ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/oopsePaper/EmpericalEnergy.tex
Revision: 973
Committed: Wed Jan 21 19:00:23 2004 UTC (20 years, 5 months ago) by mmeineke
Content type: application/x-tex
File size: 24382 byte(s)
Log Message:
minor corrections to the previous commit

File Contents

# Content
1
2 \section{\label{sec:empiricalEnergy}The Empirical Energy Functions}
3
4 \subsection{\label{sec:atomsMolecules}Atoms, Molecules and Rigid Bodies}
5
6 The basic unit of an {\sc oopse} simulation is the atom. The
7 parameters describing the atom are generalized to make the atom as
8 flexible a representation as possible. They may represent specific
9 atoms of an element, or be used for collections of atoms such as
10 methyl and carbonyl groups. The atoms are also capable of having
11 directional components associated with them (\emph{e.g.}~permanent
12 dipoles). Charges on atoms are not currently supported by {\sc oopse}.
13
14 \begin{lstlisting}[caption={[Specifier for molecules and atoms] A sample specification of the simple Ar molecule},label=sch:AtmMole]
15 molecule{
16 name = "Ar";
17 nAtoms = 1;
18 atom[0]{
19 type="Ar";
20 position( 0.0, 0.0, 0.0 );
21 }
22 }
23 \end{lstlisting}
24
25 Atoms can be collected into secondary srtructures such as rigid bodies
26 or molecules. The molecule is a way for {\sc oopse} to keep track of
27 the atoms in a simulation in logical manner. Molecular units store the
28 identities of all the atoms associated with themselves, and are
29 responsible for the evaluation of their own internal interactions
30 (\emph{i.e.}~bonds, bends, and torsions). Scheme \ref{sch:AtmMole}
31 shws how one creates a molecule in the \texttt{.mdl} files. The
32 position of the atoms given in the declaration are relative to the
33 origin of the molecule, and is used when creating a system containing
34 the molecule.
35
36 As stated previously, one of the features that sets {\sc oopse} apart
37 from most of the current molecular simulation packages is the ability
38 to handle rigid body dynamics. Rigid bodies are non-spherical
39 particles or collections of particles that have a constant internal
40 potential and move collectively.\cite{Goldstein01} They are not
41 included in most simulation packages because of the requirement to
42 propagate the orientational degrees of freedom. Until recently,
43 integrators which propagate orientational motion have been lacking.
44
45 Moving a rigid body involves determination of both the force and
46 torque applied by the surroundings, which directly affect the
47 translational and rotational motion in turn. In order to accumulate
48 the total force on a rigid body, the external forces and torques must
49 first be calculated for all the internal particles. The total force on
50 the rigid body is simply the sum of these external forces.
51 Accumulation of the total torque on the rigid body is more complex
52 than the force in that it is the torque applied on the center of mass
53 that dictates rotational motion. The torque on rigid body {\it i} is
54 \begin{equation}
55 \boldsymbol{\tau}_i=
56 \sum_{a}(\mathbf{r}_{ia}-\mathbf{r}_i)\times \mathbf{f}_{ia}
57 + \boldsymbol{\tau}_{ia},
58 \label{eq:torqueAccumulate}
59 \end{equation}
60 where $\boldsymbol{\tau}_i$ and $\mathbf{r}_i$ are the torque on and
61 position of the center of mass respectively, while $\mathbf{f}_{ia}$,
62 $\mathbf{r}_{ia}$, and $\boldsymbol{\tau}_{ia}$ are the force on,
63 position of, and torque on the component particles of the rigid body.
64
65 The summation of the total torque is done in the body fixed axis of
66 the rigid body. In order to move between the space fixed and body
67 fixed coordinate axes, parameters describing the orientation must be
68 maintained for each rigid body. At a minimum, the rotation matrix
69 (\textbf{A}) can be described by the three Euler angles ($\phi,
70 \theta,$ and $\psi$), where the elements of \textbf{A} are composed of
71 trigonometric operations involving $\phi, \theta,$ and
72 $\psi$.\cite{Goldstein01} In order to avoid numerical instabilities
73 inherent in using the Euler angles, the four parameter ``quaternion''
74 scheme is often used. The elements of \textbf{A} can be expressed as
75 arithmetic operations involving the four quaternions ($q_0, q_1, q_2,$
76 and $q_3$).\cite{allen87:csl} Use of quaternions also leads to
77 performance enhancements, particularly for very small
78 systems.\cite{Evans77}
79
80 {\sc oopse} utilizes a relatively new scheme that propagates the
81 entire nine parameter rotation matrix internally. Further discussion
82 on this choice can be found in Sec.~\ref{sec:integrate}. An example
83 definition of a riged body can be seen in Scheme
84 \ref{sch:rigidBody}. The positions in the atom definitions are the
85 placements of the atoms relative to the origin of the rigid body,
86 which itself has a position relative to the origin of the molecule.
