \end{lstlisting}
2079
2080 \section{\label{section:electrostatics}Electrostatics}
2081
2082 Because nearly all force fields involve electrostatic interactions in
2083 one form or another, OpenMD implements a number of different
2084 electrostatic summation methods. These methods are extended from the
2085 damped and cutoff-neutralized Coulombic sum originally proposed by
2086 Wolf, {\it et al.}\cite{Wolf99} One of these, the damped shifted force
2087 method, shows a remarkable ability to reproduce the energetic and
2088 dynamic characteristics exhibited by simulations employing lattice
2089 summation techniques. The basic idea is to construct well-behaved
2090 real-space summation methods using two tricks:
2091 \begin{enumerate}
2092 \item shifting through the use of image charges, and
2093 \item damping the electrostatic interaction.
2094 \end{enumerate}
2095 Starting with the original observation that the effective range of the
2096 electrostatic interaction in condensed phases is considerably less
2097 than $r^{-1}$, either the cutoff sphere neutralization or the
2098 distance-dependent damping technique could be used as a foundation for
2099 a new pairwise summation method. Wolf \textit{et al.} made the
2100 observation that charge neutralization within the cutoff sphere plays
2101 a significant role in energy convergence; therefore we will begin our
2102 analysis with the various shifted forms that maintain this charge
2103 neutralization. We can evaluate the methods of Wolf
2104 \textit{et al.} and Zahn \textit{et al.} by considering the standard
2105 shifted potential,
2106 \begin{equation}
2107 V_\textrm{SP}(r) = \begin{cases}
2108 v(r)-v_\textrm{c} &\quad r\leqslant R_\textrm{c} \\ 0 &\quad r >
2109 R_\textrm{c}
2110 \end{cases},
2111 \label{eq:shiftingPotForm}
2112 \end{equation}
2113 and shifted force,
2114 \begin{equation}
2115 V_\textrm{SF}(r) = \begin{cases}
2116 v(r)-v_\textrm{c}-\left(\frac{d v(r)}{d r}\right)_{r=R_\textrm{c}}(r-R_\textrm{c
2117 })
2118 &\quad r\leqslant R_\textrm{c} \\ 0 &\quad r > R_\textrm{c}
2119 \end{cases},
2120 \label{eq:shiftingForm}
2121 \end{equation}
2122 functions where $v(r)$ is the unshifted form of the potential, and
2123 $v_c$ is $v(R_\textrm{c})$. The Shifted Force ({\sc sf}) form ensures
2124 that both the potential and the forces goes to zero at the cutoff
2125 radius, while the Shifted Potential ({\sc sp}) form only ensures the
2126 potential is smooth at the cutoff radius
2127 ($R_\textrm{c}$).\cite{Allen87}
2128
2129 The forces associated with the shifted potential are simply the forces
2130 of the unshifted potential itself (when inside the cutoff sphere),
2131 \begin{equation}
2132 F_{\textrm{SP}} = -\left( \frac{d v(r)}{dr} \right),
2133 \end{equation}
2134 and are zero outside. Inside the cutoff sphere, the forces associated
2135 with the shifted force form can be written,
2136 \begin{equation}
2137 F_{\textrm{SF}} = -\left( \frac{d v(r)}{dr} \right) + \left(\frac{d
2138 v(r)}{dr} \right)_{r=R_\textrm{c}}.
2139 \end{equation}
2140
2141 If the potential, $v(r)$, is taken to be the normal Coulomb potential,
2142 \begin{equation}
2143 v(r) = \frac{q_i q_j}{r},
2144 \label{eq:Coulomb}
2145 \end{equation}
2146 then the Shifted Potential ({\sc sp}) forms will give Wolf {\it et
2147 al.}'s undamped prescription:
2148 \begin{equation}
2149 V_\textrm{SP}(r) =
2150 q_iq_j\left(\frac{1}{r}-\frac{1}{R_\textrm{c}}\right) \quad
2151 r\leqslant R_\textrm{c},
2152 \label{eq:SPPot}
2153 \end{equation}
2154 with associated forces,
2155 \begin{equation}
2156 F_\textrm{SP}(r) = q_iq_j\left(\frac{1}{r^2}\right) \quad r\leqslant R_\textrm{c
2157 }.
2158 \label{eq:SPForces}
2159 \end{equation}
2160 These forces are identical to the forces of the standard Coulomb
2161 interaction, and cutting these off at $R_c$ was addressed by Wolf
2162 \textit{et al.} as undesirable. They pointed out that the effect of
2163 the image charges is neglected in the forces when this form is
2164 used,\cite{Wolf99} thereby eliminating any benefit from the method in
2165 molecular dynamics. Additionally, there is a discontinuity in the
2166 forces at the cutoff radius which results in energy drift during MD
2167 simulations.
2168
2169 The shifted force ({\sc sf}) form using the normal Coulomb potential
2170 will give,
2171 \begin{equation}
2172 V_\textrm{SF}(r) = q_iq_j\left[\frac{1}{r}-\frac{1}{R_\textrm{c}}+\left(\frac{1}
2173 {R_\textrm{c}^2}\right)(r-R_\textrm{c})\right] \quad r\leqslant R_\textrm{c}.
2174 \label{eq:SFPot}
2175 \end{equation}
2176 with associated forces,
2177 \begin{equation}
2178 F_\textrm{SF}(r) = q_iq_j\left(\frac{1}{r^2}-\frac{1}{R_\textrm{c}^2}\right) \quad r\leqslant R_\textrm{c}.
2179 \label{eq:SFForces}
2180 \end{equation}
2181 This formulation has the benefits that there are no discontinuities at
2182 the cutoff radius, while the neutralizing image charges are present in
2183 both the energy and force expressions. It would be simple to add the
2184 self-neutralizing term back when computing the total energy of the
2185 system, thereby maintaining the agreement with the Madelung energies.
2186 A side effect of this treatment is the alteration in the shape of the
2187 potential that comes from the derivative term. Thus, a degree of
2188 clarity about agreement with the empirical potential is lost in order
2189 to gain functionality in dynamics simulations.
2190
2191 Wolf \textit{et al.} originally discussed the energetics of the
2192 shifted Coulomb potential (Eq. \ref{eq:SPPot}) and found that it was
2193 insufficient for accurate determination of the energy with reasonable
2194 cutoff distances. The calculated Madelung energies fluctuated around
2195 the expected value as the cutoff radius was increased, but the
2196 oscillations converged toward the correct value.\cite{Wolf99} A
2197 damping function was incorporated to accelerate the convergence; and
2198 though alternative forms for the damping function could be
2199 used,\cite{Jones56,Heyes81} the complimentary error function was
2200 chosen to mirror the effective screening used in the Ewald summation.
2201 Incorporating this error function damping into the simple Coulomb
2202 potential,
2203 \begin{equation}
2204 v(r) = \frac{\mathrm{erfc}\left(\alpha r\right)}{r},
2205 \label{eq:dampCoulomb}
2206 \end{equation}
2207 the shifted potential (eq. (\ref{eq:SPPot})) becomes
2208 \begin{equation}
2209 V_{\textrm{DSP}}(r) = q_iq_j\left(\frac{\textrm{erfc}\left(\alpha r\right)}{r}-\frac{\textrm{erfc}\left(\alpha R_\textrm{c}\right)}{R_\textrm{c}}\right) \quad r
2210 \leqslant R_\textrm{c},
2211 \label{eq:DSPPot}
2212 \end{equation}
2213 with associated forces,
2214 \begin{equation}
2215 F_{\textrm{DSP}}(r) = q_iq_j\left(\frac{\textrm{erfc}\left(\alpha r\right)}{r^2}
2216 +\frac{2\alpha}{\pi^{1/2}}\frac{\exp{\left(-\alpha^2r^2\right)}}{r}\right) \quad
2217 r\leqslant R_\textrm{c}.
2218 \label{eq:DSPForces}
2219 \end{equation}
2220 Again, this damped shifted potential suffers from a
2221 force-discontinuity at the cutoff radius, and the image charges play
2222 no role in the forces. To remedy these concerns, one may derive a
2223 {\sc sf} variant by including the derivative term in
2224 eq. (\ref{eq:shiftingForm}),
2225 \begin{equation}
2226 \begin{split}
2227 V_\mathrm{DSF}(r) = q_iq_j\Biggr{[} & \frac{\mathrm{erfc}\left(\alpha r \right)}{r} -\frac{\mathrm{erfc}\left(\alpha R_\mathrm{c} \right) }{R_\mathrm{c}} \\
2228 & \left. +\left(\frac{\mathrm{erfc}\left(\alpha
2229 R_\mathrm{c}\right)}{R_\mathrm{c}^2}+\frac{2\alpha}{\pi^{1/2}}\frac{\exp\left(-\alpha^2R_\mathrm{c}^2\right)}{R_\mathrm{c}}\right)\left(r-R_\mathrm{c}\right)
2230 \right] \quad r\leqslant R_\textrm{c}
2231 \label{eq:DSFPot}
2232 \end{split}
2233 \end{equation}
2234 The derivative of the above potential will lead to the following forces,
2235 \begin{equation}
2236 \begin{split}
2237 F_\mathrm{DSF}(r) =
2238 q_iq_j\Biggr{[}&\left(\frac{\textrm{erfc}\left(\alpha r\right)}{r^2}+\frac{2\alpha}{\pi^{1/2}}\frac{\exp{\left(-\alpha^2r^2\right)}}{r}\right) \\ &\left.-\left(\frac{\textrm{erfc}\left(\alpha R_{\textrm{c}}\right)}{R_{\textrm{c}}^2}+\frac{2\alpha}{\pi^{1/2}}\frac{\exp{\left(-\alpha^2R_{\textrm{c}}^2
2239 \right)}}{R_{\textrm{c}}}\right)\right] \quad r\leqslant R_\textrm{c}.
2240 \label{eq:DSFForces}
2241 \end{split}
2242 \end{equation}
2243 If the damping parameter $(\alpha)$ is set to zero, the undamped case,
2244 eqs. (\ref{eq:SPPot} through \ref{eq:SFForces}) are correctly
2245 recovered from eqs. (\ref{eq:DSPPot} through \ref{eq:DSFForces}).
2246
2247 It has been shown that the Damped Shifted Force method obtains nearly
2248 identical behavior to the smooth particle mesh Ewald ({\sc spme})
2249 method on a number of commonly simulated systems.\cite{Fennell06} For
2250 this reason, the default electrostatic summation method utilized by
2251 {\sc OpenMD} is the DSF (Eq. \ref{eq:DSFPot}) with a damping parameter
2252 ($\alpha$) that is set algorithmically from the cutoff radius.
2253
2254
2255 \section{\label{section:cutoffGroups}Switching Functions}
2256
2257 Calculating the the long-range interactions for $N$ atoms involves
2258 $N(N-1)/2$ evaluations of atomic distances if it is done in a brute
2259 force manner. To reduce the number of distance evaluations between
2260 pairs of atoms, {\sc OpenMD} allows the use of hard or switched
2261 cutoffs with Verlet neighbor lists.\cite{Allen87} Neutral groups which
2262 contain charges can exhibit pathological forces unless the cutoff is
2263 applied to the neutral groups evenly instead of to the individual
2264 atoms.\cite{leach01:mm} {\sc OpenMD} allows users to specify cutoff
2265 groups which may contain an arbitrary number of atoms in the molecule.
2266 Atoms in a cutoff group are treated as a single unit for the
2267 evaluation of the switching function:
2268 \begin{equation}
2269 V_{\mathrm{long-range}} = \sum_{a} \sum_{b>a} s(r_{ab}) \sum_{i \in a} \sum_{j \in b} V_{ij}(r_{ij}),
2270 \end{equation}
2271 where $r_{ab}$ is the distance between the centers of mass of the two
2272 cutoff groups ($a$ and $b$).
2273
2274 The sums over $a$ and $b$ are over the cutoff groups that are present
2275 in the simulation. Atoms which are not explicitly defined as members
2276 of a {\tt cutoffGroup} are treated as a group consisting of only one
2277 atom. The switching function, $s(r)$ is the standard cubic switching
2278 function,
2279 \begin{equation}
2280 S(r) =
2281 \begin{cases}
2282 1 & \text{if $r \le r_{\text{sw}}$},\\
2283 \frac{(r_{\text{cut}} + 2r - 3r_{\text{sw}})(r_{\text{cut}} - r)^2}
2284 {(r_{\text{cut}} - r_{\text{sw}})^3}
2285 & \text{if $r_{\text{sw}} < r \le r_{\text{cut}}$}, \\
2286 0 & \text{if $r > r_{\text{cut}}$.}
2287 \end{cases}
2288 \label{eq:dipoleSwitching}
2289 \end{equation}
2290 Here, $r_{\text{sw}}$ is the {\tt switchingRadius}, or the distance
2291 beyond which interactions are reduced, and $r_{\text{cut}}$ is the
2292 {\tt cutoffRadius}, or the distance at which interactions are
2293 truncated.
2294
2295 Users of {\sc OpenMD} do not need to specify the {\tt cutoffRadius} or
2296 {\tt switchingRadius}.
2297 If the {\tt cutoffRadius} was not explicitly set, OpenMD will attempt
2298 to guess an appropriate choice. If the system contains electrostatic
2299 atoms, the default cutoff radius is 12 \AA. In systems without
2300 electrostatic (charge or multipolar) atoms, the atom types present in the simulation will be
2301 polled for suggested cutoff values (e.g. $2.5 max(\left\{ \sigma
2302 \right\})$ for Lennard-Jones atoms. The largest suggested value
2303 that was found will be used.
2304
2305 By default, OpenMD uses shifted force potentials to force the
2306 potential energy and forces to smoothly approach zero at the cutoff
2307 radius. If the user would like to use another cutoff method
2308 they may do so by setting the {\tt cutoffMethod} parameter to:
2309 \begin{itemize}
2310 \item {\tt HARD}
2311 \item {\tt SWITCHED}
2312 \item {\tt SHIFTED\_FORCE} (default):
2313 \item {\tt TAYLOR\_SHIFTED}
2314 \item {\tt SHIFTED\_POTENTIAL}
2315 \end{itemize}
2316
2317 The {\tt switchingRadius} is set to a default value of 95\% of the
2318 {\tt cutoffRadius}. In the special case of a simulation containing
2319 {\it only} Lennard-Jones atoms, the default switching radius takes the
2320 same value as the cutoff radius, and {\sc OpenMD} will use a shifted
2321 potential to remove discontinuities in the potential at the cutoff.
2322 Both radii may be specified in the meta-data file.
2323
2324
2325 \section{\label{section:pbc}Periodic Boundary Conditions}
2326
2327 \newcommand{\roundme}{\operatorname{round}}
2328
2329 \textit{Periodic boundary conditions} are widely used to simulate bulk
2330 properties with a relatively small number of particles. In this method
2331 the simulation box is replicated throughout space to form an infinite
2332 lattice. During the simulation, when a particle moves in the primary
2333 cell, its image in other cells move in exactly the same direction with
2334 exactly the same orientation. Thus, as a particle leaves the primary
2335 cell, one of its images will enter through the opposite face. If the
2336 simulation box is large enough to avoid ``feeling'' the symmetries of
2337 the periodic lattice, surface effects can be ignored. The available
2338 periodic cells in {\sc OpenMD} are cubic, orthorhombic and
2339 parallelepiped. {\sc OpenMD} use a $3 \times 3$ matrix, $\mathsf{H}$,
2340 to describe the shape and size of the simulation box. $\mathsf{H}$ is
2341 defined:
2342 \begin{equation}
2343 \mathsf{H} = ( \mathbf{h}_x, \mathbf{h}_y, \mathbf{h}_z ),
2344 \end{equation}
2345 where $\mathbf{h}_{\alpha}$ is the column vector of the $\alpha$ axis of the
2346 box. During the course of the simulation both the size and shape of
2347 the box can be changed to allow volume fluctuations when constraining
2348 the pressure.
2349
2350 A real space vector, $\mathbf{r}$ can be transformed in to a box space
2351 vector, $\mathbf{s}$, and back through the following transformations:
2352 \begin{align}
2353 \mathbf{s} &= \mathsf{H}^{-1} \mathbf{r}, \\
2354 \mathbf{r} &= \mathsf{H} \mathbf{s}.
2355 \end{align}
2356 The vector $\mathbf{s}$ is now a vector expressed as the number of box
2357 lengths in the $\mathbf{h}_x$, $\mathbf{h}_y$, and $\mathbf{h}_z$
2358 directions. To find the minimum image of a vector $\mathbf{r}$, {\sc
2359 OpenMD} first converts it to its corresponding vector in box space, and
2360 then casts each element to lie in the range $[-0.5,0.5]$:
2361 \begin{equation}
2362 s_{i}^{\prime}=s_{i}-\roundme(s_{i}),
2363 \end{equation}
2364 where $s_i$ is the $i$th element of $\mathbf{s}$, and
2365 $\roundme(s_i)$ is given by
2366 \begin{equation}
2367 \roundme(x) =
2368 \begin{cases}
2369 \lfloor x+0.5 \rfloor & \text{if $x \ge 0$,} \\
2370 \lceil x-0.5 \rceil & \text{if $x < 0$.}
2371 \end{cases}
2372 \end{equation}
2373 Here $\lfloor x \rfloor$ is the floor operator, and gives the largest
2374 integer value that is not greater than $x$, and $\lceil x \rceil$ is
2375 the ceiling operator, and gives the smallest integer that is not less
2376 than $x$.
2377
2378 Finally, the minimum image coordinates $\mathbf{r}^{\prime}$ are
2379 obtained by transforming back to real space,
2380 \begin{equation}
2381 \mathbf{r}^{\prime}=\mathsf{H}^{-1}\mathbf{s}^{\prime}.%
2382 \end{equation}
2383 In this way, particles are allowed to diffuse freely in $\mathbf{r}$,
2384 but their minimum images, or $\mathbf{r}^{\prime}$, are used to compute
2385 the inter-atomic forces.
2386
2387 \chapter{\label{section:mechanics}Mechanics}
2388
2389 \section{\label{section:integrate}Integrating the Equations of Motion: the
2390 {\sc dlm} method}
2391
2392 The default method for integrating the equations of motion in {\sc
2393 OpenMD} is a velocity-Verlet version of the symplectic splitting method
2394 proposed by Dullweber, Leimkuhler and McLachlan
2395 ({\sc dlm}).\cite{Dullweber1997} When there are no directional atoms or
2396 rigid bodies present in the simulation, this integrator becomes the
2397 standard velocity-Verlet integrator which is known to sample the
2398 microcanonical (NVE) ensemble.\cite{Frenkel1996}
2399
2400 Previous integration methods for orientational motion have problems
2401 that are avoided in the {\sc dlm} method. Direct propagation of the Euler
2402 angles has a known $1/\sin\theta$ divergence in the equations of
2403 motion for $\phi$ and $\psi$,\cite{Allen87} leading to numerical
2404 instabilities any time one of the directional atoms or rigid bodies
2405 has an orientation near $\theta=0$ or $\theta=\pi$. Quaternion-based
2406 integration methods work well for propagating orientational motion;
2407 however, energy conservation concerns arise when using the
2408 microcanonical (NVE) ensemble. An earlier implementation of {\sc
2409 OpenMD} utilized quaternions for propagation of rotational motion;
2410 however, a detailed investigation showed that they resulted in a
2411 steady drift in the total energy, something that has been observed by
2412 Laird {\it et al.}\cite{Laird97}
2413
2414 The key difference in the integration method proposed by Dullweber
2415 \emph{et al.} is that the entire $3 \times 3$ rotation matrix is
2416 propagated from one time step to the next. In the past, this would not
2417 have been feasible, since the rotation matrix for a single body has
2418 nine elements compared with the more memory-efficient methods (using
2419 three Euler angles or 4 quaternions). Computer memory has become much
2420 less costly in recent years, and this can be translated into
2421 substantial benefits in energy conservation.
2422
2423 The basic equations of motion being integrated are derived from the
2424 Hamiltonian for conservative systems containing rigid bodies,
2425 \begin{equation}
2426 H = \sum_{i} \left( \frac{1}{2} m_i {\bf v}_i^T \cdot {\bf v}_i +
2427 \frac{1}{2} {\bf j}_i^T \cdot \overleftrightarrow{\mathsf{I}}_i^{-1} \cdot
2428 {\bf j}_i \right) +
2429 V\left(\left\{{\bf r}\right\}, \left\{\mathsf{A}\right\}\right),
2430 \end{equation}
2431 where ${\bf r}_i$ and ${\bf v}_i$ are the cartesian position vector
2432 and velocity of the center of mass of particle $i$, and ${\bf j}_i$,
2433 $\overleftrightarrow{\mathsf{I}}_i$ are the body-fixed angular
2434 momentum and moment of inertia tensor respectively, and the
2435 superscript $T$ denotes the transpose of the vector. $\mathsf{A}_i$
2436 is the $3 \times 3$ rotation matrix describing the instantaneous
2437 orientation of the particle. $V$ is the potential energy function
2438 which may depend on both the positions $\left\{{\bf r}\right\}$ and
2439 orientations $\left\{\mathsf{A}\right\}$ of all particles. The
2440 equations of motion for the particle centers of mass are derived from
2441 Hamilton's equations and are quite simple,
2442 \begin{eqnarray}
2443 \dot{{\bf r}} & = & {\bf v}, \\
2444 \dot{{\bf v}} & = & \frac{{\bf f}}{m},
2445 \end{eqnarray}
2446 where ${\bf f}$ is the instantaneous force on the center of mass
2447 of the particle,
2448 \begin{equation}
2449 {\bf f} = - \frac{\partial}{\partial
2450 {\bf r}} V(\left\{{\bf r}(t)\right\}, \left\{\mathsf{A}(t)\right\}).
2451 \end{equation}
2452
2453 The equations of motion for the orientational degrees of freedom are
2454 \begin{eqnarray}
2455 \dot{\mathsf{A}} & = & \mathsf{A} \cdot
2456 \mbox{ skew}\left(\overleftrightarrow{\mathsf{I}}^{-1} \cdot {\bf j}\right),\\
2457 \dot{{\bf j}} & = & {\bf j} \times \left( \overleftrightarrow{\mathsf{I}}^{-1}
2458 \cdot {\bf j} \right) - \mbox{ rot}\left(\mathsf{A}^{T} \cdot \frac{\partial
2459 V}{\partial \mathsf{A}} \right).
2460 \end{eqnarray}
2461 In these equations of motion, the $\mbox{skew}$ matrix of a vector
2462 ${\bf v} = \left( v_1, v_2, v_3 \right)$ is defined:
2463 \begin{equation}
2464 \mbox{skew}\left( {\bf v} \right) := \left(
2465 \begin{array}{ccc}
2466 0 & v_3 & - v_2 \\
2467 -v_3 & 0 & v_1 \\
2468 v_2 & -v_1 & 0
2469 \end{array}
2470 \right).
2471 \end{equation}
2472 The $\mbox{rot}$ notation refers to the mapping of the $3 \times 3$
2473 rotation matrix to a vector of orientations by first computing the
2474 skew-symmetric part $\left(\mathsf{A} - \mathsf{A}^{T}\right)$ and
2475 then associating this with a length 3 vector by inverting the
2476 $\mbox{skew}$ function above:
2477 \begin{equation}
2478 \mbox{rot}\left(\mathsf{A}\right) := \mbox{ skew}^{-1}\left(\mathsf{A}
2479 - \mathsf{A}^{T} \right).
2480 \end{equation}
2481 Written this way, the $\mbox{rot}$ operation creates a set of
2482 conjugate angle coordinates to the body-fixed angular momenta
2483 represented by ${\bf j}$. This equation of motion for angular momenta
2484 is equivalent to the more familiar body-fixed forms,
2485 \begin{eqnarray}
2486 \dot{j_{x}} & = & \tau^b_x(t) -
2487 \left(\overleftrightarrow{\mathsf{I}}_{yy}^{-1} - \overleftrightarrow{\mathsf{I}}_{zz}^{-1} \right) j_y j_z, \\
2488 \dot{j_{y}} & = & \tau^b_y(t) -
2489 \left(\overleftrightarrow{\mathsf{I}}_{zz}^{-1} - \overleftrightarrow{\mathsf{I}}_{xx}^{-1} \right) j_z j_x,\\
2490 \dot{j_{z}} & = & \tau^b_z(t) -
2491 \left(\overleftrightarrow{\mathsf{I}}_{xx}^{-1} - \overleftrightarrow{\mathsf{I}}_{yy}^{-1} \right) j_x j_y,
2492 \end{eqnarray}
2493 which utilize the body-fixed torques, ${\bf \tau}^b$. Torques are
2494 most easily derived in the space-fixed frame,
2495 \begin{equation}
2496 {\bf \tau}^b(t) = \mathsf{A}(t) \cdot {\bf \tau}^s(t),
2497 \end{equation}
2498 where the torques are either derived from the forces on the
2499 constituent atoms of the rigid body, or for directional atoms,
2500 directly from derivatives of the potential energy,
2501 \begin{equation}
2502 {\bf \tau}^s(t) = - \hat{\bf u}(t) \times \left( \frac{\partial}
2503 {\partial \hat{\bf u}} V\left(\left\{ {\bf r}(t) \right\}, \left\{
2504 \mathsf{A}(t) \right\}\right) \right).
2505 \end{equation}
2506 Here $\hat{\bf u}$ is a unit vector pointing along the principal axis
2507 of the particle in the space-fixed frame.
2508
2509 The {\sc dlm} method uses a Trotter factorization of the orientational
2510 propagator. This has three effects:
2511 \begin{enumerate}
2512 \item the integrator is area-preserving in phase space (i.e. it is
2513 {\it symplectic}),
2514 \item the integrator is time-{\it reversible}, making it suitable for Hybrid
2515 Monte Carlo applications, and
2516 \item the error for a single time step is of order $\mathcal{O}\left(h^4\right)$
2517 for timesteps of length $h$.
2518 \end{enumerate}
2519
2520 The integration of the equations of motion is carried out in a
2521 velocity-Verlet style 2-part algorithm, where $h= \delta t$:
2522
2523 {\tt moveA:}
2524 \begin{align*}
2525 {\bf v}\left(t + h / 2\right) &\leftarrow {\bf v}(t)
2526 + \frac{h}{2} \left( {\bf f}(t) / m \right), \\
2527 %
2528 {\bf r}(t + h) &\leftarrow {\bf r}(t)
2529 + h {\bf v}\left(t + h / 2 \right), \\
2530 %
2531 {\bf j}\left(t + h / 2 \right) &\leftarrow {\bf j}(t)
2532 + \frac{h}{2} {\bf \tau}^b(t), \\
2533 %
2534 \mathsf{A}(t + h) &\leftarrow \mathrm{rotate}\left( h {\bf j}
2535 (t + h / 2) \cdot \overleftrightarrow{\mathsf{I}}^{-1} \right).
2536 \end{align*}
2537
2538 In this context, the $\mathrm{rotate}$ function is the reversible product
2539 of the three body-fixed rotations,
2540 \begin{equation}
2541 \mathrm{rotate}({\bf a}) = \mathsf{G}_x(a_x / 2) \cdot
2542 \mathsf{G}_y(a_y / 2) \cdot \mathsf{G}_z(a_z) \cdot \mathsf{G}_y(a_y /
2543 2) \cdot \mathsf{G}_x(a_x /2),
2544 \end{equation}
2545 where each rotational propagator, $\mathsf{G}_\alpha(\theta)$, rotates
2546 both the rotation matrix ($\mathsf{A}$) and the body-fixed angular
2547 momentum (${\bf j}$) by an angle $\theta$ around body-fixed axis
2548 $\alpha$,
2549 \begin{equation}
2550 \mathsf{G}_\alpha( \theta ) = \left\{
2551 \begin{array}{lcl}
2552 \mathsf{A}(t) & \leftarrow & \mathsf{A}(0) \cdot \mathsf{R}_\alpha(\theta)^T, \\
2553 {\bf j}(t) & \leftarrow & \mathsf{R}_\alpha(\theta) \cdot {\bf j}(0).
2554 \end{array}
2555 \right.
2556 \end{equation}
2557 $\mathsf{R}_\alpha$ is a quadratic approximation to
2558 the single-axis rotation matrix. For example, in the small-angle
2559 limit, the rotation matrix around the body-fixed x-axis can be
2560 approximated as
2561 \begin{equation}
2562 \mathsf{R}_x(\theta) \approx \left(
2563 \begin{array}{ccc}
2564 1 & 0 & 0 \\
2565 0 & \frac{1-\theta^2 / 4}{1 + \theta^2 / 4} & -\frac{\theta}{1+
2566 \theta^2 / 4} \\
2567 0 & \frac{\theta}{1+
2568 \theta^2 / 4} & \frac{1-\theta^2 / 4}{1 + \theta^2 / 4}
2569 \end{array}
2570 \right).
2571 \end{equation}
2572 All other rotations follow in a straightforward manner.
2573
2574 After the first part of the propagation, the forces and body-fixed
2575 torques are calculated at the new positions and orientations
2576
2577 {\tt doForces:}
2578 \begin{align*}
2579 {\bf f}(t + h) &\leftarrow
2580 - \left(\frac{\partial V}{\partial {\bf r}}\right)_{{\bf r}(t + h)}, \\
2581 %
2582 {\bf \tau}^{s}(t + h) &\leftarrow {\bf u}(t + h)
2583 \times \frac{\partial V}{\partial {\bf u}}, \\
2584 %
2585 {\bf \tau}^{b}(t + h) &\leftarrow \mathsf{A}(t + h)
2586 \cdot {\bf \tau}^s(t + h).
2587 \end{align*}
2588
2589 {\sc OpenMD} automatically updates ${\bf u}$ when the rotation matrix
2590 $\mathsf{A}$ is calculated in {\tt moveA}. Once the forces and
2591 torques have been obtained at the new time step, the velocities can be
2592 advanced to the same time value.
2593
2594 {\tt moveB:}
2595 \begin{align*}
2596 {\bf v}\left(t + h \right) &\leftarrow {\bf v}\left(t + h / 2 \right)
2597 + \frac{h}{2} \left( {\bf f}(t + h) / m \right), \\
2598 %
2599 {\bf j}\left(t + h \right) &\leftarrow {\bf j}\left(t + h / 2 \right)
2600 + \frac{h}{2} {\bf \tau}^b(t + h) .
2601 \end{align*}
2602
2603 The matrix rotations used in the {\sc dlm} method end up being more
2604 costly computationally than the simpler arithmetic quaternion
2605 propagation. With the same time step, a 1024-molecule water simulation
2606 incurs an average 12\% increase in computation time using the {\sc
2607 dlm} method in place of quaternions. This cost is more than justified
2608 when comparing the energy conservation achieved by the two
2609 methods. Figure ~\ref{quatdlm} provides a comparative analysis of the
2610 {\sc dlm} method versus the traditional quaternion scheme.
2611
2612 \begin{figure}
2613 \centering
2614 \includegraphics[width=\linewidth]{quatvsdlm.pdf}
2615 \caption[Energy conservation analysis of the {\sc dlm} and quaternion
2616 integration methods]{Analysis of the energy conservation of the {\sc
2617 dlm} and quaternion integration methods. $\delta \mathrm{E}_1$ is the
2618 linear drift in energy over time and $\delta \mathrm{E}_0$ is the
2619 standard deviation of energy fluctuations around this drift. All
2620 simulations were of a 1024-molecule simulation of SSD water at 298 K
2621 starting from the same initial configuration. Note that the {\sc dlm}
2622 method provides more than an order of magnitude improvement in both
2623 the energy drift and the size of the energy fluctuations when compared
2624 with the quaternion method at any given time step. At time steps
2625 larger than 4 fs, the quaternion scheme resulted in rapidly rising
2626 energies which eventually lead to simulation failure. Using the {\sc
2627 dlm} method, time steps up to 8 fs can be taken before this behavior
2628 is evident.}
2629 \label{quatdlm}
2630 \end{figure}
2631
2632 In Fig.~\ref{quatdlm}, $\delta \mbox{E}_1$ is a measure of the linear
2633 energy drift in units of $\mbox{kcal mol}^{-1}$ per particle over a
2634 nanosecond of simulation time, and $\delta \mbox{E}_0$ is the standard
2635 deviation of the energy fluctuations in units of $\mbox{kcal
2636 mol}^{-1}$ per particle. In the top plot, it is apparent that the
2637 energy drift is reduced by a significant amount (2 to 3 orders of
2638 magnitude improvement at all tested time steps) by chosing the {\sc
2639 dlm} method over the simple non-symplectic quaternion integration
2640 method. In addition to this improvement in energy drift, the
2641 fluctuations in the total energy are also dampened by 1 to 2 orders of
2642 magnitude by utilizing the {\sc dlm} method.
2643
2644 Although the {\sc dlm} method is more computationally expensive than
2645 the traditional quaternion scheme for propagating a single time step,
2646 consideration of the computational cost for a long simulation with a
2647 particular level of energy conservation is in order. A plot of energy
2648 drift versus computational cost was generated
2649 (Fig.~\ref{cpuCost}). This figure provides an estimate of the CPU time
2650 required under the two integration schemes for 1 nanosecond of
2651 simulation time for the model 1024-molecule system. By chosing a
2652 desired energy drift value it is possible to determine the CPU time
2653 required for both methods. If a $\delta \mbox{E}_1$ of $1 \times
2654 10^{-3} \mbox{kcal mol}^{-1}$ per particle is desired, a nanosecond of
2655 simulation time will require ~19 hours of CPU time with the {\sc dlm}
2656 integrator, while the quaternion scheme will require ~154 hours of CPU
2657 time. This demonstrates the computational advantage of the integration
2658 scheme utilized in {\sc OpenMD}.
2659
2660 \begin{figure}
2661 \centering
2662 \includegraphics[width=\linewidth]{compCost.pdf}
2663 \caption[Energy drift as a function of required simulation run
2664 time]{Energy drift as a function of required simulation run time.
2665 $\delta \mathrm{E}_1$ is the linear drift in energy over time.
2666 Simulations were performed on a single 2.5 GHz Pentium 4
2667 processor. Simulation time comparisons can be made by tracing
2668 horizontally from one curve to the other. For example, a simulation
2669 that takes ~24 hours using the {\sc dlm} method will take roughly 210
2670 hours using the simple quaternion method if the same degree of energy
2671 conservation is desired.}
2672 \label{cpuCost}
2673 \end{figure}
2674
2675 There is only one specific keyword relevant to the default integrator,
2676 and that is the time step for integrating the equations of motion.
2677
2678 \begin{center}
2679 \begin{tabular}{llll}
2680 {\bf variable} & {\bf Meta-data keyword} & {\bf units} & {\bf
2681 default value} \\
2682 $h$ & {\tt dt = 2.0;} & fs & none
2683 \end{tabular}
2684 \end{center}
2685
2686 \section{\label{sec:extended}Extended Systems for other Ensembles}
2687
2688 {\sc OpenMD} implements a number of extended system integrators for
2689 sampling from other ensembles relevant to chemical physics. The
2690 integrator can be selected with the {\tt ensemble} keyword in the
2691 meta-data file:
2692
2693 \begin{center}
2694 \begin{tabular}{lll}
2695 {\bf Integrator} & {\bf Ensemble} & {\bf Meta-data instruction} \\
2696 NVE & microcanonical & {\tt ensemble = NVE; } \\
2697 NVT & canonical & {\tt ensemble = NVT; } \\
2698 NPTi & isobaric-isothermal & {\tt ensemble = NPTi;} \\
2699 & (with isotropic volume changes) & \\
2700 NPTf & isobaric-isothermal & {\tt ensemble = NPTf;} \\
2701 & (with changes to box shape) & \\
2702 NPTxyz & approximate isobaric-isothermal & {\tt ensemble = NPTxyz;} \\
2703 & (with separate barostats on each box dimension) & \\
2704 LD & Langevin Dynamics & {\tt ensemble = LD;} \\
2705 & (approximates the effects of an implicit solvent) & \\
2706 LangevinHull & Non-periodic Langevin Dynamics & {\tt ensemble = LangevinHull;} \\
2707 & (Langevin Dynamics for molecules on convex hull;\\
2708 & Newtonian for interior molecules) & \\
2709 \end{tabular}
2710 \end{center}
2711
2712 The relatively well-known Nos\'e-Hoover thermostat\cite{Hoover85} is
2713 implemented in {\sc OpenMD}'s NVT integrator. This method couples an
2714 extra degree of freedom (the thermostat) to the kinetic energy of the
2715 system and it has been shown to sample the canonical distribution in
2716 the system degrees of freedom while conserving a quantity that is, to
2717 within a constant, the Helmholtz free energy.\cite{melchionna93}
2718
2719 NPT algorithms attempt to maintain constant pressure in the system by
2720 coupling the volume of the system to a barostat. {\sc OpenMD} contains
2721 three different constant pressure algorithms. The first two, NPTi and
2722 NPTf have been shown to conserve a quantity that is, to within a
2723 constant, the Gibbs free energy.\cite{melchionna93} The Melchionna
2724 modification to the Hoover barostat is implemented in both NPTi and
2725 NPTf. NPTi allows only isotropic changes in the simulation box, while
2726 box {\it shape} variations are allowed in NPTf. The NPTxyz integrator
2727 has {\it not} been shown to sample from the isobaric-isothermal
2728 ensemble. It is useful, however, in that it maintains orthogonality
2729 for the axes of the simulation box while attempting to equalize
2730 pressure along the three perpendicular directions in the box.
2731
2732 Each of the extended system integrators requires additional keywords
2733 to set target values for the thermodynamic state variables that are
2734 being held constant. Keywords are also required to set the
2735 characteristic decay times for the dynamics of the extended
2736 variables.
2737
2738 \begin{center}
2739 \begin{tabular}{llll}
2740 {\bf variable} & {\bf Meta-data instruction} & {\bf units} & {\bf
2741 default value} \\
2742 $T_{\mathrm{target}}$ & {\tt targetTemperature = 300;} & K & none \\
2743 $P_{\mathrm{target}}$ & {\tt targetPressure = 1;} & atm & none \\
2744 $\tau_T$ & {\tt tauThermostat = 1e3;} & fs & none \\
2745 $\tau_B$ & {\tt tauBarostat = 5e3;} & fs & none \\
2746 & {\tt resetTime = 200;} & fs & none \\
2747 & {\tt useInitialExtendedSystemState = true;} & logical &
2748 true
2749 \end{tabular}
2750 \end{center}
2751
2752 Two additional keywords can be used to either clear the extended
2753 system variables periodically ({\tt resetTime}), or to maintain the
2754 state of the extended system variables between simulations ({\tt
2755 useInitialExtendedSystemState}). More details on these variables
2756 and their use in the integrators follows below.