87
88 \begin{lstlisting}[caption={[Defining rigid bodies]A sample definition of a rigid body},label={sch:rigidBody}]
89 molecule{
90 name = "TIP3P_water";
91 nRigidBodies = 1;
92 rigidBody[0]{
93 nAtoms = 3;
94 atom[0]{
95 type = "O_TIP3P";
96 position( 0.0, 0.0, -0.06556 );
97 }
98 atom[1]{
99 type = "H_TIP3P";
100 position( 0.0, 0.75695, 0.52032 );
101 }
102 atom[2]{
103 type = "H_TIP3P";
104 position( 0.0, -0.75695, 0.52032 );
105 }
106 position( 0.0, 0.0, 0.0 );
107 orientation( 0.0, 0.0, 1.0 );
108 }
109 }
110 \end{lstlisting}
111
112 \subsection{\label{sec:LJPot}The Lennard Jones Potential}
113
114 The most basic force field implemented in {\sc oopse} is the
115 Lennard-Jones potential, which mimics the van der Waals interaction at
116 long distances, and uses an empirical repulsion at short
117 distances. The Lennard-Jones potential is given by:
118 \begin{equation}
119 V_{\text{LJ}}(r_{ij}) =
120 4\epsilon_{ij} \biggl[
121 \biggl(\frac{\sigma_{ij}}{r_{ij}}\biggr)^{12}
122 - \biggl(\frac{\sigma_{ij}}{r_{ij}}\biggr)^{6}
123 \biggr]
124 \label{eq:lennardJonesPot}
125 \end{equation}
126 Where $r_{ij}$ is the distance between particles $i$ and $j$,
127 $\sigma_{ij}$ scales the length of the interaction, and
128 $\epsilon_{ij}$ scales the well depth of the potential. Scheme
129 \ref{sch:LJFF} gives and example partial \texttt{.bass} file that
130 shows a system of 108 Ar particles simulated with the Lennard-Jones
131 force field.
132
133 \begin{lstlisting}[caption={[Invocation of the Lennard-Jones force field] A sample system using the Lennard-Jones force field.},label={sch:LJFF}]
134
135 /*
136 * The Ar molecule is specified
137 * external to the.bass file
138 */
139
140 #include "argon.mdl"
141
142 nComponents = 1;
143 component{
144 type = "Ar";
145 nMol = 108;
146 }
147
148 /*
149 * The initial configuration is generated
150 * before the simulation is invoked.
151 */
152
153 initialConfig = "./argon.init";
154
155 forceField = "LJ";
156 \end{lstlisting}
157
158 Because this potential is calculated between all pairs, the force
159 evaluation can become computationally expensive for large systems. To
160 keep the pair evaluations to a manageable number, {\sc oopse} employs
161 a cut-off radius.\cite{allen87:csl} The cutoff radius is set to be
162 $2.5\sigma_{ii}$, where $\sigma_{ii}$ is the largest Lennard-Jones
163 length parameter present in the simulation. Truncating the calculation
164 at $r_{\text{cut}}$ introduces a discontinuity into the potential
165 energy. To offset this discontinuity, the energy value at
166 $r_{\text{cut}}$ is subtracted from the potential. This causes the
167 potential to go to zero smoothly at the cut-off radius.
168
169 Interactions between dissimilar particles requires the generation of
170 cross term parameters for $\sigma$ and $\epsilon$. These are
171 calculated through the Lorentz-Berthelot mixing
172 rules:\cite{allen87:csl}
173 \begin{equation}
174 \sigma_{ij} = \frac{1}{2}[\sigma_{ii} + \sigma_{jj}]
175 \label{eq:sigmaMix}
176 \end{equation}
177 and
178 \begin{equation}
179 \epsilon_{ij} = \sqrt{\epsilon_{ii} \epsilon_{jj}}
180 \label{eq:epsilonMix}
181 \end{equation}
182
183
184
185 \subsection{\label{sec:DUFF}Dipolar Unified-Atom Force Field}
186
187 The dipolar unified-atom force field ({\sc duff}) was developed to
188 simulate lipid bilayers. The simulations require a model capable of
189 forming bilayers, while still being sufficiently computationally
190 efficient to allow large systems ($\approx$100's of phospholipids,
191 $\approx$1000's of waters) to be simulated for long times
192 ($\approx$10's of nanoseconds).
193
194 With this goal in mind, {\sc duff} has no point
195 charges. Charge-neutral distributions were replaced with dipoles,
196 while most atoms and groups of atoms were reduced to Lennard-Jones
197 interaction sites. This simplification cuts the length scale of long
198 range interactions from $\frac{1}{r}$ to $\frac{1}{r^3}$, allowing us
199 to avoid the computationally expensive Ewald sum. Instead, we can use
200 neighbor-lists, reaction field, and cutoff radii for the dipolar
201 interactions.
202
203 As an example, lipid head-groups in {\sc duff} are represented as
204 point dipole interaction sites. By placing a dipole of 20.6~Debye at
205 the head group center of mass, our model mimics the head group of
206 phosphatidylcholine.\cite{Cevc87} Additionally, a large Lennard-Jones
207 site is located at the pseudoatom's center of mass. The model is
208 illustrated by the dark grey atom in Fig.~\ref{fig:lipidModel}. The
209 water model we use to complement the dipoles of the lipids is our
210 reparameterization of the soft sticky dipole (SSD) model of Ichiye
211 \emph{et al.}\cite{liu96:new_model}
212
213 \begin{figure}
214 \epsfxsize=\linewidth
215 \epsfbox{lipidModel.eps}
216 \caption{A representation of the lipid model. $\phi$ is the torsion angle, $\theta$ %
217 is the bend angle, $\mu$ is the dipole moment of the head group, and n
218 is the chain length.}
219 \label{fig:lipidModel}
220 \end{figure}
221
222 We have used a set of scalable parameters to model the alkyl groups
223 with Lennard-Jones sites. For this, we have borrowed parameters from
224 the TraPPE force field of Siepmann
225 \emph{et al}.\cite{Siepmann1998} TraPPE is a unified-atom
226 representation of n-alkanes, which is parametrized against phase
227 equilibria using Gibbs ensemble Monte Carlo simulation
228 techniques.\cite{Siepmann1998} One of the advantages of TraPPE is that
229 it generalizes the types of atoms in an alkyl chain to keep the number
230 of pseudoatoms to a minimum; the parameters for an atom such as
231 $\text{CH}_2$ do not change depending on what species are bonded to
232 it.
233
234 TraPPE also constrains all bonds to be of fixed length. Typically,
235 bond vibrations are the fastest motions in a molecular dynamic
236 simulation. Small time steps between force evaluations must be used to
237 ensure adequate sampling of the bond potential to ensure conservation
238 of energy. By constraining the bond lengths, larger time steps may be
239 used when integrating the equations of motion. A simulation using {\sc
240 duff} is illustrated in Scheme \ref{sch:DUFF}.
241
242 \begin{lstlisting}[caption={[Invocation of {\sc duff}]Sample \texttt{.bass} file showing a simulation utilizing {\sc duff}},label={sch:DUFF}]
243
244 #include "water.mdl"
245 #include "lipid.mdl"
246
247 nComponents = 2;
248 component{
249 type = "simpleLipid_16";
250 nMol = 60;
251 }
252
253 component{
254 type = "SSD_water";
255 nMol = 1936;
256 }
257
258 initialConfig = "bilayer.init";
259
260 forceField = "DUFF";
261
262 \end{lstlisting}
263
264 \subsubsection{\label{subSec:energyFunctions}{\sc duff} Energy Functions}
265
266 The total potential energy function in {\sc duff} is
267 \begin{equation}
268 V = \sum^{N}_{I=1} V^{I}_{\text{Internal}}
269 + \sum^{N}_{I=1} \sum_{J>I} V^{IJ}_{\text{Cross}}
270 \label{eq:totalPotential}
271 \end{equation}
272 Where $V^{I}_{\text{Internal}}$ is the internal potential of molecule $I$:
273 \begin{equation}
274 V^{I}_{\text{Internal}} =
275 \sum_{\theta_{ijk} \in I} V_{\text{bend}}(\theta_{ijk})
276 + \sum_{\phi_{ijkl} \in I} V_{\text{torsion}}(\phi_{ijkl})
277 + \sum_{i \in I} \sum_{(j>i+4) \in I}
278 \biggl[ V_{\text{LJ}}(r_{ij}) + V_{\text{dipole}}
279 (\mathbf{r}_{ij},\boldsymbol{\Omega}_{i},\boldsymbol{\Omega}_{j})
280 \biggr]
281 \label{eq:internalPotential}
282 \end{equation}
283 Here $V_{\text{bend}}$ is the bend potential for all 1, 3 bonded pairs
284 within the molecule $I$, and $V_{\text{torsion}}$ is the torsion potential
285 for all 1, 4 bonded pairs. The pairwise portions of the internal
286 potential are excluded for pairs that are closer than three bonds,
287 i.e.~atom pairs farther away than a torsion are included in the
288 pair-wise loop.