2757
2758 \section{\label{section:noseHooverThermo}Nos\'{e}-Hoover Thermostatting}
2759
2760 The Nos\'e-Hoover equations of motion are given by\cite{Hoover85}
2761 \begin{eqnarray}
2762 \dot{{\bf r}} & = & {\bf v}, \\
2763 \dot{{\bf v}} & = & \frac{{\bf f}}{m} - \chi {\bf v} ,\\
2764 \dot{\mathsf{A}} & = & \mathsf{A} \cdot
2765 \mbox{ skew}\left(\overleftrightarrow{\mathsf{I}}^{-1} \cdot {\bf j}\right), \\
2766 \dot{{\bf j}} & = & {\bf j} \times \left( \overleftrightarrow{\mathsf{I}}^{-1}
2767 \cdot {\bf j} \right) - \mbox{ rot}\left(\mathsf{A}^{T} \cdot \frac{\partial
2768 V}{\partial \mathsf{A}} \right) - \chi {\bf j}.
2769 \label{eq:nosehoovereom}
2770 \end{eqnarray}
2771
2772 $\chi$ is an ``extra'' variable included in the extended system, and
2773 it is propagated using the first order equation of motion
2774 \begin{equation}
2775 \dot{\chi} = \frac{1}{\tau_{T}^2} \left( \frac{T}{T_{\mathrm{target}}} - 1 \right).
2776 \label{eq:nosehooverext}
2777 \end{equation}
2778
2779 The instantaneous temperature $T$ is proportional to the total kinetic
2780 energy (both translational and orientational) and is given by
2781 \begin{equation}
2782 T = \frac{2 K}{f k_B}
2783 \end{equation}
2784 Here, $f$ is the total number of degrees of freedom in the system,
2785 \begin{equation}
2786 f = 3 N + 2 N_{\mathrm{linear}} + 3 N_{\mathrm{non-linear}} - N_{\mathrm{constraints}},
2787 \end{equation}
2788 and $K$ is the total kinetic energy,
2789 \begin{equation}
2790 K = \sum_{i=1}^{N} \frac{1}{2} m_i {\bf v}_i^T \cdot {\bf v}_i +
2791 \sum_{i=1}^{N_{\mathrm{linear}}+N_{\mathrm{non-linear}}} \frac{1}{2} {\bf j}_i^T \cdot
2792 \overleftrightarrow{\mathsf{I}}_i^{-1} \cdot {\bf j}_i.
2793 \end{equation}
2794 $N_{\mathrm{linear}}$ is the number of linear rotors (i.e. with two
2795 non-zero moments of inertia), and $N_{\mathrm{non-linear}}$ is the
2796 number of non-linear rotors (i.e. with three non-zero moments of
2797 inertia).
2798
2799 In eq.(\ref{eq:nosehooverext}), $\tau_T$ is the time constant for
2800 relaxation of the temperature to the target value. To set values for
2801 $\tau_T$ or $T_{\mathrm{target}}$ in a simulation, one would use the
2802 {\tt tauThermostat} and {\tt targetTemperature} keywords in the
2803 meta-data file. The units for {\tt tauThermostat} are fs, and the
2804 units for the {\tt targetTemperature} are degrees K. The integration
2805 of the equations of motion is carried out in a velocity-Verlet style 2
2806 part algorithm:
2807
2808 {\tt moveA:}
2809 \begin{align*}
2810 T(t) &\leftarrow \left\{{\bf v}(t)\right\}, \left\{{\bf j}(t)\right\} ,\\
2811 %
2812 {\bf v}\left(t + h / 2\right) &\leftarrow {\bf v}(t)
2813 + \frac{h}{2} \left( \frac{{\bf f}(t)}{m} - {\bf v}(t)
2814 \chi(t)\right), \\
2815 %
2816 {\bf r}(t + h) &\leftarrow {\bf r}(t)
2817 + h {\bf v}\left(t + h / 2 \right) ,\\
2818 %
2819 {\bf j}\left(t + h / 2 \right) &\leftarrow {\bf j}(t)
2820 + \frac{h}{2} \left( {\bf \tau}^b(t) - {\bf j}(t)
2821 \chi(t) \right) ,\\
2822 %
2823 \mathsf{A}(t + h) &\leftarrow \mathrm{rotate}
2824 \left(h * {\bf j}(t + h / 2)
2825 \overleftrightarrow{\mathsf{I}}^{-1} \right) ,\\
2826 %
2827 \chi\left(t + h / 2 \right) &\leftarrow \chi(t)
2828 + \frac{h}{2 \tau_T^2} \left( \frac{T(t)}
2829 {T_{\mathrm{target}}} - 1 \right) .
2830 \end{align*}
2831
2832 Here $\mathrm{rotate}(h * {\bf j}
2833 \overleftrightarrow{\mathsf{I}}^{-1})$ is the same symplectic Trotter
2834 factorization of the three rotation operations that was discussed in
2835 the section on the {\sc dlm} integrator. Note that this operation modifies
2836 both the rotation matrix $\mathsf{A}$ and the angular momentum ${\bf
2837 j}$. {\tt moveA} propagates velocities by a half time step, and
2838 positional degrees of freedom by a full time step. The new positions
2839 (and orientations) are then used to calculate a new set of forces and
2840 torques in exactly the same way they are calculated in the {\tt
2841 doForces} portion of the {\sc dlm} integrator.
2842
2843 Once the forces and torques have been obtained at the new time step,
2844 the temperature, velocities, and the extended system variable can be
2845 advanced to the same time value.
2846
2847 {\tt moveB:}
2848 \begin{align*}
2849 T(t + h) &\leftarrow \left\{{\bf v}(t + h)\right\},
2850 \left\{{\bf j}(t + h)\right\}, \\
2851 %
2852 \chi\left(t + h \right) &\leftarrow \chi\left(t + h /
2853 2 \right) + \frac{h}{2 \tau_T^2} \left( \frac{T(t+h)}
2854 {T_{\mathrm{target}}} - 1 \right), \\
2855 %
2856 {\bf v}\left(t + h \right) &\leftarrow {\bf v}\left(t
2857 + h / 2 \right) + \frac{h}{2} \left(
2858 \frac{{\bf f}(t + h)}{m} - {\bf v}(t + h)
2859 \chi(t h)\right) ,\\
2860 %
2861 {\bf j}\left(t + h \right) &\leftarrow {\bf j}\left(t
2862 + h / 2 \right) + \frac{h}{2}
2863 \left( {\bf \tau}^b(t + h) - {\bf j}(t + h)
2864 \chi(t + h) \right) .
2865 \end{align*}
2866
2867 Since ${\bf v}(t + h)$ and ${\bf j}(t + h)$ are required to calculate
2868 $T(t + h)$ as well as $\chi(t + h)$, they indirectly depend on their
2869 own values at time $t + h$. {\tt moveB} is therefore done in an
2870 iterative fashion until $\chi(t + h)$ becomes self-consistent. The
2871 relative tolerance for the self-consistency check defaults to a value
2872 of $\mbox{10}^{-6}$, but {\sc OpenMD} will terminate the iteration
2873 after 4 loops even if the consistency check has not been satisfied.
2874
2875 The Nos\'e-Hoover algorithm is known to conserve a Hamiltonian for the
2876 extended system that is, to within a constant, identical to the
2877 Helmholtz free energy,\cite{melchionna93}
2878 \begin{equation}
2879 H_{\mathrm{NVT}} = V + K + f k_B T_{\mathrm{target}} \left(
2880 \frac{\tau_{T}^2 \chi^2(t)}{2} + \int_{0}^{t} \chi(t^\prime) dt^\prime
2881 \right).
2882 \end{equation}
2883 Poor choices of $h$ or $\tau_T$ can result in non-conservation
2884 of $H_{\mathrm{NVT}}$, so the conserved quantity is maintained in the
2885 last column of the {\tt .stat} file to allow checks on the quality of
2886 the integration.
2887
2888 Bond constraints are applied at the end of both the {\tt moveA} and
2889 {\tt moveB} portions of the algorithm. Details on the constraint
2890 algorithms are given in section \ref{section:rattle}.
2891
2892 \section{\label{sec:NPTi}Constant-pressure integration with
2893 isotropic box deformations (NPTi)}
2894
2895 To carry out isobaric-isothermal ensemble calculations, {\sc OpenMD}
2896 implements the Melchionna modifications to the Nos\'e-Hoover-Andersen
2897 equations of motion.\cite{melchionna93} The equations of motion are
2898 the same as NVT with the following exceptions:
2899
2900 \begin{eqnarray}
2901 \dot{{\bf r}} & = & {\bf v} + \eta \left( {\bf r} - {\bf R}_0 \right), \\
2902 \dot{{\bf v}} & = & \frac{{\bf f}}{m} - (\eta + \chi) {\bf v}, \\
2903 \dot{\eta} & = & \frac{1}{\tau_{B}^2 f k_B T_{\mathrm{target}}} V \left( P -
2904 P_{\mathrm{target}} \right), \\
2905 \dot{\mathcal{V}} & = & 3 \mathcal{V} \eta .
2906 \label{eq:melchionna1}
2907 \end{eqnarray}
2908
2909 $\chi$ and $\eta$ are the ``extra'' degrees of freedom in the extended
2910 system. $\chi$ is a thermostat, and it has the same function as it
2911 does in the Nos\'e-Hoover NVT integrator. $\eta$ is a barostat which
2912 controls changes to the volume of the simulation box. ${\bf R}_0$ is
2913 the location of the center of mass for the entire system, and
2914 $\mathcal{V}$ is the volume of the simulation box. At any time, the
2915 volume can be calculated from the determinant of the matrix which
2916 describes the box shape:
2917 \begin{equation}
2918 \mathcal{V} = \det(\mathsf{H}).
2919 \end{equation}
2920
2921 The NPTi integrator requires an instantaneous pressure. This quantity
2922 is calculated via the pressure tensor,
2923 \begin{equation}
2924 \overleftrightarrow{\mathsf{P}}(t) = \frac{1}{\mathcal{V}(t)} \left(
2925 \sum_{i=1}^{N} m_i {\bf v}_i(t) \otimes {\bf v}_i(t) \right) +
2926 \overleftrightarrow{\mathsf{W}}(t).
2927 \end{equation}
2928 The kinetic contribution to the pressure tensor utilizes the {\it
2929 outer} product of the velocities, denoted by the $\otimes$ symbol. The
2930 stress tensor is calculated from another outer product of the
2931 inter-atomic separation vectors (${\bf r}_{ij} = {\bf r}_j - {\bf
2932 r}_i$) with the forces between the same two atoms,
2933 \begin{equation}
2934 \overleftrightarrow{\mathsf{W}}(t) = \sum_{i} \sum_{j>i} {\bf r}_{ij}(t)
2935 \otimes {\bf f}_{ij}(t).
2936 \end{equation}
2937 In systems containing cutoff groups, the stress tensor is computed
2938 between the centers-of-mass of the cutoff groups:
2939 \begin{equation}
2940 \overleftrightarrow{\mathsf{W}}(t) = \sum_{a} \sum_{b} {\bf r}_{ab}(t)
2941 \otimes {\bf f}_{ab}(t).
2942 \end{equation}
2943 where ${\bf r}_{ab}$ is the distance between the centers of mass, and
2944 \begin{equation}
2945 {\bf f}_{ab} = s(r_{ab}) \sum_{i \in a} \sum_{j \in b} {\bf f}_{ij} +
2946 s^{\prime}(r_{ab}) \frac{{\bf r}_{ab}}{|r_{ab}|} \sum_{i \in a} \sum_{j
2947 \in b} V_{ij}({\bf r}_{ij}).
2948 \end{equation}
2949
2950 The instantaneous pressure is then simply obtained from the trace of
2951 the pressure tensor,
2952 \begin{equation}
2953 P(t) = \frac{1}{3} \mathrm{Tr} \left( \overleftrightarrow{\mathsf{P}}(t)
2954 \right).
2955 \end{equation}
2956
2957 In eq.(\ref{eq:melchionna1}), $\tau_B$ is the time constant for
2958 relaxation of the pressure to the target value. To set values for
2959 $\tau_B$ or $P_{\mathrm{target}}$ in a simulation, one would use the
2960 {\tt tauBarostat} and {\tt targetPressure} keywords in the meta-data
2961 file. The units for {\tt tauBarostat} are fs, and the units for the
2962 {\tt targetPressure} are atmospheres. Like in the NVT integrator, the
2963 integration of the equations of motion is carried out in a
2964 velocity-Verlet style two part algorithm with only the following
2965 differences:
2966
2967 {\tt moveA:}
2968 \begin{align*}
2969 P(t) &\leftarrow \left\{{\bf r}(t)\right\}, \left\{{\bf v}(t)\right\} ,\\
2970 %
2971 {\bf v}\left(t + h / 2\right) &\leftarrow {\bf v}(t)
2972 + \frac{h}{2} \left( \frac{{\bf f}(t)}{m} - {\bf v}(t)
2973 \left(\chi(t) + \eta(t) \right) \right), \\
2974 %
2975 \eta(t + h / 2) &\leftarrow \eta(t) + \frac{h
2976 \mathcal{V}(t)}{2 N k_B T(t) \tau_B^2} \left( P(t)
2977 - P_{\mathrm{target}} \right), \\
2978 %
2979 {\bf r}(t + h) &\leftarrow {\bf r}(t) + h
2980 \left\{ {\bf v}\left(t + h / 2 \right)
2981 + \eta(t + h / 2)\left[ {\bf r}(t + h)
2982 - {\bf R}_0 \right] \right\} ,\\
2983 %
2984 \mathsf{H}(t + h) &\leftarrow e^{-h \eta(t + h / 2)}
2985 \mathsf{H}(t).
2986 \end{align*}
2987
2988 The propagation of positions to time $t + h$
2989 depends on the positions at the same time. {\sc OpenMD} carries out
2990 this step iteratively (with a limit of 5 passes through the iterative
2991 loop). Also, the simulation box $\mathsf{H}$ is scaled uniformly for
2992 one full time step by an exponential factor that depends on the value
2993 of $\eta$ at time $t +
2994 h / 2$. Reshaping the box uniformly also scales the volume of
2995 the box by
2996 \begin{equation}
2997 \mathcal{V}(t + h) \leftarrow e^{ - 3 h \eta(t + h /2)} \times
2998 \mathcal{V}(t).
2999 \end{equation}
3000
3001 The {\tt doForces} step for the NPTi integrator is exactly the same as
3002 in both the {\sc dlm} and NVT integrators. Once the forces and torques have
3003 been obtained at the new time step, the velocities can be advanced to
3004 the same time value.
3005
3006 {\tt moveB:}
3007 \begin{align*}
3008 P(t + h) &\leftarrow \left\{{\bf r}(t + h)\right\},
3009 \left\{{\bf v}(t + h)\right\}, \\
3010 %
3011 \eta(t + h) &\leftarrow \eta(t + h / 2) +
3012 \frac{h \mathcal{V}(t + h)}{2 N k_B T(t + h)
3013 \tau_B^2} \left( P(t + h) - P_{\mathrm{target}} \right), \\
3014 %
3015 {\bf v}\left(t + h \right) &\leftarrow {\bf v}\left(t
3016 + h / 2 \right) + \frac{h}{2} \left(
3017 \frac{{\bf f}(t + h)}{m} - {\bf v}(t + h)
3018 (\chi(t + h) + \eta(t + h)) \right) ,\\
3019 %
3020 {\bf j}\left(t + h \right) &\leftarrow {\bf j}\left(t
3021 + h / 2 \right) + \frac{h}{2} \left( {\bf
3022 \tau}^b(t + h) - {\bf j}(t + h)
3023 \chi(t + h) \right) .
3024 \end{align*}
3025
3026 Once again, since ${\bf v}(t + h)$ and ${\bf j}(t + h)$ are required
3027 to calculate $T(t + h)$, $P(t + h)$, $\chi(t + h)$, and $\eta(t +
3028 h)$, they indirectly depend on their own values at time $t + h$. {\tt
3029 moveB} is therefore done in an iterative fashion until $\chi(t + h)$
3030 and $\eta(t + h)$ become self-consistent. The relative tolerance for
3031 the self-consistency check defaults to a value of $\mbox{10}^{-6}$,
3032 but {\sc OpenMD} will terminate the iteration after 4 loops even if the
3033 consistency check has not been satisfied.
3034
3035 The Melchionna modification of the Nos\'e-Hoover-Andersen algorithm is
3036 known to conserve a Hamiltonian for the extended system that is, to
3037 within a constant, identical to the Gibbs free energy,
3038 \begin{equation}
3039 H_{\mathrm{NPTi}} = V + K + f k_B T_{\mathrm{target}} \left(
3040 \frac{\tau_{T}^2 \chi^2(t)}{2} + \int_{0}^{t} \chi(t^\prime) dt^\prime
3041 \right) + P_{\mathrm{target}} \mathcal{V}(t).
3042 \end{equation}
3043 Poor choices of $\delta t$, $\tau_T$, or $\tau_B$ can result in
3044 non-conservation of $H_{\mathrm{NPTi}}$, so the conserved quantity is
3045 maintained in the last column of the {\tt .stat} file to allow checks
3046 on the quality of the integration. It is also known that this
3047 algorithm samples the equilibrium distribution for the enthalpy
3048 (including contributions for the thermostat and barostat),
3049 \begin{equation}
3050 H_{\mathrm{NPTi}} = V + K + \frac{f k_B T_{\mathrm{target}}}{2} \left(
3051 \chi^2 \tau_T^2 + \eta^2 \tau_B^2 \right) + P_{\mathrm{target}}
3052 \mathcal{V}(t).
3053 \end{equation}
3054
3055 Bond constraints are applied at the end of both the {\tt moveA} and
3056 {\tt moveB} portions of the algorithm. Details on the constraint
3057 algorithms are given in section \ref{section:rattle}.
3058
3059 \section{\label{sec:NPTf}Constant-pressure integration with a
3060 flexible box (NPTf)}
3061
3062 There is a relatively simple generalization of the
3063 Nos\'e-Hoover-Andersen method to include changes in the simulation box
3064 {\it shape} as well as in the volume of the box. This method utilizes
3065 the full $3 \times 3$ pressure tensor and introduces a tensor of
3066 extended variables ($\overleftrightarrow{\eta}$) to control changes to
3067 the box shape. The equations of motion for this method differ from
3068 those of NPTi as follows:
3069 \begin{eqnarray}
3070 \dot{{\bf r}} & = & {\bf v} + \overleftrightarrow{\eta} \cdot \left( {\bf r} - {\bf R}_0 \right), \\
3071 \dot{{\bf v}} & = & \frac{{\bf f}}{m} - (\overleftrightarrow{\eta} +
3072 \chi \cdot \mathsf{1}) {\bf v}, \\
3073 \dot{\overleftrightarrow{\eta}} & = & \frac{1}{\tau_{B}^2 f k_B
3074 T_{\mathrm{target}}} V \left( \overleftrightarrow{\mathsf{P}} - P_{\mathrm{target}}\mathsf{1} \right) ,\\
3075 \dot{\mathsf{H}} & = & \overleftrightarrow{\eta} \cdot \mathsf{H} .
3076 \label{eq:melchionna2}
3077 \end{eqnarray}
3078
3079 Here, $\mathsf{1}$ is the unit matrix and $\overleftrightarrow{\mathsf{P}}$
3080 is the pressure tensor. Again, the volume, $\mathcal{V} = \det
3081 \mathsf{H}$.