289
290
291 The bend potential of a molecule is represented by the following function:
292 \begin{equation}
293 V_{\text{bend}}(\theta_{ijk}) = k_{\theta}( \theta_{ijk} - \theta_0 )^2 \label{eq:bendPot}
294 \end{equation}
295 Where $\theta_{ijk}$ is the angle defined by atoms $i$, $j$, and $k$
296 (see Fig.~\ref{fig:lipidModel}), $\theta_0$ is the equilibrium
297 bond angle, and $k_{\theta}$ is the force constant which determines the
298 strength of the harmonic bend. The parameters for $k_{\theta}$ and
299 $\theta_0$ are borrowed from those in TraPPE.\cite{Siepmann1998}
300
301 The torsion potential and parameters are also borrowed from TraPPE. It is
302 of the form:
303 \begin{equation}
304 V_{\text{torsion}}(\phi) = c_1[1 + \cos \phi]
305 + c_2[1 + \cos(2\phi)]
306 + c_3[1 + \cos(3\phi)]
307 \label{eq:origTorsionPot}
308 \end{equation}
309 Here $\phi$ is the angle defined by four bonded neighbors $i$,
310 $j$, $k$, and $l$ (again, see Fig.~\ref{fig:lipidModel}). For
311 computational efficiency, the torsion potential has been recast after
312 the method of CHARMM,\cite{charmm1983} in which the angle series is
313 converted to a power series of the form:
314 \begin{equation}
315 V_{\text{torsion}}(\phi) =
316 k_3 \cos^3 \phi + k_2 \cos^2 \phi + k_1 \cos \phi + k_0
317 \label{eq:torsionPot}
318 \end{equation}
319 Where:
320 \begin{align*}
321 k_0 &= c_1 + c_3 \\
322 k_1 &= c_1 - 3c_3 \\
323 k_2 &= 2 c_2 \\
324 k_3 &= 4c_3
325 \end{align*}
326 By recasting the potential as a power series, repeated trigonometric
327 evaluations are avoided during the calculation of the potential energy.
328
329
330 The cross potential between molecules $I$ and $J$, $V^{IJ}_{\text{Cross}}$, is
331 as follows:
332 \begin{equation}
333 V^{IJ}_{\text{Cross}} =
334 \sum_{i \in I} \sum_{j \in J}
335 \biggl[ V_{\text{LJ}}(r_{ij}) + V_{\text{dipole}}
336 (\mathbf{r}_{ij},\boldsymbol{\Omega}_{i},\boldsymbol{\Omega}_{j})
337 + V_{\text{sticky}}
338 (\mathbf{r}_{ij},\boldsymbol{\Omega}_{i},\boldsymbol{\Omega}_{j})
339 \biggr]
340 \label{eq:crossPotentail}
341 \end{equation}
342 Where $V_{\text{LJ}}$ is the Lennard Jones potential,
343 $V_{\text{dipole}}$ is the dipole dipole potential, and
344 $V_{\text{sticky}}$ is the sticky potential defined by the SSD model
345 (Sec.~\ref{sec:SSD}). Note that not all atom types include all
346 interactions.
347
348 The dipole-dipole potential has the following form:
349 \begin{equation}
350 V_{\text{dipole}}(\mathbf{r}_{ij},\boldsymbol{\Omega}_{i},
351 \boldsymbol{\Omega}_{j}) = \frac{|\mu_i||\mu_j|}{4\pi\epsilon_{0}r_{ij}^{3}} \biggl[
352 \boldsymbol{\hat{u}}_{i} \cdot \boldsymbol{\hat{u}}_{j}
353 -
354 \frac{3(\boldsymbol{\hat{u}}_i \cdot \mathbf{r}_{ij}) %
355 (\boldsymbol{\hat{u}}_j \cdot \mathbf{r}_{ij}) }
356 {r^{2}_{ij}} \biggr]
357 \label{eq:dipolePot}
358 \end{equation}
359 Here $\mathbf{r}_{ij}$ is the vector starting at atom $i$ pointing
360 towards $j$, and $\boldsymbol{\Omega}_i$ and $\boldsymbol{\Omega}_j$
361 are the orientational degrees of freedom for atoms $i$ and $j$
362 respectively. $|\mu_i|$ is the magnitude of the dipole moment of atom
363 $i$, $\boldsymbol{\hat{u}}_i$ is the standard unit orientation
364 vector of $\boldsymbol{\Omega}_i$, and $\boldsymbol{\hat{r}}_{ij}$ is
365 the unit vector pointing along $\mathbf{r}_{ij}$.
366
367
368 \subsubsection{\label{sec:SSD}The {\sc duff} Water Models: SSD/E and SSD/RF}
369
370 In the interest of computational efficiency, the default solvent used
371 by {\sc oopse} is the extended Soft Sticky Dipole (SSD/E) water
372 model.\cite{Gezelter04} The original SSD was developed by Ichiye
373 \emph{et al.}\cite{liu96:new_model} as a modified form of the hard-sphere
374 water model proposed by Bratko, Blum, and
375 Luzar.\cite{Bratko85,Bratko95} It consists of a single point dipole
376 with a Lennard-Jones core and a sticky potential that directs the
377 particles to assume the proper hydrogen bond orientation in the first
378 solvation shell. Thus, the interaction between two SSD water molecules
379 \emph{i} and \emph{j} is given by the potential
380 \begin{equation}
381 V_{ij} =
382 V_{ij}^{LJ} (r_{ij})\ + V_{ij}^{dp}
383 (\mathbf{r}_{ij},\boldsymbol{\Omega}_i,\boldsymbol{\Omega}_j)\ +
384 V_{ij}^{sp}
385 (\mathbf{r}_{ij},\boldsymbol{\Omega}_i,\boldsymbol{\Omega}_j),
386 \label{eq:ssdPot}
387 \end{equation}
388 where the $\mathbf{r}_{ij}$ is the position vector between molecules
389 \emph{i} and \emph{j} with magnitude equal to the distance $r_{ij}$, and
390 $\boldsymbol{\Omega}_i$ and $\boldsymbol{\Omega}_j$ represent the
391 orientations of the respective molecules. The Lennard-Jones and dipole
392 parts of the potential are given by equations \ref{eq:lennardJonesPot}
393 and \ref{eq:dipolePot} respectively. The sticky part is described by
394 the following,
395 \begin{equation}
396 u_{ij}^{sp}(\mathbf{r}_{ij},\boldsymbol{\Omega}_i,\boldsymbol{\Omega}_j)=
397 \frac{\nu_0}{2}[s(r_{ij})w(\mathbf{r}_{ij},
398 \boldsymbol{\Omega}_i,\boldsymbol{\Omega}_j) +
399 s^\prime(r_{ij})w^\prime(\mathbf{r}_{ij},
400 \boldsymbol{\Omega}_i,\boldsymbol{\Omega}_j)]\ ,
401 \label{eq:stickyPot}
402 \end{equation}
403 where $\nu_0$ is a strength parameter for the sticky potential, and
404 $s$ and $s^\prime$ are cubic switching functions which turn off the
405 sticky interaction beyond the first solvation shell. The $w$ function
406 can be thought of as an attractive potential with tetrahedral
407 geometry:
408 \begin{equation}
409 w({\bf r}_{ij},{\bf \Omega}_i,{\bf \Omega}_j)=
410 \sin\theta_{ij}\sin2\theta_{ij}\cos2\phi_{ij},
411 \label{eq:stickyW}
412 \end{equation}