3082
3083 The propagation of the equations of motion is nearly identical to the
3084 NPTi integration:
3085
3086 {\tt moveA:}
3087 \begin{align*}
3088 \overleftrightarrow{\mathsf{P}}(t) &\leftarrow \left\{{\bf r}(t)\right\},
3089 \left\{{\bf v}(t)\right\} ,\\
3090 %
3091 {\bf v}\left(t + h / 2\right) &\leftarrow {\bf v}(t)
3092 + \frac{h}{2} \left( \frac{{\bf f}(t)}{m} -
3093 \left(\chi(t)\mathsf{1} + \overleftrightarrow{\eta}(t) \right) \cdot
3094 {\bf v}(t) \right), \\
3095 %
3096 \overleftrightarrow{\eta}(t + h / 2) &\leftarrow
3097 \overleftrightarrow{\eta}(t) + \frac{h \mathcal{V}(t)}{2 N k_B
3098 T(t) \tau_B^2} \left( \overleftrightarrow{\mathsf{P}}(t)
3099 - P_{\mathrm{target}}\mathsf{1} \right), \\
3100 %
3101 {\bf r}(t + h) &\leftarrow {\bf r}(t) + h \left\{ {\bf v}
3102 \left(t + h / 2 \right) + \overleftrightarrow{\eta}(t +
3103 h / 2) \cdot \left[ {\bf r}(t + h)
3104 - {\bf R}_0 \right] \right\}, \\
3105 %
3106 \mathsf{H}(t + h) &\leftarrow \mathsf{H}(t) \cdot e^{-h
3107 \overleftrightarrow{\eta}(t + h / 2)} .
3108 \end{align*}
3109 {\sc OpenMD} uses a power series expansion truncated at second order
3110 for the exponential operation which scales the simulation box.
3111
3112 The {\tt moveB} portion of the algorithm is largely unchanged from the
3113 NPTi integrator:
3114
3115 {\tt moveB:}
3116 \begin{align*}
3117 \overleftrightarrow{\mathsf{P}}(t + h) &\leftarrow \left\{{\bf r}
3118 (t + h)\right\}, \left\{{\bf v}(t
3119 + h)\right\}, \left\{{\bf f}(t + h)\right\} ,\\
3120 %
3121 \overleftrightarrow{\eta}(t + h) &\leftarrow
3122 \overleftrightarrow{\eta}(t + h / 2) +
3123 \frac{h \mathcal{V}(t + h)}{2 N k_B T(t + h)
3124 \tau_B^2} \left( \overleftrightarrow{P}(t + h)
3125 - P_{\mathrm{target}}\mathsf{1} \right) ,\\
3126 %
3127 {\bf v}\left(t + h \right) &\leftarrow {\bf v}\left(t
3128 + h / 2 \right) + \frac{h}{2} \left(
3129 \frac{{\bf f}(t + h)}{m} -
3130 (\chi(t + h)\mathsf{1} + \overleftrightarrow{\eta}(t
3131 + h)) \right) \cdot {\bf v}(t + h), \\
3132 \end{align*}
3133
3134 The iterative schemes for both {\tt moveA} and {\tt moveB} are
3135 identical to those described for the NPTi integrator.
3136
3137 The NPTf integrator is known to conserve the following Hamiltonian:
3138 \begin{equation}
3139 H_{\mathrm{NPTf}} = V + K + f k_B T_{\mathrm{target}} \left(
3140 \frac{\tau_{T}^2 \chi^2(t)}{2} + \int_{0}^{t} \chi(t^\prime) dt^\prime
3141 \right) + P_{\mathrm{target}} \mathcal{V}(t) + \frac{f k_B
3142 T_{\mathrm{target}}}{2}
3143 \mathrm{Tr}\left[\overleftrightarrow{\eta}(t)\right]^2 \tau_B^2.
3144 \end{equation}
3145
3146 This integrator must be used with care, particularly in liquid
3147 simulations. Liquids have very small restoring forces in the
3148 off-diagonal directions, and the simulation box can very quickly form
3149 elongated and sheared geometries which become smaller than the cutoff
3150 radius. The NPTf integrator finds most use in simulating crystals or
3151 liquid crystals which assume non-orthorhombic geometries.
3152
3153 \section{\label{nptxyz}Constant pressure in 3 axes (NPTxyz)}
3154
3155 There is one additional extended system integrator which is somewhat
3156 simpler than the NPTf method described above. In this case, the three
3157 axes have independent barostats which each attempt to preserve the
3158 target pressure along the box walls perpendicular to that particular
3159 axis. The lengths of the box axes are allowed to fluctuate
3160 independently, but the angle between the box axes does not change.
3161 The equations of motion are identical to those described above, but
3162 only the {\it diagonal} elements of $\overleftrightarrow{\eta}$ are
3163 computed. The off-diagonal elements are set to zero (even when the
3164 pressure tensor has non-zero off-diagonal elements).
3165
3166 It should be noted that the NPTxyz integrator is {\it not} known to
3167 preserve any Hamiltonian of interest to the chemical physics
3168 community. The integrator is extremely useful, however, in generating
3169 initial conditions for other integration methods. It {\it is} suitable
3170 for use with liquid simulations, or in cases where there is
3171 orientational anisotropy in the system (i.e. in lipid bilayer
3172 simulations).
3173
3174 \section{Langevin Dynamics (LD)\label{LDRB}}
3175
3176 {\sc OpenMD} implements a Langevin integrator in order to perform
3177 molecular dynamics simulations in implicit solvent environments. This
3178 can result in substantial performance gains when the detailed dynamics
3179 of the solvent is not important. Since {\sc OpenMD} also handles rigid
3180 bodies of arbitrary composition and shape, the Langevin integrator is
3181 by necessity somewhat more complex than in other simulation packages.
3182
3183 Consider the Langevin equations of motion in generalized coordinates
3184 \begin{equation}
3185 {\bf M} \dot{{\bf V}}(t) = {\bf F}_{s}(t) +
3186 {\bf F}_{f}(t) + {\bf F}_{r}(t)
3187 \label{LDGeneralizedForm}
3188 \end{equation}
3189 where ${\bf M}$ is a $6 \times 6$ diagonal mass matrix (which
3190 includes the mass of the rigid body as well as the moments of inertia
3191 in the body-fixed frame) and ${\bf V}$ is a generalized velocity,
3192 ${\bf V} =
3193 \left\{{\bf v},{\bf \omega}\right\}$. The right side of
3194 Eq. \ref{LDGeneralizedForm} consists of three generalized forces: a
3195 system force (${\bf F}_{s}$), a frictional or dissipative force (${\bf
3196 F}_{f}$) and a stochastic force (${\bf F}_{r}$). While the evolution
3197 of the system in Newtonian mechanics is typically done in the lab
3198 frame, it is convenient to handle the dynamics of rigid bodies in
3199 body-fixed frames. Thus the friction and random forces on each
3200 substructure are calculated in a body-fixed frame and may converted
3201 back to the lab frame using that substructure's rotation matrix (${\bf
3202 Q}$):
3203 \begin{equation}
3204 {\bf F}_{f,r} =
3205 \left( \begin{array}{c}
3206 {\bf f}_{f,r} \\
3207 {\bf \tau}_{f,r}
3208 \end{array} \right)
3209 =
3210 \left( \begin{array}{c}
3211 {\bf Q}^{T} {\bf f}^{~b}_{f,r} \\
3212 {\bf Q}^{T} {\bf \tau}^{~b}_{f,r}
3213 \end{array} \right)
3214 \end{equation}
3215 The body-fixed friction force, ${\bf F}_{f}^{~b}$, is proportional to
3216 the (body-fixed) velocity at the center of resistance
3217 ${\bf v}_{R}^{~b}$ and the angular velocity ${\bf \omega}$
3218 \begin{equation}
3219 {\bf F}_{f}^{~b}(t) = \left( \begin{array}{l}
3220 {\bf f}_{f}^{~b}(t) \\
3221 {\bf \tau}_{f}^{~b}(t) \\
3222 \end{array} \right) = - \left( \begin{array}{*{20}c}
3223 \Xi_{R}^{tt} & \Xi_{R}^{rt} \\
3224 \Xi_{R}^{tr} & \Xi_{R}^{rr} \\
3225 \end{array} \right)\left( \begin{array}{l}
3226 {\bf v}_{R}^{~b}(t) \\
3227 {\bf \omega}(t) \\
3228 \end{array} \right),
3229 \end{equation}
3230 while the random force, ${\bf F}_{r}$, is a Gaussian stochastic
3231 variable with zero mean and variance,
3232 \begin{equation}
3233 \left\langle {{\bf F}_{r}(t) ({\bf F}_{r}(t'))^T } \right\rangle =
3234 \left\langle {{\bf F}_{r}^{~b} (t) ({\bf F}_{r}^{~b} (t'))^T } \right\rangle =
3235 2 k_B T \Xi_R \delta(t - t'). \label{eq:randomForce}
3236 \end{equation}
3237 $\Xi_R$ is the $6\times6$ resistance tensor at the center of
3238 resistance.
3239
3240 For atoms and ellipsoids, there are good approximations for this
3241 tensor that are based on Stokes' law. For arbitrary rigid bodies, the
3242 resistance tensor must be pre-computed before Langevin dynamics can be
3243 used. The {\sc OpenMD} distribution contains a utitilty program called
3244 Hydro that performs this computation.
3245
3246 Once this tensor is known for a given {\tt integrableObject},
3247 obtaining a stochastic vector that has the properties in
3248 Eq. (\ref{eq:randomForce}) can be done efficiently by carrying out a
3249 one-time Cholesky decomposition to obtain the square root matrix of
3250 the resistance tensor,
3251 \begin{equation}
3252 \Xi_R = {\bf S} {\bf S}^{T},
3253 \label{eq:Cholesky}
3254 \end{equation}
3255 where ${\bf S}$ is a lower triangular matrix.\cite{Schlick2002} A
3256 vector with the statistics required for the random force can then be
3257 obtained by multiplying ${\bf S}$ onto a random 6-vector ${\bf Z}$ which
3258 has elements chosen from a Gaussian distribution, such that:
3259 \begin{equation}
3260 \langle {\bf Z}_i \rangle = 0, \hspace{1in} \langle {\bf Z}_i \cdot
3261 {\bf Z}_j \rangle = \frac{2 k_B T}{\delta t} \delta_{ij},
3262 \end{equation}
3263 where $\delta t$ is the timestep in use during the simulation. The
3264 random force, ${\bf F}_{r}^{~b} = {\bf S} {\bf Z}$, can be shown to have the
3265 correct properties required by Eq. (\ref{eq:randomForce}).
3266
3267 The equation of motion for the translational velocity of the center of
3268 mass (${\bf v}$) can be written as
3269 \begin{equation}
3270 m \dot{{\bf v}} (t) = {\bf f}_{s}(t) + {\bf f}_{f}(t) +
3271 {\bf f}_{r}(t)
3272 \end{equation}
3273 Since the frictional and random forces are applied at the center of
3274 resistance, which generally does not coincide with the center of mass,
3275 extra torques are exerted at the center of mass. Thus, the net
3276 body-fixed torque at the center of mass, $\tau^{~b}(t)$,
3277 is given by
3278 \begin{equation}
3279 \tau^{~b} \leftarrow \tau_{s}^{~b} + \tau_{f}^{~b} + \tau_{r}^{~b} + {\bf r}_{MR} \times \left( {\bf f}_{f}^{~b} + {\bf f}_{r}^{~b} \right)
3280 \end{equation}
3281 where ${\bf r}_{MR}$ is the vector from the center of mass to the center of
3282 resistance. Instead of integrating the angular velocity in lab-fixed
3283 frame, we consider the equation of motion for the angular momentum
3284 (${\bf j}$) in the body-fixed frame
3285 \begin{equation}
3286 \frac{\partial}{\partial t}{\bf j}(t) = \tau^{~b}(t)
3287 \end{equation}
3288 By embedding the friction and random forces into the the total force
3289 and torque, {\sc OpenMD} integrates the Langevin equations of motion
3290 for a rigid body of arbitrary shape in a velocity-Verlet style 2-part
3291 algorithm, where $h = \delta t$:
3292
3293 {\tt move A:}
3294 \begin{align*}
3295 {\bf v}\left(t + h / 2\right) &\leftarrow {\bf v}(t)
3296 + \frac{h}{2} \left( {\bf f}(t) / m \right), \\
3297 %
3298 {\bf r}(t + h) &\leftarrow {\bf r}(t)
3299 + h {\bf v}\left(t + h / 2 \right), \\
3300 %
3301 {\bf j}\left(t + h / 2 \right) &\leftarrow {\bf j}(t)
3302 + \frac{h}{2} {\bf \tau}^{~b}(t), \\
3303 %
3304 {\bf Q}(t + h) &\leftarrow \mathrm{rotate}\left( h {\bf j}
3305 (t + h / 2) \cdot \overleftrightarrow{\mathsf{I}}^{-1} \right).
3306 \end{align*}
3307 In this context, $\overleftrightarrow{\mathsf{I}}$ is the diagonal
3308 moment of inertia tensor, and the $\mathrm{rotate}$ function is the
3309 reversible product of the three body-fixed rotations,
3310 \begin{equation}
3311 \mathrm{rotate}({\bf a}) = \mathsf{G}_x(a_x / 2) \cdot
3312 \mathsf{G}_y(a_y / 2) \cdot \mathsf{G}_z(a_z) \cdot \mathsf{G}_y(a_y
3313 / 2) \cdot \mathsf{G}_x(a_x /2),
3314 \end{equation}
3315 where each rotational propagator, $\mathsf{G}_\alpha(\theta)$,
3316 rotates both the rotation matrix ($\mathbf{Q}$) and the body-fixed
3317 angular momentum (${\bf j}$) by an angle $\theta$ around body-fixed
3318 axis $\alpha$,
3319 \begin{equation}
3320 \mathsf{G}_\alpha( \theta ) = \left\{
3321 \begin{array}{lcl}
3322 \mathbf{Q}(t) & \leftarrow & \mathbf{Q}(0) \cdot \mathsf{R}_\alpha(\theta)^T, \\
3323 {\bf j}(t) & \leftarrow & \mathsf{R}_\alpha(\theta) \cdot {\bf
3324 j}(0).
3325 \end{array}
3326 \right.
3327 \end{equation}
3328 $\mathsf{R}_\alpha$ is a quadratic approximation to the single-axis
3329 rotation matrix. For example, in the small-angle limit, the
3330 rotation matrix around the body-fixed x-axis can be approximated as
3331 \begin{equation}
3332 \mathsf{R}_x(\theta) \approx \left(
3333 \begin{array}{ccc}
3334 1 & 0 & 0 \\
3335 0 & \frac{1-\theta^2 / 4}{1 + \theta^2 / 4} & -\frac{\theta}{1+
3336 \theta^2 / 4} \\
3337 0 & \frac{\theta}{1+ \theta^2 / 4} & \frac{1-\theta^2 / 4}{1 +
3338 \theta^2 / 4}
3339 \end{array}
3340 \right).
3341 \end{equation}
3342 All other rotations follow in a straightforward manner. After the
3343 first part of the propagation, the forces and body-fixed torques are
3344 calculated at the new positions and orientations. The system forces
3345 and torques are derivatives of the total potential energy function
3346 ($U$) with respect to the rigid body positions (${\bf r}$) and the
3347 columns of the transposed rotation matrix ${\bf Q}^T = \left({\bf
3348 u}_x, {\bf u}_y, {\bf u}_z \right)$:
3349
3350 {\tt Forces:}
3351 \begin{align*}
3352 {\bf f}_{s}(t + h) & \leftarrow
3353 - \left(\frac{\partial U}{\partial {\bf r}}\right)_{{\bf r}(t + h)} \\
3354 %
3355 {\bf \tau}_{s}(t + h) &\leftarrow {\bf u}(t + h)
3356 \times \frac{\partial U}{\partial {\bf u}} \\
3357 %
3358 {\bf v}^{b}_{R}(t+h) & \leftarrow \mathbf{Q}(t+h) \cdot \left({\bf v}(t+h) + {\bf \omega}(t+h) \times {\bf r}_{MR} \right) \\
3359 %
3360 {\bf f}_{R,f}^{b}(t+h) & \leftarrow - \Xi_{R}^{tt} \cdot
3361 {\bf v}^{b}_{R}(t+h) - \Xi_{R}^{rt} \cdot {\bf \omega}(t+h) \\
3362 %
3363 {\bf \tau}_{R,f}^{b}(t+h) & \leftarrow - \Xi_{R}^{tr} \cdot
3364 {\bf v}^{b}_{R}(t+h) - \Xi_{R}^{rr} \cdot {\bf \omega}(t+h) \\
3365 %
3366 Z & \leftarrow {\tt GaussianNormal}(2 k_B T / h, 6) \\
3367 {\bf F}_{R,r}^{b}(t+h) & \leftarrow {\bf S} \cdot Z \\
3368 %
3369 {\bf f}(t+h) & \leftarrow {\bf f}_{s}(t+h) + \mathbf{Q}^{T}(t+h)
3370 \cdot \left({\bf f}_{R,f}^{~b} + {\bf f}_{R,r}^{~b} \right) \\
3371 %
3372 \tau(t+h) & \leftarrow \tau_{s}(t+h) + \mathbf{Q}^{T}(t+h) \cdot \left(\tau_{R,f}^{~b} + \tau_{R,r}^{~b} \right) + {\bf r}_{MR} \times \left({\bf f}_{f}(t+h) + {\bf f}_{r}(t+h) \right) \\
3373 \tau^{~b}(t+h) & \leftarrow \mathbf{Q}(t+h) \cdot \tau(t+h) \\
3374 \end{align*}
3375 Frictional (and random) forces and torques must be computed at the
3376 center of resistance, so there are additional steps required to find
3377 the body-fixed velocity (${\bf v}_{R}^{~b}$) at this location. Mapping
3378 the frictional and random forces at the center of resistance back to
3379 the center of mass also introduces an additional term in the torque
3380 one obtains at the center of mass.
3381
3382 Once the forces and torques have been obtained at the new time step,
3383 the velocities can be advanced to the same time value.
3384
3385 {\tt move B:}
3386 \begin{align*}
3387 {\bf v}\left(t + h \right) &\leftarrow {\bf v}\left(t + h / 2
3388 \right)
3389 + \frac{h}{2} \left( {\bf f}(t + h) / m \right), \\
3390 %
3391 {\bf j}\left(t + h \right) &\leftarrow {\bf j}\left(t + h / 2
3392 \right)
3393 + \frac{h}{2} {\bf \tau}^{~b}(t + h) .
3394 \end{align*}
3395
3396 The viscosity of the implicit solvent must be specified using the {\tt
3397 viscosity} keyword in the meta-data file if the Langevin integrator is
3398 selected. For simple particles (spheres and ellipsoids), no further
3399 parameters are necessary. Since there are no analytic solutions for
3400 the resistance tensors for composite rigid bodies, the approximate
3401 tensors for these objects must also be specified in order to use
3402 Langevin dynamics. The meta-data file must therefore point to another
3403 file which contains the information about the hydrodynamic properties
3404 of all complex rigid bodies being used during the simulation. The
3405 {\tt HydroPropFile} keyword is used to specify the name of this
3406 file. A {\tt HydroPropFile} should be precalculated using the Hydro
3407 program that is included in the {\sc OpenMD} distribution.
3408
3409 \begin{longtable}[c]{ABG}
3410 \caption{Meta-data Keywords: Required parameters for the Langevin integrator}
3411 \\
3412 {\bf keyword} & {\bf units} & {\bf use} \\ \hline
3413 \endhead
3414 \hline
3415 \endfoot
3416 {\tt viscosity} & poise & Sets the value of viscosity of the implicit
3417 solvent \\
3418 {\tt targetTemp} & K & Sets the target temperature of the system.
3419 This parameter must be specified to use Langevin dynamics. \\
3420 {\tt HydroPropFile} & string & Specifies the name of the resistance
3421 tensor (usually a {\tt .diff} file) which is precalculated by {\tt
3422 Hydro}. This keyword is not necessary if the simulation contains only
3423 simple bodies (spheres and ellipsoids). \\
3424 {\tt beadSize} & $\mbox{\AA}$ & Sets the diameter of the beads to use
3425 when the {\tt RoughShell} model is used to approximate the resistance
3426 tensor.