413 while the $w^\prime$ function counters the normal aligned and
414 anti-aligned structures favored by point dipoles:
415 \begin{equation}
416 w^\prime({\bf r}_{ij},{\bf \Omega}_i,{\bf \Omega}_j)=
417 (\cos\theta_{ij}-0.6)^2(\cos\theta_{ij}+0.8)^2-w^0,
418 \label{eq:stickyWprime}
419 \end{equation}
420 It should be noted that $w$ is proportional to the sum of the $Y_3^2$
421 and $Y_3^{-2}$ spherical harmonics (a linear combination which
422 enhances the tetrahedral geometry for hydrogen bonded structures),
423 while $w^\prime$ is a purely empirical function. A more detailed
424 description of the functional parts and variables in this potential
425 can be found in the original SSD
426 articles.\cite{liu96:new_model,liu96:monte_carlo,chandra99:ssd_md,Ichiye03}
427
428 Since SSD is a single-point {\it dipolar} model, the force
429 calculations are simplified significantly relative to the standard
430 {\it charged} multi-point models. In the original Monte Carlo
431 simulations using this model, Ichiye {\it et al.} reported that using
432 SSD decreased computer time by a factor of 6-7 compared to other
433 models.\cite{liu96:new_model} What is most impressive is that these savings
434 did not come at the expense of accurate depiction of the liquid state
435 properties. Indeed, SSD maintains reasonable agreement with the Soper
436 diffraction data for the structural features of liquid
437 water.\cite{Soper86,liu96:new_model} Additionally, the dynamical properties
438 exhibited by SSD agree with experiment better than those of more
439 computationally expensive models (like TIP3P and
440 SPC/E).\cite{chandra99:ssd_md} The combination of speed and accurate depiction
441 of solvent properties makes SSD a very attractive model for the
442 simulation of large scale biochemical simulations.
443
444 Recent constant pressure simulations revealed issues in the original
445 SSD model that led to lower than expected densities at all target
446 pressures.\cite{Ichiye03,Gezelter04} The default model in {\sc oopse}
447 is therefore SSD/E, a density corrected derivative of SSD that
448 exhibits improved liquid structure and transport behavior. If the use
449 of a reaction field long-range interaction correction is desired, it
450 is recommended that the parameters be modified to those of the SSD/RF
451 model. Solvent parameters can be easily modified in an accompanying
452 {\sc BASS} file as illustrated in the scheme below. A table of the
453 parameter values and the drawbacks and benefits of the different
454 density corrected SSD models can be found in reference
455 \ref{Gezelter04}.
456
457 \begin{lstlisting}[caption={[A simulation of {\sc ssd} water]An example file showing a simulation including {\sc ssd} water.},label={sch:ssd}]
458
459 #include "water.mdl"
460
461 nComponents = 1;
462 component{
463 type = "SSD_water";
464 nMol = 864;
465 }
466
467 initialConfig = "liquidWater.init";
468
469 forceField = "DUFF";
470
471 /*
472 * The reactionField flag toggles reaction
473 * field corrections.
474 */
475
476 reactionField = false; // defaults to false
477 dielectric = 80.0; // dielectric for reaction field
478
479 /*
480 * The following two flags set the cutoff
481 * radius for the electrostatic forces
482 * as well as the skin thickness of the switching
483 * function.