3427 \label{table:ldParameters}
3428 \end{longtable}
3429
3430 \section{Constant Pressure without periodic boundary conditions (The LangevinHull)}
3431
3432 The Langevin Hull\cite{Vardeman2011} uses an external bath at a fixed constant pressure
3433 ($P$) and temperature ($T$) with an effective solvent viscosity
3434 ($\eta$). This bath interacts only with the objects on the exterior
3435 hull of the system. Defining the hull of the atoms in a simulation is
3436 done in a manner similar to the approach of Kohanoff, Caro and
3437 Finnis.\cite{Kohanoff:2005qm} That is, any instantaneous configuration
3438 of the atoms in the system is considered as a point cloud in three
3439 dimensional space. Delaunay triangulation is used to find all facets
3440 between coplanar
3441 neighbors.\cite{delaunay,springerlink:10.1007/BF00977785} In highly
3442 symmetric point clouds, facets can contain many atoms, but in all but
3443 the most symmetric of cases, the facets are simple triangles in
3444 3-space which contain exactly three atoms.
3445
3446 The convex hull is the set of facets that have {\it no concave
3447 corners} at an atomic site.\cite{Barber96,EDELSBRUNNER:1994oq} This
3448 eliminates all facets on the interior of the point cloud, leaving only
3449 those exposed to the bath. Sites on the convex hull are dynamic; as
3450 molecules re-enter the cluster, all interactions between atoms on that
3451 molecule and the external bath are removed. Since the edge is
3452 determined dynamically as the simulation progresses, no {\it a priori}
3453 geometry is defined. The pressure and temperature bath interacts only
3454 with the atoms on the edge and not with atoms interior to the
3455 simulation.
3456
3457 Atomic sites in the interior of the simulation move under standard
3458 Newtonian dynamics,
3459 \begin{equation}
3460 m_i \dot{\mathbf v}_i(t)=-{\mathbf \nabla}_i U,
3461 \label{eq:Newton}
3462 \end{equation}
3463 where $m_i$ is the mass of site $i$, ${\mathbf v}_i(t)$ is the
3464 instantaneous velocity of site $i$ at time $t$, and $U$ is the total
3465 potential energy. For atoms on the exterior of the cluster
3466 (i.e. those that occupy one of the vertices of the convex hull), the
3467 equation of motion is modified with an external force, ${\mathbf
3468 F}_i^{\mathrm ext}$:
3469 \begin{equation}
3470 m_i \dot{\mathbf v}_i(t)=-{\mathbf \nabla}_i U + {\mathbf F}_i^{\mathrm ext}.
3471 \end{equation}
3472
3473 The external bath interacts indirectly with the atomic sites through
3474 the intermediary of the hull facets. Since each vertex (or atom)
3475 provides one corner of a triangular facet, the force on the facets are
3476 divided equally to each vertex. However, each vertex can participate
3477 in multiple facets, so the resultant force is a sum over all facets
3478 $f$ containing vertex $i$:
3479 \begin{equation}
3480 {\mathbf F}_{i}^{\mathrm ext} = \sum_{\begin{array}{c}\mathrm{facets\
3481 } f \\ \mathrm{containing\ } i\end{array}} \frac{1}{3}\ {\mathbf
3482 F}_f^{\mathrm ext}
3483 \end{equation}
3484
3485 The external pressure bath applies a force to the facets of the convex
3486 hull in direct proportion to the area of the facet, while the thermal
3487 coupling depends on the solvent temperature, viscosity and the size
3488 and shape of each facet. The thermal interactions are expressed as a
3489 standard Langevin description of the forces,
3490 \begin{equation}
3491 \begin{array}{rclclcl}
3492 {\mathbf F}_f^{\text{ext}} & = & \text{external pressure} & + & \text{drag force} & + & \text{random force} \\
3493 & = & -\hat{n}_f P A_f & - & \Xi_f(t) {\mathbf v}_f(t) & + & {\mathbf R}_f(t)
3494 \end{array}
3495 \end{equation}
3496 Here, $A_f$ and $\hat{n}_f$ are the area and (outward-facing) normal
3497 vectors for facet $f$, respectively. ${\mathbf v}_f(t)$ is the
3498 velocity of the facet centroid,
3499 \begin{equation}
3500 {\mathbf v}_f(t) = \frac{1}{3} \sum_{i=1}^{3} {\mathbf v}_i,
3501 \end{equation}
3502 and $\Xi_f(t)$ is an approximate ($3 \times 3$) resistance tensor that
3503 depends on the geometry and surface area of facet $f$ and the
3504 viscosity of the bath. The resistance tensor is related to the
3505 fluctuations of the random force, $\mathbf{R}(t)$, by the
3506 fluctuation-dissipation theorem (see Eq. \ref{eq:randomForce}).
3507
3508 Once the resistance tensor is known for a given facet, a stochastic
3509 vector that has the properties in Eq. (\ref{eq:randomForce}) can be
3510 calculated efficiently by carrying out a Cholesky decomposition to
3511 obtain the square root matrix of the resistance tensor (see
3512 Eq. \ref{eq:Cholesky}).
3513
3514 Our treatment of the resistance tensor for the Langevin Hull facets is
3515 approximate. $\Xi_f$ for a rigid triangular plate would normally be
3516 treated as a $6 \times 6$ tensor that includes translational and
3517 rotational drag as well as translational-rotational coupling. The
3518 computation of resistance tensors for rigid bodies has been detailed
3519 elsewhere,\cite{JoseGarciadelaTorre02012000,Garcia-de-la-Torre:2001wd,GarciadelaTorreJ2002,Sun:2008fk}
3520 but the standard approach involving bead approximations would be
3521 prohibitively expensive if it were recomputed at each step in a
3522 molecular dynamics simulation.
3523
3524 Instead, we are utilizing an approximate resistance tensor obtained by
3525 first constructing the Oseen tensor for the interaction of the
3526 centroid of the facet ($f$) with each of the subfacets $\ell=1,2,3$,
3527 \begin{equation}
3528 T_{\ell f}=\frac{A_\ell}{8\pi\eta R_{\ell f}}\left(I +
3529 \frac{\mathbf{R}_{\ell f}\mathbf{R}_{\ell f}^T}{R_{\ell f}^2}\right)
3530 \end{equation}
3531 Here, $A_\ell$ is the area of subfacet $\ell$ which is a triangle
3532 containing two of the vertices of the facet along with the centroid.
3533 $\mathbf{R}_{\ell f}$ is the vector between the centroid of facet $f$
3534 and the centroid of sub-facet $\ell$, and $I$ is the ($3 \times 3$)
3535 identity matrix. $\eta$ is the viscosity of the external bath.
3536
3537 The tensors for each of the sub-facets are added together, and the
3538 resulting matrix is inverted to give a $3 \times 3$ resistance tensor
3539 for translations of the triangular facet,
3540 \begin{equation}
3541 \Xi_f(t) =\left[\sum_{i=1}^3 T_{if}\right]^{-1}.
3542 \end{equation}
3543 Note that this treatment ignores rotations (and
3544 translational-rotational coupling) of the facet. In compact systems,
3545 the facets stay relatively fixed in orientation between
3546 configurations, so this appears to be a reasonably good approximation.
3547
3548 At each
3549 molecular dynamics time step, the following process is carried out:
3550 \begin{enumerate}
3551 \item The standard inter-atomic forces ($\nabla_iU$) are computed.
3552 \item Delaunay triangulation is carried out using the current atomic
3553 configuration.
3554 \item The convex hull is computed and facets are identified.
3555 \item For each facet:
3556 \begin{itemize}
3557 \item[a.] The force from the pressure bath ($-\hat{n}_fPA_f$) is
3558 computed.
3559 \item[b.] The resistance tensor ($\Xi_f(t)$) is computed using the
3560 viscosity ($\eta$) of the bath.
3561 \item[c.] Facet drag ($-\Xi_f(t) \mathbf{v}_f(t)$) forces are
3562 computed.
3563 \item[d.] Random forces ($\mathbf{R}_f(t)$) are computed using the
3564 resistance tensor and the temperature ($T$) of the bath.
3565 \end{itemize}
3566 \item The facet forces are divided equally among the vertex atoms.
3567 \item Atomic positions and velocities are propagated.
3568 \end{enumerate}
3569 The Delaunay triangulation and computation of the convex hull are done
3570 using calls to the qhull library,\cite{Qhull} and for this reason, if
3571 qhull is not detected during the build, the Langevin Hull integrator
3572 will not be available. There is a minimal penalty for computing the
3573 convex hull and resistance tensors at each step in the molecular
3574 dynamics simulation (roughly 0.02 $\times$ cost of a single force
3575 evaluation).
3576
3577 \begin{longtable}[c]{GBF}
3578 \caption{Meta-data Keywords: Required parameters for the Langevin Hull integrator}
3579 \\
3580 {\bf keyword} & {\bf units} & {\bf use} \\ \hline
3581 \endhead
3582 \hline
3583 \endfoot
3584 {\tt viscosity} & poise & Sets the value of viscosity of the implicit
3585 solven . \\
3586 {\tt targetTemp} & K & Sets the target temperature of the system.
3587 This parameter must be specified to use Langevin Hull dynamics. \\
3588 {\tt targetPressure} & atm & Sets the target pressure of the system.
3589 This parameter must be specified to use Langevin Hull dynamics. \\
3590 {\tt usePeriodicBoundaryConditions} & logical & Turns off periodic boundary conditions.
3591 This parameter must be set to \tt false \\
3592 \label{table:lhullParameters}
3593 \end{longtable}
3594
3595
3596 \section{\label{sec:constraints}Constraint Methods}
3597
3598 \subsection{\label{section:rattle}The {\sc rattle} Method for Bond
3599 Constraints}
3600
3601 In order to satisfy the constraints of fixed bond lengths within {\sc
3602 OpenMD}, we have implemented the {\sc rattle} algorithm of
3603 Andersen.\cite{andersen83} {\sc rattle} is a velocity-Verlet
3604 formulation of the {\sc shake} method\cite{ryckaert77} for iteratively
3605 solving the Lagrange multipliers which maintain the holonomic
3606 constraints. Both methods are covered in depth in the
3607 literature,\cite{leach01:mm,Allen87} and a detailed description of
3608 this method would be redundant.
3609
3610 \subsection{\label{section:zcons}The Z-Constraint Method}
3611
3612 A force auto-correlation method based on the fluctuation-dissipation
3613 theorem was developed by Roux and Karplus to investigate the dynamics
3614 of ions inside ion channels.\cite{Roux91} The time-dependent friction
3615 coefficient can be calculated from the deviation of the instantaneous
3616 force from its mean value:
3617 \begin{equation}
3618 \xi(z,t)=\langle\delta F(z,t)\delta F(z,0)\rangle/k_{B}T,
3619 \end{equation}
3620 where%
3621 \begin{equation}
3622 \delta F(z,t)=F(z,t)-\langle F(z,t)\rangle.
3623 \end{equation}
3624
3625 If the time-dependent friction decays rapidly, the static friction
3626 coefficient can be approximated by
3627 \begin{equation}
3628 \xi_{\text{static}}(z)=\int_{0}^{\infty}\langle\delta F(z,t)\delta F(z,0)\rangle dt.
3629 \end{equation}
3630
3631 This allows the diffusion constant to then be calculated through the
3632 Einstein relation:\cite{Marrink94}
3633 \begin{equation}
3634 D(z)=\frac{k_{B}T}{\xi_{\text{static}}(z)}=\frac{(k_{B}T)^{2}}{\int_{0}^{\infty
3635 }\langle\delta F(z,t)\delta F(z,0)\rangle dt}.%
3636 \end{equation}
3637
3638 The Z-Constraint method, which fixes the $z$ coordinates of a few
3639 ``tagged'' molecules with respect to the center of the mass of the
3640 system is a technique that was proposed to obtain the forces required
3641 for the force auto-correlation calculation.\cite{Marrink94} However,
3642 simply resetting the coordinate will move the center of the mass of
3643 the whole system. To avoid this problem, we have developed a new
3644 method that is utilized in {\sc OpenMD}. Instead of resetting the
3645 coordinates, we reset the forces of $z$-constrained molecules and
3646 subtract the total constraint forces from the rest of the system after
3647 the force calculation at each time step.
3648
3649 After the force calculation, the total force on molecule $\alpha$ is:
3650 \begin{equation}
3651 G_{\alpha} = \sum_i F_{\alpha i},
3652 \label{eq:zc1}
3653 \end{equation}
3654 where $F_{\alpha i}$ is the force in the $z$ direction on atom $i$ in
3655 $z$-constrained molecule $\alpha$. The forces on the atoms in the
3656 $z$-constrained molecule are then adjusted to remove the total force
3657 on molecule $\alpha$:
3658 \begin{equation}
3659 F_{\alpha i} = F_{\alpha i} -
3660 \frac{m_{\alpha i} G_{\alpha}}{\sum_i m_{\alpha i}}.
3661 \end{equation}
3662 Here, $m_{\alpha i}$ is the mass of atom $i$ in the $z$-constrained
3663 molecule. After the forces have been adjusted, the velocities must
3664 also be modified to subtract out molecule $\alpha$'s center-of-mass
3665 velocity in the $z$ direction.
3666 \begin{equation}
3667 v_{\alpha i} = v_{\alpha i} -
3668 \frac{\sum_i m_{\alpha i} v_{\alpha i}}{\sum_i m_{\alpha i}},
3669 \end{equation}
3670 where $v_{\alpha i}$ is the velocity of atom $i$ in the $z$ direction.
3671 Lastly, all of the accumulated constraint forces must be subtracted
3672 from the rest of the unconstrained system to keep the system center of
3673 mass of the entire system from drifting.
3674 \begin{equation}
3675 F_{\beta i} = F_{\beta i} - \frac{m_{\beta i} \sum_{\alpha} G_{\alpha}}
3676 {\sum_{\beta}\sum_i m_{\beta i}},
3677 \end{equation}
3678 where $\beta$ denotes all {\it unconstrained} molecules in the
3679 system. Similarly, the velocities of the unconstrained molecules must
3680 also be scaled:
3681 \begin{equation}
3682 v_{\beta i} = v_{\beta i} + \sum_{\alpha} \frac{\sum_i m_{\alpha i}
3683 v_{\alpha i}}{\sum_i m_{\alpha i}}.
3684 \end{equation}
3685
3686 This method will pin down the centers-of-mass of all of the
3687 $z$-constrained molecules, and will also keep the entire system fixed
3688 at the original system center-of-mass location.
3689
3690 At the very beginning of the simulation, the molecules may not be at
3691 their desired positions. To steer a $z$-constrained molecule to its
3692 specified position, a simple harmonic potential is used:
3693 \begin{equation}
3694 U(t)=\frac{1}{2}k_{\text{Harmonic}}(z(t)-z_{\text{cons}})^{2},%
3695 \end{equation}
3696 where $k_{\text{Harmonic}}$ is an harmonic force constant, $z(t)$ is
3697 the current $z$ coordinate of the center of mass of the constrained
3698 molecule, and $z_{\text{cons}}$ is the desired constraint
3699 position. The harmonic force operating on the $z$-constrained molecule
3700 at time $t$ can be calculated by
3701 \begin{equation}
3702 F_{z_{\text{Harmonic}}}(t)=-\frac{\partial U(t)}{\partial z(t)}=
3703 -k_{\text{Harmonic}}(z(t)-z_{\text{cons}}).
3704 \end{equation}
3705
3706 The user may also specify the use of a constant velocity method
3707 (steered molecular dynamics) to move the molecules to their desired
3708 initial positions. Based on concepts from atomic force microscopy,
3709 {\sc smd} has been used to study many processes which occur via rare
3710 events on the time scale of a few hundreds of picoseconds. For
3711 example,{\sc smd} has been used to observe the dissociation of
3712 Streptavidin-biotin Complex.\cite{smd}
3713
3714 To use of the $z$-constraint method in an {\sc OpenMD} simulation, the
3715 molecules must be specified using the {\tt nZconstraints} keyword in
3716 the meta-data file. The other parameters for modifying the behavior
3717 of the $z$-constraint method are listed in table~\ref{table:zconParams}.
3718
3719 \begin{longtable}[c]{ABCD}
3720 \caption{Meta-data Keywords: Z-Constraint Parameters}
3721 \\
3722 {\bf keyword} & {\bf units} & {\bf use} & {\bf remarks} \\ \hline
3723 \endhead
3724 \hline
3725 \endfoot
3726 {\tt zconsTime} & fs & Sets the frequency at which the {\tt .fz} file
3727 is written & \\
3728 {\tt zconsForcePolicy} & string & The strategy for subtracting
3729 the $z$-constraint force from the {\it unconstrained} molecules & Possible
3730 strategies are {\tt BYMASS} and {\tt BYNUMBER}. The default
3731 strategy is {\tt BYMASS}\\
3732 {\tt zconsGap} & $\mbox{\AA}$ & Sets the distance between two adjacent
3733 constraint positions&Used mainly to move molecules through a
3734 simulation to estimate potentials of mean force. \\
3735 {\tt zconsFixtime} & fs & Sets the length of time the $z$-constraint
3736 molecule is held fixed & {\tt zconsFixtime} must be set if {\tt
3737 zconsGap} is set\\
3738 {\tt zconsUsingSMD} & logical & Flag for using Steered Molecular
3739 Dynamics to move the molecules to the correct constrained positions &
3740 Harmonic Forces are used by default
3741 \label{table:zconParams}
3742 \end{longtable}
3743
3744 % \chapter{\label{section:restraints}Restraints}
3745 % Restraints are external potentials that are added to a system to
3746 % keep particular molecules or collections of particles close to some
3747 % reference structure. A restraint can be a collective
3748
3749 \chapter{\label{section:thermInt}Thermodynamic Integration}
3750
3751 Thermodynamic integration is an established technique that has been
3752 used extensively in the calculation of free energies for condensed
3753 phases of
3754 materials.\cite{Frenkel84,Hermens88,Meijer90,Baez95a,Vlot99}. This
3755 method uses a sequence of simulations during which the system of
3756 interest is converted into a reference system for which the free
3757 energy is known analytically ($A_0$). The difference in potential
3758 energy between the reference system and the system of interest
3759 ($\Delta V$) is then integrated in order to determine the free energy
3760 difference between the two states:
3761 \begin{equation}
3762 A = A_0 + \int_0^1 \left\langle \Delta V \right\rangle_\lambda
3763 d\lambda.
3764 \label{eq:thermInt}
3765 \end{equation}
3766 Here, $\lambda$ is the parameter that governs the transformation
3767 between the reference system and the system of interest. For
3768 crystalline phases, an harmonically-restrained (Einstein) crystal is
3769 chosen as the reference state, while for liquid phases, the ideal gas
3770 is taken as the reference state.
3771
3772 In an Einstein crystal, the molecules are restrained at their ideal
3773 lattice locations and orientations. Using harmonic restraints, as
3774 applied by B\`{a}ez and Clancy, the total potential for this reference
3775 crystal ($V_\mathrm{EC}$) is the sum of all the harmonic restraints,
3776 \begin{equation}
3777 V_\mathrm{EC} = \sum_{i} \left[ \frac{K_\mathrm{v}}{2} (r_i - r_i^\circ)^2 +
3778 \frac{K_\theta}{2} (\theta_i - \theta_i^\circ)^2 +
3779 \frac{K_\omega}{2}(\omega_i - \omega_i^\circ)^2 \right],
3780 \end{equation}
3781 where $K_\mathrm{v}$, $K_\mathrm{\theta}$, and $K_\mathrm{\omega}$ are
3782 the spring constants restraining translational motion and deflection
3783 of and rotation around the principle axis of the molecule
3784 respectively. The values of $\theta$ range from $0$ to $\pi$, while
3785 $\omega$ ranges from $-\pi$ to $\pi$.
3786
3787 The partition function for a molecular crystal restrained in this
3788 fashion can be evaluated analytically, and the Helmholtz Free Energy
3789 ({\it A}) is given by
3790 \begin{eqnarray}
3791 \frac{A}{N} = \frac{E_m}{N}\ -\ kT\ln \left (\frac{kT}{h\nu}\right )^3&-&kT\ln \left
3792 [\pi^\frac{1}{2}\left (\frac{8\pi^2I_\mathrm{A}kT}{h^2}\right
3793 )^\frac{1}{2}\left (\frac{8\pi^2I_\mathrm{B}kT}{h^2}\right
3794 )^\frac{1}{2}\left (\frac{8\pi^2I_\mathrm{C}kT}{h^2}\right
3795 )^\frac{1}{2}\right ] \nonumber \\ &-& kT\ln \left [\frac{kT}{2(\pi
3796 K_\omega K_\theta)^{\frac{1}{2}}}\exp\left
3797 (-\frac{kT}{2K_\theta}\right)\int_0^{\left (\frac{kT}{2K_\theta}\right
3798 )^\frac{1}{2}}\exp(t^2)\mathrm{d}t\right ],
3799 \label{ecFreeEnergy}
3800 \end{eqnarray}
3801 where $2\pi\nu = (K_\mathrm{v}/m)^{1/2}$, and $E_m$ is the minimum
3802 potential energy of the ideal crystal.\cite{Baez95a}
3803
3804 {\sc OpenMD} can perform the simulations that aid the user in
3805 constructing the thermodynamic path from the molecular system to one
3806 of the reference systems. To do this, the user sets the value of
3807 $\lambda$ (between 0 \& 1) in the meta-data file. If the system of
3808 interest is crystalline, {\sc OpenMD} must be able to find the {\it
3809 reference} configuration of the system in a file called {\tt
3810 idealCrystal.in} in the directory from which the simulation was run.
3811 This file is a standard {\tt .dump} file, but all information about
3812 velocities and angular momenta are discarded when the file is read.
3813
3814 The configuration found in the {\tt idealCrystal.in} file is used for
3815 the reference positions and molecular orientations of the Einstein
3816 crystal. To complete the specification of the Einstein crystal, a set
3817 of force constants must also be specified; one for displacments of the
3818 molecular centers of mass, and two for displacements from the ideal
3819 orientations of the molecules.
3820
3821 To construct a thermodynamic integration path, the user would run a
3822 sequence of $N$ simulations, each with a different value of lambda
3823 between $0$ and $1$. When {\tt useSolidThermInt} is set to {\tt true}
3824 in the meta-data file, two additional energy columns are reported in
3825 the {\tt .stat} file for the simulation. The first, {\tt vRaw}, is
3826 the unperturbed energy for the configuration, and the second, {\tt
3827 vHarm}, is the energy of the harmonic (Einstein) system in an
3828 identical configuration. The total potential energy of the
3829 configuration is a linear combination of {\tt vRaw} and {\tt vHarm}
3830 weighted by the value of $\lambda$.
3831
3832 From a running average of the difference between {\tt vRaw} and {\tt
3833 vHarm}, the user can obtain the integrand in Eq. (\ref{eq:thermInt})
3834 for fixed value of $\lambda$.
3835
3836 There are two additional files with the suffixes {\tt .zang0} and {\tt
3837 .zang} generated by {\sc OpenMD} during the first run of a solid
3838 thermodynamic integration. These files contain the initial ({\tt
3839 .zang0}) and final ({\tt .zang}) values of the angular displacement
3840 coordinates for each of the molecules. These are particularly useful
3841 when chaining a number of simulations (with successive values of
3842 $\lambda$) together.
3843
3844 For {\it liquid} thermodynamic integrations, the reference system is
3845 the ideal gas (with a potential exactly equal to 0), so the {\tt
3846 .stat} file contains only the standard columns. The potential energy
3847 column contains the potential of the {\it unperturbed} system (and not
3848 the $\lambda$-weighted potential. This allows the user to use the
3849 potential energy directly as the $\Delta V$ in the integrand of
3850 Eq. (\ref{eq:thermInt}).
3851
3852 Meta-data parameters concerning thermodynamic integrations are given in
3853 Table~\ref{table:thermIntParams}
3854
3855 \begin{longtable}[c]{ABCD}
3856 \caption{Meta-data Keywords: Thermodynamic Integration Parameters}
3857 \\
3858 {\bf keyword} & {\bf units} & {\bf use} & {\bf remarks} \\ \hline
3859 \endhead
3860 \hline
3861 \endfoot
3862 {\tt useSolidThermInt} & logical & perform thermodynamic integration
3863 to an Einstein crystal? & default is ``false'' \\
3864 {\tt useLiquidThermInt} & logical & perform thermodynamic integration
3865 to an ideal gas? & default is ``false'' \\
3866 {\tt thermodynamicIntegrationLambda} & & & \\
3867 & double & transformation
3868 parameter & Sets how far along the thermodynamic integration path the
3869 simulation will be. \\
3870 {\tt thermodynamicIntegrationK} & & & \\
3871 & double & & power of $\lambda$
3872 governing shape of integration pathway \\
3873 {\tt thermIntDistSpringConst} & & & \\
3874 & $\mbox{kcal~mol}^{-1} \mbox{\AA}^{-2}$
3875 & & spring constant for translations in Einstein crystal \\
3876 {\tt thermIntThetaSpringConst} & & & \\
3877 & $\mbox{kcal~mol}^{-1}
3878 \mbox{rad}^{-2}$ & & spring constant for deflection away from z-axis
3879 in Einstein crystal \\
3880 {\tt thermIntOmegaSpringConst} & & & \\
3881 & $\mbox{kcal~mol}^{-1}
3882 \mbox{rad}^{-2}$ & & spring constant for rotation around z-axis in
3883 Einstein crystal
3884 \label{table:thermIntParams}
3885 \end{longtable}
3886
3887 \chapter{\label{section:rnemd}Reverse Non-Equilibrium Molecular Dynamics (RNEMD)}
3888
3889 There are many ways to compute transport properties from molecular
3890 dynamics simulations. Equilibrium Molecular Dynamics (EMD)
3891 simulations can be used by computing relevant time correlation
3892 functions and assuming linear response theory holds. For some transport properties (notably thermal conductivity), EMD approaches
3893 are subject to noise and poor convergence of the relevant
3894 correlation functions. Traditional Non-equilibrium Molecular Dynamics
3895 (NEMD) methods impose a gradient (e.g. thermal or momentum) on a
3896 simulation. However, the resulting flux is often difficult to
3897 measure. Furthermore, problems arise for NEMD simulations of
3898 heterogeneous systems, such as phase-phase boundaries or interfaces,
3899 where the type of gradient to enforce at the boundary between
3900 materials is unclear.
3901
3902 {\it Reverse} Non-Equilibrium Molecular Dynamics (RNEMD) methods adopt
3903 a different approach in that an unphysical {\it flux} is imposed
3904 between different regions or ``slabs'' of the simulation box. The
3905 response of the system is to develop a temperature or momentum {\it
3906 gradient} between the two regions. Since the amount of the applied
3907 flux is known exactly, and the measurement of gradient is generally
3908 less complicated, imposed-flux methods typically take shorter
3909 simulation times to obtain converged results for transport properties.
3910
3911 \begin{figure}
3912 \includegraphics[width=\linewidth]{rnemdDemo}
3913 \caption{The (VSS) RNEMD approach imposes unphysical transfer of both
3914 linear momentum and kinetic energy between a ``hot'' slab and a
3915 ``cold'' slab in the simulation box. The system responds to this
3916 imposed flux by generating both momentum and temperature gradients.
3917 The slope of the gradients can then be used to compute transport
3918 properties (e.g. shear viscosity and thermal conductivity).}
3919 \label{rnemdDemo}
3920 \end{figure}
3921
3922 \section{\label{section:algo}Three algorithms for carrying out RNEMD simulations}
3923 \subsection{\label{subsection:swapping}The swapping algorithm}
3924 The original ``swapping'' approaches by M\"{u}ller-Plathe {\it et
3925 al.}\cite{ISI:000080382700030,MullerPlathe:1997xw} can be understood
3926 as a sequence of imaginary elastic collisions between particles in
3927 opposite slabs. In each collision, the entire momentum vectors of
3928 both particles may be exchanged to generate a thermal
3929 flux. Alternatively, a single component of the momentum vectors may be
3930 exchanged to generate a shear flux. This algorithm turns out to be
3931 quite useful in many simulations. However, the M\"{u}ller-Plathe
3932 swapping approach perturbs the system away from ideal
3933 Maxwell-Boltzmann distributions, and this may leads to undesirable
3934 side-effects when the applied flux becomes large.\cite{Maginn:2010}
3935 This limits the applicability of the swapping algorithm, so in OpenMD,
3936 we have implemented two additional algorithms for RNEMD in addition to the
3937 original swapping approach.
3938
3939 \subsection{\label{subsection:nivs}Non-Isotropic Velocity Scaling (NIVS)}
3940 Instead of having momentum exchange between {\it individual particles}
3941 in each slab, the NIVS algorithm applies velocity scaling to all of
3942 the selected particles in both slabs.\cite{kuang:164101} A combination of linear
3943 momentum, kinetic energy, and flux constraint equations governs the
3944 amount of velocity scaling performed at each step. Interested readers
3945 should consult ref. \citealp{kuang:164101} for further details on the
3946 methodology.
3947
3948 NIVS has been shown to be very effective at producing thermal
3949 gradients and for computing thermal conductivities, particularly for
3950 heterogeneous interfaces. Although the NIVS algorithm can also be
3951 applied to impose a directional momentum flux, thermal anisotropy was
3952 observed in relatively high flux simulations, and the method is not
3953 suitable for imposing a shear flux or for computing shear viscosities.
3954
3955 \subsection{\label{subsection:vss}Velocity Shearing and Scaling (VSS)}
3956 The third RNEMD algorithm implemented in OpenMD utilizes a series of
3957 simultaneous velocity shearing and scaling exchanges between the two
3958 slabs.\cite{2012MolPh.110..691K} This method results in a set of simpler equations to satisfy
3959 the conservation constraints while creating a desired flux between the
3960 two slabs.
3961
3962 The VSS approach is versatile in that it may be used to implement both
3963 thermal and shear transport either separately or simultaneously.
3964 Perturbations of velocities away from the ideal Maxwell-Boltzmann
3965 distributions are minimal, and thermal anisotropy is kept to a
3966 minimum. This ability to generate simultaneous thermal and shear
3967 fluxes has been utilized to map out the shear viscosity of SPC/E water
3968 over a wide range of temperatures (90~K) just with a single simulation.
3969 VSS-RNEMD also allows the directional momentum flux to have
3970 arbitary directions, which could aid in the study of anisotropic solid
3971 surfaces in contact with liquid environments.
3972
3973 \section{\label{section:usingRNEMD}Using OpenMD to perform a RNEMD simulation}
3974 \subsection{\label{section:rnemdParams} What the user needs to specify}
3975 To carry out a RNEMD simulation,
3976 a user must specify a number of parameters in the MetaData (.md) file.
3977 Because the RNEMD methods have a large number of parameters, these
3978 must be enclosed in a {\it separate} {\tt RNEMD\{...\}} block. The most important
3979 parameters to specify are the {\tt useRNEMD}, {\tt fluxType} and flux
3980 parameters. Most other parameters (summarized in table
3981 \ref{table:rnemd}) have reasonable default values. {\tt fluxType}
3982 sets up the kind of exchange that will be carried out between the two
3983 slabs (either Kinetic Energy ({\tt KE}) or momentum ({\tt Px, Py, Pz,
3984 Pvector}), or some combination of these). The flux is specified
3985 with the use of three possible parameters: {\tt kineticFlux} for
3986 kinetic energy exchange, as well as {\tt momentumFlux} or {\tt
3987 momentumFluxVector} for simulations with directional exchange.
3988
3989 \subsection{\label{section:rnemdResults} Processing the results}
3990 OpenMD will generate a {\tt .rnemd}
3991 file with the same prefix as the original {\tt .md} file. This file
3992 contains a running average of properties of interest computed within a
3993 set of bins that divide the simulation cell along the $z$-axis. The
3994 first column of the {\tt .rnemd} file is the $z$ coordinate of the
3995 center of each bin, while following columns may contain the average
3996 temperature, velocity, or density within each bin. The output format
3997 in the {\tt .rnemd} file can be altered with the {\tt outputFields},
3998 {\tt outputBins}, and {\tt outputFileName} parameters. A report at
3999 the top of the {\tt .rnemd} file contains the current exchange totals
4000 as well as the average flux applied during the simulation. Using the
4001 slope of the temperature or velocity gradient obtaine from the {\tt
4002 .rnemd} file along with the applied flux, the user can very simply
4003 arrive at estimates of thermal conductivities ($\lambda$),
4004 \begin{equation}
4005 J_z = -\lambda \frac{\partial T}{\partial z},
4006 \end{equation}
4007 and shear viscosities ($\eta$),
4008 \begin{equation}
4009 j_z(p_x) = -\eta \frac{\partial \langle v_x \rangle}{\partial z}.
4010 \end{equation}
4011 Here, the quantities on the left hand side are the actual flux values
4012 (in the header of the {\tt .rnemd} file), while the slopes are
4013 obtained from linear fits to the gradients observed in the {\tt
4014 .rnemd} file.
4015
4016 More complicated simulations (including interfaces) require a bit more
4017 care. Here the second derivative may be required to compute the
4018 interfacial thermal conductance,
4019 \begin{align}
4020 G^\prime &= \left(\nabla\lambda \cdot \mathbf{\hat{n}}\right)_{z_0} \\
4021 &= \frac{\partial}{\partial z}\left(-\frac{J_z}{
4022 \left(\frac{\partial T}{\partial z}\right)}\right)_{z_0} \\
4023 &= J_z\left(\frac{\partial^2 T}{\partial z^2}\right)_{z_0} \Big/
4024 \left(\frac{\partial T}{\partial z}\right)_{z_0}^2.
4025 \label{derivativeG}
4026 \end{align}
4027 where $z_0$ is the location of the interface between two materials and
4028 $\mathbf{\hat{n}}$ is a unit vector normal to the interface. We
4029 suggest that users interested in interfacial conductance consult
4030 reference \citealp{kuang:AuThl} for other approaches to computing $G$.
4031 Users interested in {\it friction coefficients} at heterogeneous
4032 interfaces may also find reference \citealp{2012MolPh.110..691K}
4033 useful.
4034
4035 \newpage
4036
4037 \begin{longtable}[c]{JKLM}
4038 \caption{Meta-data Keywords: Parameters for RNEMD simulations}\\
4039 \multicolumn{4}{c}{The following keywords must be enclosed inside a {\tt RNEMD\{...\}} block.}
4040 \\ \hline
4041 {\bf keyword} & {\bf units} & {\bf use} & {\bf remarks} \\ \hline
4042 \endhead
4043 \hline
4044 \endfoot
4045 {\tt useRNEMD} & logical & perform RNEMD? & default is ``false'' \\
4046 {\tt objectSelection} & string & see section \ref{section:syntax}
4047 for selection syntax & default is ``select all'' \\
4048 {\tt method} & string & exchange method & one of the following:
4049 {\tt Swap, NIVS,} or {\tt VSS} (default is {\tt VSS}) \\
4050 {\tt fluxType} & string & what is being exchanged between slabs? & one
4051 of the following: {\tt KE, Px, Py, Pz, Pvector, KE+Px, KE+Py, KE+Pvector} \\
4052 {\tt kineticFlux} & kcal mol$^{-1}$ \AA$^{-2}$ fs$^{-1}$ & specify the kinetic energy flux & \\
4053 {\tt momentumFlux} & amu \AA$^{-1}$ fs$^{-2}$ & specify the momentum flux & \\
4054 {\tt momentumFluxVector} & amu \AA$^{-1}$ fs$^{-2}$ & specify the momentum flux when
4055 {\tt Pvector} is part of the exchange & Vector3d input\\
4056 {\tt exchangeTime} & fs & how often to perform the exchange & default is 100 fs\\
4057
4058 {\tt slabWidth} & $\mbox{\AA}$ & width of the two exchange slabs & default is $\mathsf{H}_{zz} / 10.0$ \\
4059 {\tt slabAcenter} & $\mbox{\AA}$ & center of the end slab & default is 0 \\
4060 {\tt slabBcenter} & $\mbox{\AA}$ & center of the middle slab & default is $\mathsf{H}_{zz} / 2$ \\
4061 {\tt outputFileName} & string & file name for output histograms & default is the same prefix as the
4062 .md file, but with the {\tt .rnemd} extension \\
4063 {\tt outputBins} & int & number of $z$-bins in the output histogram &
4064 default is 20 \\
4065 {\tt outputFields} & string & columns to print in the {\tt .rnemd}
4066 file where each column is separated by a pipe ($\mid$) symbol. & Allowed column names are: {\sc z, temperature, velocity, density} \\
4067 \label{table:rnemd}
4068 \end{longtable}
4069
4070 \chapter{\label{section:minimizer}Energy Minimization}
4071
4072 Energy minimization is used to identify local configurations that are stable
4073 points on the potential energy surface. There is a vast literature on
4074 energy minimization algorithms have been developed to search for the
4075 global energy minimum as well as to find local structures which are
4076 stable fixed points on the surface. We have included two simple
4077 minimization algorithms: steepest descent, ({\sc sd}) and conjugate
4078 gradient ({\sc cg}) to help users find reasonable local minima from
4079 their initial configurations. Since {\sc OpenMD} handles atoms and
4080 rigid bodies which have orientational coordinates as well as
4081 translational coordinates, there is some subtlety to the choice of
4082 parameters for minimization algorithms.
4083
4084 Given a coordinate set $x_{k}$ and a search direction $d_{k}$, a line
4085 search algorithm is performed along $d_{k}$ to produce
4086 $x_{k+1}=x_{k}+$ $\lambda _{k}d_{k}$. In the steepest descent ({\sc
4087 sd}) algorithm,%
4088 \begin{equation}
4089 d_{k}=-\nabla V(x_{k}).
4090 \end{equation}
4091 The gradient and the direction of next step are always orthogonal.
4092 This may cause oscillatory behavior in narrow valleys. To overcome
4093 this problem, the Fletcher-Reeves variant~\cite{FletcherReeves} of the
4094 conjugate gradient ({\sc cg}) algorithm is used to generate $d_{k+1}$
4095 via simple recursion:
4096 \begin{equation}
4097 d_{k+1} =-\nabla V(x_{k+1})+\gamma_{k}d_{k}
4098 \end{equation}
4099 where
4100 \begin{equation}
4101 \gamma_{k} =\frac{\nabla V(x_{k+1})^{T}\nabla V(x_{k+1})}{\nabla
4102 V(x_{k})^{T}\nabla V(x_{k})}.
4103 \end{equation}
4104
4105 The Polak-Ribiere variant~\cite{PolakRibiere} of the conjugate
4106 gradient ($\gamma_{k}$) is defined as%
4107 \begin{equation}
4108 \gamma_{k}=\frac{[\nabla V(x_{k+1})-\nabla V(x)]^{T}\nabla V(x_{k+1})}{\nabla
4109 V(x_{k})^{T}\nabla V(x_{k})}%
4110 \end{equation}
4111 It is widely agreed that the Polak-Ribiere variant gives better
4112 convergence than the Fletcher-Reeves variant, so the conjugate
4113 gradient approach implemented in {\sc OpenMD} is the Polak-Ribiere
4114 variant.