484 */
485
486 electrostaticCutoffRadius = 9.2;
487 electrostaticSkinThickness = 1.38;
488
489 \end{lstlisting}
490
491
492 \subsection{\label{sec:eam}Embedded Atom Method}
493
494 Several other molecular dynamics packages\cite{dynamo86} exist which have the
495 capacity to simulate metallic systems, including some that have
496 parallel computational abilities\cite{plimpton93}. Potentials that
497 describe bonding transition metal
498 systems\cite{Finnis84,Ercolessi88,Chen90,Qi99,Ercolessi02} have a
499 attractive interaction which models ``Embedding''
500 a positively charged metal ion in the electron density due to the
501 free valance ``sea'' of electrons created by the surrounding atoms in
502 the system. A mostly repulsive pairwise part of the potential
503 describes the interaction of the positively charged metal core ions
504 with one another. A particular potential description called the
505 Embedded Atom Method\cite{Daw84,FBD86,johnson89,Lu97}({\sc eam}) that has
506 particularly wide adoption has been selected for inclusion in {\sc oopse}. A
507 good review of {\sc eam} and other metallic potential formulations was done
508 by Voter.\cite{voter}
509
510 The {\sc eam} potential has the form:
511 \begin{eqnarray}
512 V & = & \sum_{i} F_{i}\left[\rho_{i}\right] + \sum_{i} \sum_{j \neq i}
513 \phi_{ij}({\bf r}_{ij}) \\
514 \rho_{i} & = & \sum_{j \neq i} f_{j}({\bf r}_{ij})
515 \end{eqnarray}S
516
517 where $F_{i} $ is the embedding function that equates the energy required to embed a
518 positively-charged core ion $i$ into a linear superposition of
519 spherically averaged atomic electron densities given by
520 $\rho_{i}$. $\phi_{ij}$ is a primarily repulsive pairwise interaction
521 between atoms $i$ and $j$. In the original formulation of
522 {\sc eam} cite{Daw84}, $\phi_{ij}$ was an entirely repulsive term, however
523 in later refinements to EAM have shown that non-uniqueness between $F$
524 and $\phi$ allow for more general forms for $\phi$.\cite{Daw89}
525 There is a cutoff distance, $r_{cut}$, which limits the
526 summations in the {\sc eam} equation to the few dozen atoms
527 surrounding atom $i$ for both the density $\rho$ and pairwise $\phi$
528 interactions. Foiles et al. fit EAM potentials for fcc metals Cu, Ag, Au, Ni, Pd, Pt and alloys of these metals\cite{FDB86}. These potential fits are in the DYNAMO 86 format and are included with {\sc oopse}.
529
530
531 \subsection{\label{Sec:pbc}Periodic Boundary Conditions}
532
533 \newcommand{\roundme}{\operatorname{round}}
534
535 \textit{Periodic boundary conditions} are widely used to simulate truly
536 macroscopic systems with a relatively small number of particles. The
537 simulation box is replicated throughout space to form an infinite lattice.
538 During the simulation, when a particle moves in the primary cell, its image in
539 other boxes move in exactly the same direction with exactly the same
540 orientation.Thus, as a particle leaves the primary cell, one of its images
541 will enter through the opposite face.If the simulation box is large enough to
542 avoid \textquotedblleft feeling\textquotedblright\ the symmetries of the
543 periodic lattice, surface effects can be ignored. Cubic, orthorhombic and
544 parallelepiped are the available periodic cells In OOPSE. We use a matrix to
545 describe the property of the simulation box. Therefore, both the size and
546 shape of the simulation box can be changed during the simulation. The
547 transformation from box space vector $\mathbf{s}$ to its corresponding real
548 space vector $\mathbf{r}$ is defined by
549 \begin{equation}
550 \mathbf{r}=\underline{\mathbf{H}}\cdot\mathbf{s}%
551 \end{equation}
552
553
554 where $H=(h_{x},h_{y},h_{z})$ is a transformation matrix made up of the three
555 box axis vectors. $h_{x},h_{y}$ and $h_{z}$ represent the three sides of the
556 simulation box respectively.
557
558 To find the minimum image of a vector $\mathbf{r}$, we convert the real vector
559 to its corresponding vector in box space first, \bigskip%
560 \begin{equation}
561 \mathbf{s}=\underline{\mathbf{H}}^{-1}\cdot\mathbf{r}%
562 \end{equation}
563 And then, each element of $\mathbf{s}$ is wrapped to lie between -0.5 to 0.5,
564 \begin{equation}
565 s_{i}^{\prime}=s_{i}-\roundme(s_{i})
566 \end{equation}
567 where
568
569 %
570
571 \begin{equation}
572 \roundme(x)=\left\{
573 \begin{array}{cc}%
574 \lfloor{x+0.5}\rfloor & \text{if \ }x\geqslant0\\
575 \lceil{x-0.5}\rceil & \text{otherwise}%
576 \end{array}
577 \right.
578 \end{equation}
579
580
581 For example, $\roundme(3.6)=4$,$\roundme(3.1)=3$, $\roundme(-3.6)=-4$, $\roundme(-3.1)=-3$.
582
583 Finally, we obtain the minimum image coordinates $\mathbf{r}^{\prime}$ by
584 transforming back to real space,%
585
586 \begin{equation}
587 \mathbf{r}^{\prime}=\underline{\mathbf{H}}^{-1}\cdot\mathbf{s}^{\prime}%
588 \end{equation}
589