4115
4116 The conjugate gradient method assumes that the conformation is close
4117 enough to a local minimum that the potential energy surface is very
4118 nearly quadratic. When the initial structure is far from the minimum,
4119 the steepest descent method can be superior to the conjugate gradient
4120 method. Hence, the steepest descent method is often used for the first
4121 10-100 steps of minimization. Another useful feature of minimization
4122 methods in {\sc OpenMD} is that a modified {\sc shake} algorithm can be
4123 applied during the minimization to constraint the bond lengths if this
4124 is required by the force field. Meta-data parameters concerning the
4125 minimizer are given in Table~\ref{table:minimizeParams}
4126
4127 \begin{longtable}[c]{ABCD}
4128 \caption{Meta-data Keywords: Energy Minimizer Parameters}
4129 \\
4130 {\bf keyword} & {\bf units} & {\bf use} & {\bf remarks} \\ \hline
4131 \endhead
4132 \hline
4133 \endfoot
4134 {\tt minimizer} & string & selects the minimization method to be used
4135 & either {\tt CG} (conjugate gradient) or {\tt SD} (steepest
4136 descent) \\
4137 {\tt minimizerMaxIter} & steps & Sets the maximum number of iterations
4138 for the energy minimization & The default value is 200\\
4139 {\tt minimizerWriteFreq} & steps & Sets the frequency with which the {\tt .dump} and {\tt .stat} files are writtern during energy minimization & \\
4140 {\tt minimizerStepSize} & $\mbox{\AA}$ & Sets the step size for the
4141 line search & The default value is 0.01\\
4142 {\tt minimizerFTol} & $\mbox{kcal mol}^{-1}$ & Sets the energy tolerance
4143 for stopping the minimziation. & The default value is $10^{-8}$\\
4144 {\tt minimizerGTol} & $\mbox{kcal mol}^{-1}\mbox{\AA}^{-1}$ & Sets the
4145 gradient tolerance for stopping the minimization. & The default value
4146 is $10^{-8}$\\
4147 {\tt minimizerLSTol} & $\mbox{kcal mol}^{-1}$ & Sets line search
4148 tolerance for terminating each step of the minimization. & The default
4149 value is $10^{-8}$\\
4150 {\tt minimizerLSMaxIter} & steps & Sets the maximum number of
4151 iterations for each line search & The default value is 50\\
4152 \label{table:minimizeParams}
4153 \end{longtable}
4154
4155 \chapter{\label{section:anal}Analysis of Physical Properties}
4156
4157 {\sc OpenMD} includes a few utility programs which compute properties
4158 from the dump files that are generated during a molecular dynamics
4159 simulation. These programs are:
4160
4161 \begin{description}
4162 \item[{\bf Dump2XYZ}] Converts an {\sc OpenMD} dump file into a file
4163 suitable for viewing in a molecular dynamics viewer like Jmol
4164 \item[{\bf StaticProps}] Computes static properties like the pair
4165 distribution function, $g(r)$.
4166 \item[{\bf DynamicProps}] Computes time correlation functions like the
4167 velocity autocorrelation function, $\langle v(t) \cdot v(0)\rangle$,
4168 or the mean square displacement $\langle |r(t) - r(0)|^{2} \rangle$.
4169 \end{description}
4170
4171 These programs often need to operate on a subset of the data contained
4172 within a dump file. For example, if you want only the {\it oxygen-oxygen}
4173 pair distribution from a water simulation, or if you want to make a
4174 movie including only the water molecules within a 6 angstrom radius of
4175 lipid head groups, you need a way to specify your selection to these
4176 utility programs. {\sc OpenMD} has a selection syntax which allows you to
4177 specify the selection in a compact form in order to generate only the
4178 data you want. For example a common use of the StaticProps command
4179 would be:
4180
4181 {\tt StaticProps -i tp4.dump -{}-gofr -{}-sele1="select O*" -{}-sele2="select O*"}
4182
4183 This command computes the oxygen-oxygen pair distribution function,
4184 $g_{OO}(r)$, from a file named {\tt tp4.dump}. In order to understand
4185 this selection syntax and to make full use of the selection
4186 capabilities of the analysis programs, it is necessary to understand a
4187 few of the core concepts that are used to perform simulations.
4188
4189 \section{\label{section:concepts}Concepts}
4190
4191 {\sc OpenMD} manipulates both traditional atoms as well as some objects that
4192 {\it behave like atoms}. These objects can be rigid collections of
4193 atoms or atoms which have orientational degrees of freedom. Here is a
4194 diagram of the class heirarchy:
4195
4196 \begin{figure}
4197 \centering
4198 \includegraphics[width=3in]{heirarchy.pdf}
4199 \caption[Class heirarchy for StuntDoubles in {\sc OpenMD}]{ \\ The
4200 class heirarchy of StuntDoubles in {\sc OpenMD}. The selection
4201 syntax allows the user to select any of the objects that are descended
4202 from a StuntDouble.}
4203 \label{fig:heirarchy}
4204 \end{figure}
4205
4206 \begin{itemize}
4207 \item A {\bf StuntDouble} is {\it any} object that can be manipulated by the
4208 integrators and minimizers.
4209 \item An {\bf Atom} is a fundamental point-particle that can be moved around during a simulation.
4210 \item A {\bf DirectionalAtom} is an atom which has {\it orientational} as well as translational degrees of freedom.
4211 \item A {\bf RigidBody} is a collection of {\bf Atom}s or {\bf
4212 DirectionalAtom}s which behaves as a single unit.
4213 \end{itemize}
4214
4215 Every Molecule, Atom and DirectionalAtom in {\sc OpenMD} have their own names
4216 which are specified in the {\tt .md} file. In contrast, RigidBodies are
4217 denoted by their membership and index inside a particular molecule:
4218 [MoleculeName]\_RB\_[index] (the contents inside the brackets
4219 depend on the specifics of the simulation). The names of rigid bodies are
4220 generated automatically. For example, the name of the first rigid body
4221 in a DMPC molecule is DMPC\_RB\_0.
4222
4223 \section{\label{section:syntax}Syntax of the Select Command}
4224
4225 The most general form of the select command is: {\tt select {\it expression}}
4226
4227 This expression represents an arbitrary set of StuntDoubles (Atoms or
4228 RigidBodies) in {\sc OpenMD}. Expressions are composed of either name
4229 expressions, index expressions, predefined sets, user-defined
4230 expressions, comparison operators, within expressions, or logical
4231 combinations of the above expression types. Expressions can be
4232 combined using parentheses and the Boolean operators.
4233
4234 \subsection{\label{section:logical}Logical expressions}
4235
4236 The logical operators allow complex queries to be constructed out of
4237 simpler ones using the standard boolean connectives {\bf and}, {\bf
4238 or}, {\bf not}. Parentheses can be used to alter the precedence of the
4239 operators.
4240
4241 \begin{center}
4242 \begin{tabular}{|ll|}
4243 \hline
4244 {\bf logical operator} & {\bf equivalent operator} \\
4245 \hline
4246 and & ``\&'', ``\&\&'' \\
4247 or & ``$|$'', ``$||$'', ``,'' \\
4248 not & ``!'' \\
4249 \hline
4250 \end{tabular}
4251 \end{center}
4252
4253 \subsection{\label{section:name}Name expressions}
4254
4255 \begin{center}
4256 \begin{tabular}{|llp{3in}|}
4257 \hline
4258 {\bf type of expression} & {\bf examples} & {\bf translation of
4259 examples} \\
4260 \hline
4261 expression without ``.'' & select DMPC & select all StuntDoubles
4262 belonging to all DMPC molecules \\
4263 & select C* & select all atoms which have atom types beginning with C
4264 \\
4265 & select DMPC\_RB\_* & select all RigidBodies in DMPC molecules (but
4266 only select the rigid bodies, and not the atoms belonging to them). \\
4267 \hline
4268 expression has one ``.'' & select TIP3P.O\_TIP3P & select the O\_TIP3P
4269 atoms belonging to TIP3P molecules \\
4270 & select DMPC\_RB\_O.PO4 & select the PO4 atoms belonging to
4271 the first
4272 RigidBody in each DMPC molecule \\
4273 & select DMPC.20 & select the twentieth StuntDouble in each DMPC
4274 molecule \\
4275 \hline
4276 expression has two ``.''s & select DMPC.DMPC\_RB\_?.* &
4277 select all atoms
4278 belonging to all rigid bodies within all DMPC molecules \\
4279 \hline
4280 \end{tabular}
4281 \end{center}
4282
4283 \subsection{\label{section:index}Index expressions}
4284
4285 \begin{center}
4286 \begin{tabular}{|lp{4in}|}
4287 \hline
4288 {\bf examples} & {\bf translation of examples} \\
4289 \hline
4290 select 20 & select all of the StuntDoubles belonging to Molecule 20 \\
4291 select 20 to 30 & select all of the StuntDoubles belonging to
4292 molecules which have global indices between 20 (inclusive) and 30
4293 (exclusive) \\
4294 \hline
4295 \end{tabular}
4296 \end{center}
4297
4298 \subsection{\label{section:predefined}Predefined sets}
4299
4300 \begin{center}
4301 \begin{tabular}{|ll|}
4302 \hline
4303 {\bf keyword} & {\bf description} \\
4304 \hline
4305 all & select all StuntDoubles \\
4306 none & select none of the StuntDoubles \\
4307 \hline
4308 \end{tabular}
4309 \end{center}
4310
4311 \subsection{\label{section:userdefined}User-defined expressions}
4312
4313 Users can define arbitrary terms to represent groups of StuntDoubles,
4314 and then use the define terms in select commands. The general form for
4315 the define command is: {\bf define {\it term expression}}
4316
4317 Once defined, the user can specify such terms in boolean expressions
4318
4319 {\tt define SSDWATER SSD or SSD1 or SSDRF}
4320
4321 {\tt select SSDWATER}
4322
4323 \subsection{\label{section:comparison}Comparison expressions}
4324
4325 StuntDoubles can be selected by using comparision operators on their
4326 properties. The general form for the comparison command is: a property
4327 name, followed by a comparision operator and then a number.
4328
4329 \begin{center}
4330 \begin{tabular}{|l|l|}
4331 \hline
4332 {\bf property} & mass, charge \\
4333 {\bf comparison operator} & ``$>$'', ``$<$'', ``$=$'', ``$>=$'',
4334 ``$<=$'', ``$!=$'' \\
4335 \hline
4336 \end{tabular}
4337 \end{center}
4338
4339 For example, the phrase {\tt select mass > 16.0 and charge < -2}
4340 would select StuntDoubles which have mass greater than 16.0 and charges
4341 less than -2.
4342
4343 \subsection{\label{section:within}Within expressions}
4344
4345 The ``within'' keyword allows the user to select all StuntDoubles
4346 within the specified distance (in Angstroms) from a selection,
4347 including the selected atom itself. The general form for within
4348 selection is: {\tt select within(distance, expression)}
4349
4350 For example, the phrase {\tt select within(2.5, PO4 or NC4)} would
4351 select all StuntDoubles which are within 2.5 angstroms of PO4 or NC4
4352 atoms.
4353
4354 \section{\label{section:tools}Tools which use the selection command}
4355
4356 \subsection{\label{section:Dump2XYZ}Dump2XYZ}
4357
4358 Dump2XYZ can transform an {\sc OpenMD} dump file into a xyz file which can
4359 be opened by other molecular dynamics viewers such as Jmol and
4360 VMD. The options available for Dump2XYZ are as follows:
4361
4362
4363 \begin{longtable}[c]{|EFG|}
4364 \caption{Dump2XYZ Command-line Options}
4365 \\ \hline
4366 {\bf option} & {\bf verbose option} & {\bf behavior} \\ \hline
4367 \endhead
4368 \hline
4369 \endfoot
4370 -h & {\tt -{}-help} & Print help and exit \\
4371 -V & {\tt -{}-version} & Print version and exit \\
4372 -i & {\tt -{}-input=filename} & input dump file \\
4373 -o & {\tt -{}-output=filename} & output file name \\
4374 -n & {\tt -{}-frame=INT} & print every n frame (default=`1') \\
4375 -w & {\tt -{}-water} & skip the the waters (default=off) \\
4376 -m & {\tt -{}-periodicBox} & map to the periodic box (default=off)\\
4377 -z & {\tt -{}-zconstraint} & replace the atom types of zconstraint molecules (default=off) \\
4378 -r & {\tt -{}-rigidbody} & add a pseudo COM atom to rigidbody (default=off) \\
4379 -t & {\tt -{}-watertype} & replace the atom type of water model (default=on) \\
4380 -b & {\tt -{}-basetype} & using base atom type
4381 (default=off) \\
4382 -v& {\tt -{}-velocities} & Print velocities in xyz file (default=off)\\
4383 -f& {\tt -{}-forces} & Print forces xyz file (default=off)\\
4384 -u& {\tt -{}-vectors} & Print vectors (dipoles, etc) in xyz file
4385 (default=off)\\
4386 -c& {\tt -{}-charges} & Print charges in xyz file (default=off)\\
4387 -e& {\tt -{}-efield} & Print electric field vector in xyz file
4388 (default=off)\\
4389 & {\tt -{}-repeatX=INT} & The number of images to repeat in the x direction (default=`0') \\
4390 & {\tt -{}-repeatY=INT} & The number of images to repeat in the y direction (default=`0') \\
4391 & {\tt -{}-repeatZ=INT} & The number of images to repeat in the z direction (default=`0') \\
4392 -s & {\tt -{}-selection=selection script} & By specifying {\tt -{}-selection}=``selection command'' with Dump2XYZ, the user can select an arbitrary set of StuntDoubles to be
4393 converted. \\
4394 & {\tt -{}-originsele} & By specifying {\tt -{}-originsele}=``selection command'' with Dump2XYZ, the user can re-center the origin of the system around a specific StuntDouble \\
4395 & {\tt -{}-refsele} & In order to rotate the system, {\tt -{}-originsele} and {\tt -{}-refsele} must be given to define the new coordinate set. A StuntDouble which contains a dipole (the direction of the dipole is always (0, 0, 1) in body frame) is specified by {\tt -{}-originsele}. The new x-z plane is defined by the direction of the dipole and the StuntDouble is specified by {\tt -{}-refsele}.
4396 \end{longtable}
4397
4398
4399 \subsection{\label{section:StaticProps}StaticProps}
4400
4401 {\tt StaticProps} can compute properties which are averaged over some
4402 or all of the configurations that are contained within a dump file.
4403 The most common example of a static property that can be computed is
4404 the pair distribution function between atoms of type $A$ and other
4405 atoms of type $B$, $g_{AB}(r)$. StaticProps can also be used to
4406 compute the density distributions of other molecules in a reference
4407 frame {\it fixed to the body-fixed reference frame} of a selected atom
4408 or rigid body.
4409
4410 There are five seperate radial distribution functions availiable in
4411 {\sc OpenMD}. Since every radial distrbution function invlove the calculation
4412 between pairs of bodies, {\tt -{}-sele1} and {\tt -{}-sele2} must be specified to tell
4413 StaticProps which bodies to include in the calculation.
4414
4415 \begin{description}
4416 \item[{\tt -{}-gofr}] Computes the pair distribution function,
4417 \begin{equation*}
4418 g_{AB}(r) = \frac{1}{\rho_B}\frac{1}{N_A} \langle \sum_{i \in A}
4419 \sum_{j \in B} \delta(r - r_{ij}) \rangle
4420 \end{equation*}
4421 \item[{\tt -{}-r\_theta}] Computes the angle-dependent pair distribution
4422 function. The angle is defined by the intermolecular vector $\vec{r}$ and
4423 $z$-axis of DirectionalAtom A,
4424 \begin{equation*}
4425 g_{AB}(r, \cos \theta) = \frac{1}{\rho_B}\frac{1}{N_A} \langle \sum_{i \in A}
4426 \sum_{j \in B} \delta(r - r_{ij}) \delta(\cos \theta_{ij} - \cos \theta)\rangle
4427 \end{equation*}
4428 \item[{\tt -{}-r\_omega}] Computes the angle-dependent pair distribution
4429 function. The angle is defined by the $z$-axes of the two
4430 DirectionalAtoms A and B.
4431 \begin{equation*}
4432 g_{AB}(r, \cos \omega) = \frac{1}{\rho_B}\frac{1}{N_A} \langle \sum_{i \in A}
4433 \sum_{j \in B} \delta(r - r_{ij}) \delta(\cos \omega_{ij} - \cos \omega)\rangle
4434 \end{equation*}
4435 \item[{\tt -{}-theta\_omega}] Computes the pair distribution in the angular
4436 space $\theta, \omega$ defined by the two angles mentioned above.
4437 \begin{equation*}
4438 g_{AB}(\cos\theta, \cos \omega) = \frac{1}{\rho_B}\frac{1}{N_A} \langle \sum_{i \in A}
4439 \sum_{j \in B} \langle \delta(\cos \theta_{ij} - \cos \theta)
4440 \delta(\cos \omega_{ij} - \cos \omega)\rangle
4441 \end{equation*}
4442 \item[{\tt -{}-gxyz}] Calculates the density distribution of particles of type
4443 B in the body frame of particle A. Therefore, {\tt -{}-originsele} and
4444 {\tt -{}-refsele} must be given to define A's internal coordinate set as
4445 the reference frame for the calculation.
4446 \end{description}
4447
4448 The vectors (and angles) associated with these angular pair
4449 distribution functions are most easily seen in the figure below:
4450
4451 \begin{figure}
4452 \centering
4453 \includegraphics[width=3in]{definition.pdf}
4454 \caption[Definitions of the angles between directional objects]{ \\ Any
4455 two directional objects (DirectionalAtoms and RigidBodies) have a set
4456 of two angles ($\theta$, and $\omega$) between the z-axes of their
4457 body-fixed frames.}
4458 \label{fig:gofr}
4459 \end{figure}
4460
4461 The options available for {\tt StaticProps} are as follows:
4462 \begin{longtable}[c]{|EFG|}
4463 \caption{StaticProps Command-line Options}
4464 \\ \hline
4465 {\bf option} & {\bf verbose option} & {\bf behavior} \\ \hline
4466 \endhead
4467 \hline
4468 \endfoot
4469 -h& {\tt -{}-help} & Print help and exit \\
4470 -V& {\tt -{}-version} & Print version and exit \\
4471 -i& {\tt -{}-input=filename} & input dump file \\
4472 -o& {\tt -{}-output=filename} & output file name \\
4473 -n& {\tt -{}-step=INT} & process every n frame (default=`1') \\
4474 -r& {\tt -{}-nrbins=INT} & number of bins for distance (default=`100') \\
4475 -a& {\tt -{}-nanglebins=INT} & number of bins for cos(angle) (default= `50') \\
4476 -l& {\tt -{}-length=DOUBLE} & maximum length (Defaults to 1/2 smallest length of first frame) \\
4477 & {\tt -{}-sele1=selection script} & select the first StuntDouble set \\
4478 & {\tt -{}-sele2=selection script} & select the second StuntDouble set \\
4479 & {\tt -{}-sele3=selection script} & select the third StuntDouble set \\
4480 & {\tt -{}-refsele=selection script} & select reference (can only
4481 be used with {\tt -{}-gxyz}) \\
4482 & {\tt -{}-comsele=selection script}
4483 & select stunt doubles for center-of-mass
4484 reference point\\
4485 & {\tt -{}-seleoffset=INT} & global index offset for a second object (used
4486 to define a vector between sites in molecule)\\
4487
4488 & {\tt -{}-molname=STRING} & molecule name \\
4489 & {\tt -{}-begin=INT} & begin internal index \\
4490 & {\tt -{}-end=INT} & end internal index \\
4491 & {\tt -{}-radius=DOUBLE} & nanoparticle radius\\
4492 \hline
4493 \multicolumn{3}{|l|}{One option from the following group of options is required:} \\
4494 \hline
4495 & {\tt -{}-bo} & bond order parameter ({\tt -{}-rcut} must be specified) \\
4496 & {\tt -{}-bor} & bond order parameter as a function of
4497 radius ({\tt -{}-rcut} must be specified) \\
4498 & {\tt -{}-bad} & $N(\theta)$ bond angle density within ({\tt -{}-rcut} must be specified) \\
4499 & {\tt -{}-count} & count of molecules matching selection
4500 criteria (and associated statistics) \\
4501 -g& {\tt -{}-gofr} & $g(r)$ \\
4502 & {\tt -{}-gofz} & $g(z)$ \\
4503 & {\tt -{}-r\_theta} & $g(r, \cos(\theta))$ \\
4504 & {\tt -{}-r\_omega} & $g(r, \cos(\omega))$ \\
4505 & {\tt -{}-r\_z} & $g(r, z)$ \\
4506 & {\tt -{}-theta\_omega} & $g(\cos(\theta), \cos(\omega))$ \\
4507 & {\tt -{}-gxyz} & $g(x, y, z)$ \\
4508 & {\tt -{}-twodgofr} & 2D $g(r)$ (Slab width {\tt -{}-dz} must be specified)\\
4509 -p& {\tt -{}-p2} & $P_2$ order parameter ({\tt -{}-sele1} must be specified, {\tt -{}-sele2} is optional) \\
4510 & {\tt -{}-rp2} & Ripple order parameter ({\tt -{}-sele1} and {\tt -{}-sele2} must be specified) \\
4511 & {\tt -{}-scd} & $S_{CD}$ order parameter(either {\tt -{}-sele1}, {\tt -{}-sele2}, {\tt -{}-sele3} are specified or {\tt -{}-molname}, {\tt -{}-begin}, {\tt -{}-end} are specified) \\
4512 -d& {\tt -{}-density} & density plot \\
4513 & {\tt -{}-slab\_density} & slab density \\
4514 & {\tt -{}-p\_angle} & $p(\cos(\theta))$ ($\theta$
4515 is the angle between molecular axis and radial vector from origin\\
4516 & {\tt -{}-hxy} & Calculates the undulation spectrum, $h(x,y)$, of an interface \\
4517 & {\tt -{}-rho\_r} & $\rho(r)$\\
4518 & {\tt -{}-angle\_r} & $\theta(r)$ (spatially resolves the
4519 angle between the molecular axis and the radial vector from the
4520 origin)\\
4521 & {\tt -{}-hullvol} & hull volume of nanoparticle\\
4522 & {\tt -{}-rodlength} & length of nanorod\\
4523 -Q& {\tt -{}-tet\_param} & tetrahedrality order parameter ($Q$)\\
4524 & {\tt -{}-tet\_param\_z} & spatially-resolved tetrahedrality order
4525 parameter $Q(z)$\\
4526 & {\tt -{}-rnemdz} & slab-resolved RNEMD statistics (temperature,
4527 density, velocity)\\
4528 & {\tt -{}-rnemdr} & shell-resolved RNEMD statistics (temperature,
4529 density, angular velocity)
4530 \end{longtable}
4531
4532 \subsection{\label{section:DynamicProps}DynamicProps}
4533
4534 {\tt DynamicProps} computes time correlation functions from the
4535 configurations stored in a dump file. Typical examples of time
4536 correlation functions are the mean square displacement and the
4537 velocity autocorrelation functions. Once again, the selection syntax
4538 can be used to specify the StuntDoubles that will be used for the
4539 calculation. A general time correlation function can be thought of
4540 as:
4541 \begin{equation}
4542 C_{AB}(t) = \langle \vec{u}_A(t) \cdot \vec{v}_B(0) \rangle
4543 \end{equation}
4544 where $\vec{u}_A(t)$ is a vector property associated with an atom of
4545 type $A$ at time $t$, and $\vec{v}_B(t^{\prime})$ is a different vector
4546 property associated with an atom of type $B$ at a different time
4547 $t^{\prime}$. In most autocorrelation functions, the vector properties
4548 ($\vec{v}$ and $\vec{u}$) and the types of atoms ($A$ and $B$) are
4549 identical, and the three calculations built in to {\tt DynamicProps}
4550 make these assumptions. It is possible, however, to make simple
4551 modifications to the {\tt DynamicProps} code to allow the use of {\it
4552 cross} time correlation functions (i.e. with different vectors). The
4553 ability to use two selection scripts to select different types of
4554 atoms is already present in the code.
4555
4556 The options available for DynamicProps are as follows:
4557 \begin{longtable}[c]{|EFG|}
4558 \caption{DynamicProps Command-line Options}
4559 \\ \hline
4560 {\bf option} & {\bf verbose option} & {\bf behavior} \\ \hline
4561 \endhead
4562 \hline
4563 \endfoot
4564 -h& {\tt -{}-help} & Print help and exit \\
4565 -V& {\tt -{}-version} & Print version and exit \\
4566 -i& {\tt -{}-input=filename} & input dump file \\
4567 -o& {\tt -{}-output=filename} & output file name \\
4568 & {\tt -{}-sele1=selection script} & select first StuntDouble set \\
4569 & {\tt -{}-sele2=selection script} & select second StuntDouble set (if sele2 is not set, use script from sele1) \\
4570 & {\tt -{}-order=INT} & Lengendre Polynomial Order\\
4571 -z& {\tt -{}-nzbins=INT} & Number of $z$ bins (default=`100`)\\
4572 -m& {\tt -{}-memory=memory specification}
4573 &Available memory
4574 (default=`2G`)\\
4575 \hline
4576 \multicolumn{3}{|l|}{One option from the following group of options is required:} \\
4577 \hline
4578 -s& {\tt -{}-selecorr} & selection correlation function \\
4579 -r& {\tt -{}-rcorr} & compute mean squared displacement \\
4580 -v& {\tt -{}-vcorr} & velocity autocorrelation function \\
4581 -d& {\tt -{}-dcorr} & dipole correlation function \\
4582 -l& {\tt -{}-lcorr} & Lengendre correlation function \\
4583 & {\tt -{}-lcorrZ} & Lengendre correlation function binned by $z$ \\
4584 & {\tt -{}-cohZ} & Lengendre correlation function for OH bond vectors binned by $z$\\
4585 -M& {\tt -{}-sdcorr} & System dipole correlation function\\
4586 & {\tt -{}-r\_rcorr} & Radial mean squared displacement\\
4587 & {\tt -{}-thetacorr} & Angular mean squared displacement\\
4588 & {\tt -{}-drcorr} & Directional mean squared displacement for particles with unit vectors\\
4589 & {\tt -{}-helfandEcorr} & Helfand moment for thermal conductvity\\
4590 -p& {\tt -{}-momentum} & Helfand momentum for viscosity\\
4591 & {\tt -{}-stresscorr} & Stress tensor correlation function
4592 \end{longtable}
4593
4594 \chapter{\label{section:PreparingInput} Preparing Input Configurations}
4595
4596 {\sc OpenMD} version 4 comes with a few utility programs to aid in
4597 setting up initial configuration and meta-data files. Usually, a user
4598 is interested in either importing a structure from some other format
4599 (usually XYZ or PDB), or in building an initial configuration in some
4600 perfect crystalline lattice. The programs bundled with {\sc OpenMD}
4601 which import coordinate files are {\tt atom2md}, {\tt xyz2md}, and
4602 {\tt pdb2md}. The programs which generate perfect crystals are called
4603 {\tt SimpleBuilder} and {\tt RandomBuilder}
4604
4605 \section{\label{section:atom2md}atom2md, xyz2md, and pdb2md}
4606
4607 {\tt atom2md}, {\tt xyz2md}, and {\tt pdb2md} attempt to construct
4608 {\tt .md} files from a single file containing only atomic coordinate
4609 information. To do this task, they make reasonable guesses about
4610 bonding from the distance between atoms in the coordinate, and attempt
4611 to identify other terms in the potential energy from the topology of
4612 the graph of discovered bonds. This procedure is not perfect, and the
4613 user should check the discovered bonding topology that is contained in
4614 the {\tt $<$MetaData$>$} block in the file that is generated.
4615
4616 Typically, the user would run:
4617
4618 {\tt atom2md $<$input spec$>$ [Options]}
4619
4620 Here {\tt $<$input spec$>$} can be used to specify the type of file being
4621 used for configuration input. I.e. using {\tt -ipdb} specifies that the
4622 input file contains coordinate information in the PDB format.
4623
4624 The options available for atom2md are as follows:
4625 \begin{longtable}[c]{|HI|}
4626 \caption{atom2md Command-line Options}
4627 \\ \hline
4628 {\bf option} & {\bf behavior} \\ \hline
4629 \endhead
4630 \hline
4631 \endfoot
4632 -f \# & Start import at molecule \# specified \\
4633 -l \# & End import at molecule \# specified \\
4634 -t & All input files describe a single molecule \\
4635 -e & Continue with next object after error, if possible \\
4636 -z & Compress the output with gzip \\
4637 -H & Outputs this help text \\
4638 -Hxxx & (xxx is file format ID e.g. -Hpdb) gives format info \\
4639 -Hall & Outputs details of all formats \\
4640 -V & Outputs version number \\
4641 \hline
4642 \multicolumn{2}{|l|}{The following file formats are recognized:}\\
4643 \hline
4644 ent & Protein Data Bank format \\
4645 in & {\sc OpenMD} cartesian coordinates format \\
4646 pdb & Protein Data Bank format \\
4647 prep & Amber Prep format \\
4648 xyz & XYZ cartesian coordinates format \\
4649 \hline
4650 \multicolumn{2}{|l|}{More specific info and options are available
4651 using -H$<$format-type$>$, e.g. -Hpdb}
4652 \end{longtable}
4653
4654 The specific programs {\tt xyz2md} and {\tt pdb2md} are identical
4655 to {\tt atom2md}, but they use a specific input format and do not
4656 expect the the input specifier on the command line.
4657
4658
4659 \section{\label{section:SimpleBuilder}SimpleBuilder}
4660
4661 {\tt SimpleBuilder} creates simple lattice structures. It requires an
4662 initial, but skeletal {\sc OpenMD} file to specify the components that are to
4663 be placed on the lattice. The total number of placed molecules will
4664 be shown at the top of the configuration file that is generated, and
4665 that number may not match the original meta-data file, so a new
4666 meta-data file is also generated which matches the lattice structure.
4667
4668 The options available for SimpleBuilder are as follows:
4669 \begin{longtable}[c]{|EFG|}
4670 \caption{SimpleBuilder Command-line Options}
4671 \\ \hline
4672 {\bf option} & {\bf verbose option} & {\bf behavior} \\ \hline
4673 \endhead
4674 \hline
4675 \endfoot
4676 -h& {\tt -{}-help} & Print help and exit\\
4677 -V& {\tt -{}-version} & Print version and exit\\
4678 -o& {\tt -{}-output=STRING} & Output file name\\
4679 & {\tt -{}-density=DOUBLE} & density ($\mathrm{g~cm}^{-3}$)\\
4680 & {\tt -{}-nx=INT} & number of unit cells in x\\
4681 & {\tt -{}-ny=INT} & number of unit cells in y\\
4682 & {\tt -{}-nz=INT} & number of unit cells in z
4683 \end{longtable}
4684
4685 \section{\label{section:icosahedralBuilder}icosahedralBuilder}
4686
4687 {\tt icosahedralBuilder} creates single-component geometric solids
4688 that can be useful in simulating nanostructures. Like the other
4689 builders, it requires an initial, but skeletal {\sc OpenMD} file to
4690 specify the component that is to be placed on the lattice. The total
4691 number of placed molecules will be shown at the top of the
4692 configuration file that is generated, and that number may not match
4693 the original meta-data file, so a new meta-data file is also generated
4694 which matches the lattice structure.
4695
4696 The options available for icosahedralBuilder are as follows:
4697 \begin{longtable}[c]{|EFG|}
4698 \caption{icosahedralBuilder Command-line Options}
4699 \\ \hline
4700 {\bf option} & {\bf verbose option} & {\bf behavior} \\ \hline
4701 \endhead
4702 \hline
4703 \endfoot
4704 -h& {\tt -{}-help} & Print help and exit\\
4705 -V& {\tt -{}-version} & Print version and exit\\
4706 -o& {\tt -{}-output=STRING} & Output file name\\
4707 -n& {\tt -{}-shells=INT} & Nanoparticle shells\\
4708 -d& {\tt -{}-latticeConstant=DOUBLE} & Lattice spacing in Angstroms for cubic lattice.\\
4709 -c& {\tt -{}-columnAtoms=INT} & Number of atoms along central
4710 column (Decahedron only)\\
4711 -t& {\tt -{}-twinAtoms=INT} & Number of atoms along twin
4712 boundary (Decahedron only) \\
4713 -p& {\tt -{}-truncatedPlanes=INT} & Number of truncated planes (Curling-stone Decahedron only)\\
4714 \hline
4715 \multicolumn{3}{|l|}{One option from the following group of options is required:} \\
4716 \hline
4717 & {\tt -{}-ico} & Create an Icosahedral cluster \\
4718 & {\tt -{}-deca} & Create a regualar Decahedral cluster\\
4719 & {\tt -{}-ino} & Create an Ino Decahedral cluster\\
4720 & {\tt -{}-marks} & Create a Marks Decahedral cluster\\
4721 & {\tt -{}-stone} & Create a Curling-stone Decahedral cluster
4722 \end{longtable}
4723
4724
4725 \section{\label{section:Hydro}Hydro}
4726 {\tt Hydro} generates resistance tensor ({\tt .diff}) files which are
4727 required when using the Langevin integrator using complex rigid
4728 bodies. {\tt Hydro} supports two approximate models: the {\tt
4729 BeadModel} and {\tt RoughShell}. Additionally, {\tt Hydro} can
4730 generate resistance tensor files using analytic solutions for simple
4731 shapes. To generate a {\tt }.diff file, a meta-data file is needed as
4732 the input file. Since the resistance tensor depends on these
4733 quantities, the {\tt viscosity} of the solvent and the temperature
4734 ({\tt targetTemp}) of the system must be defined in meta-data file. If
4735 the approximate model in use is the {\tt RoughShell} model the {\tt
4736 beadSize} (the diameter of the small beads used to approximate the
4737 surface of the body) must also be specified.
4738
4739 The options available for Hydro are as follows:
4740 \begin{longtable}[c]{|EFG|}
4741 \caption{Hydro Command-line Options}
4742 \\ \hline
4743 {\bf option} & {\bf verbose option} & {\bf behavior} \\ \hline
4744 \endhead
4745 \hline
4746 \endfoot
4747 -h& {\tt -{}-help} & Print help and exit\\
4748 -V& {\tt -{}-version} & Print version and exit\\
4749 -i& {\tt -{}-input=filename} & input MetaData (md) file\\
4750 -o& {\tt -{}-output=STRING} & Output file name\\
4751 & {\tt -{}-model=STRING} & hydrodynamics model (supports both
4752 {\tt RoughShell} and {\tt BeadModel})\\
4753 -b& {\tt -{}-beads} & generate the beads only,
4754 hydrodynamic calculations will not be performed (default=off)\\
4755 \end{longtable}
4756
4757
4758
4759
4760
4761 \chapter{\label{section:parallelization} Parallel Simulation Implementation}
4762
4763 Although processor power is continually improving, it is still
4764 unreasonable to simulate systems of more than 10,000 atoms on a single
4765 processor. To facilitate study of larger system sizes or smaller
4766 systems for longer time scales, parallel methods were developed to
4767 allow multiple CPU's to share the simulation workload. Three general
4768 categories of parallel decomposition methods have been developed:
4769 these are the atomic,\cite{Fox88} spatial~\cite{plimpton95} and
4770 force~\cite{Paradyn} decomposition methods.
4771
4772 Algorithmically simplest of the three methods is atomic decomposition,
4773 where $N$ particles in a simulation are split among $P$ processors for
4774 the duration of the simulation. Computational cost scales as an
4775 optimal $\mathcal{O}(N/P)$ for atomic decomposition. Unfortunately, all
4776 processors must communicate positions and forces with all other
4777 processors at every force evaluation, leading the communication costs
4778 to scale as an unfavorable $\mathcal{O}(N)$, \emph{independent of the
4779 number of processors}. This communication bottleneck led to the
4780 development of spatial and force decomposition methods, in which
4781 communication among processors scales much more favorably. Spatial or
4782 domain decomposition divides the physical spatial domain into 3D boxes
4783 in which each processor is responsible for calculation of forces and
4784 positions of particles located in its box. Particles are reassigned to
4785 different processors as they move through simulation space. To
4786 calculate forces on a given particle, a processor must simply know the
4787 positions of particles within some cutoff radius located on nearby
4788 processors rather than the positions of particles on all
4789 processors. Both communication between processors and computation
4790 scale as $\mathcal{O}(N/P)$ in the spatial method. However, spatial
4791 decomposition adds algorithmic complexity to the simulation code and
4792 is not very efficient for small $N$, since the overall communication
4793 scales as the surface to volume ratio $\mathcal{O}(N/P)^{2/3}$ in
4794 three dimensions.
4795
4796 The parallelization method used in {\sc OpenMD} is the force
4797 decomposition method.\cite{hendrickson:95} Force decomposition assigns
4798 particles to processors based on a block decomposition of the force
4799 matrix. Processors are split into an optimally square grid forming row
4800 and column processor groups. Forces are calculated on particles in a
4801 given row by particles located in that processor's column
4802 assignment. One deviation from the algorithm described by Hendrickson
4803 {\it et al.} is the use of column ordering based on the row indexes
4804 preventing the need for a transpose operation necessitating a second
4805 communication step when gathering the final force components. Force
4806 decomposition is less complex to implement than the spatial method but
4807 still scales computationally as $\mathcal{O}(N/P)$ and scales as
4808 $\mathcal{O}(N/\sqrt{P})$ in communication cost. Plimpton has also
4809 found that force decompositions scale more favorably than spatial
4810 decompositions for systems up to 10,000 atoms and favorably compete
4811 with spatial methods up to 100,000 atoms.\cite{plimpton95}
4812
4813 \chapter{\label{section:conclusion}Conclusion}
4814
4815 We have presented a new parallel simulation program called {\sc
4816 OpenMD}. This program offers some novel capabilities, but mostly makes
4817 available a library of modern object-oriented code for the scientific
4818 community to use freely. Notably, {\sc OpenMD} can handle symplectic
4819 integration of objects (atoms and rigid bodies) which have
4820 orientational degrees of freedom. It can also work with transition
4821 metal force fields and point-dipoles. It is capable of scaling across
4822 multiple processors through the use of force based decomposition. It
4823 also implements several advanced integrators allowing the end user
4824 control over temperature and pressure. In addition, it is capable of
4825 integrating constrained dynamics through both the {\sc rattle}
4826 algorithm and the $z$-constraint method.
4827
4828 We encourage other researchers to download and apply this program to
4829 their own research problems. By making the code available, we hope to
4830 encourage other researchers to contribute their own code and make it a
4831 more powerful package for everyone in the molecular dynamics community
4832 to use. All source code for {\sc OpenMD} is available for download at
4833 {\tt http://openmd.net}.
4834
4835 \chapter{Acknowledgments}
4836
4837 Development of {\sc OpenMD} was funded by a New Faculty Award from the
4838 Camille and Henry Dreyfus Foundation and by the National Science
4839 Foundation under grant CHE-0134881. Computation time was provided by
4840 the Notre Dame Bunch-of-Boxes (B.o.B) computer cluster under NSF grant
4841 DMR-0079647.
4842
4843
4844 \bibliographystyle{aip}
4845 \bibliography{openmdDoc}
4846
4847 \end{document}