ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/oopsePaper/EmpericalEnergy.tex
Revision: 961
Committed: Mon Jan 19 17:24:52 2004 UTC (20 years, 5 months ago) by chrisfen
Content type: application/x-tex
File size: 21051 byte(s)
Log Message:
Adjusted the SSD part of the paper...

File Contents

# Content
1
2 \section{\label{sec:empericalEnergy}The Emperical 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 parameters
7 describing the atom are generalized to make the atom as flexible a
8 representation as possible. They may represent specific atoms of an
9 element, or be used for collections of atoms such as a methyl
10 group. The atoms are also capable of having a directional component
11 associated with them, often in the form of a dipole. Charges on atoms
12 are not currently suported by {\sc oopse}.
13
14 The second most basic building block of a simulation is the
15 molecule. The molecule is a way for {\sc oopse} to keep track of the
16 atoms in a simulation in logical manner. This particular unit will
17 store the identities of all the atoms associated with itself and is
18 responsible for the evaluation of its own bonded interaction
19 (i.e.~bonds, bends, and torsions).
20
21 As stated previously, one of the features that sets {\sc oopse} apart
22 from most of the current molecular simulation packages is the ability
23 to handle rigid body dynamics. Rigid bodies are non-spherical
24 particles or collections of particles that have a constant internal
25 potential and move collectively.\cite{Goldstein01} They are not
26 included in most simulation packages because of the requirement to
27 propagate the orientational degrees of freedom. Until recently,
28 integrators which propagate orientational motion have been lacking.
29
30 Moving a rigid body involves determination of both the force and
31 torque applied by the surroundings, which directly affect the
32 translational and rotational motion in turn. In order to accumulate
33 the total force on a rigid body, the external forces and torques must
34 first be calculated for all the internal particles. The total force on
35 the rigid body is simply the sum of these external forces.
36 Accumulation of the total torque on the rigid body is more complex
37 than the force in that it is the torque applied on the center of mass
38 that dictates rotational motion. The torque on rigid body {\it i} is
39 \begin{equation}
40 \boldsymbol{\tau}_i=
41 \sum_{a}(\mathbf{r}_{ia}-\mathbf{r}_i)\times \mathbf{f}_{ia}
42 + \boldsymbol{\tau}_{ia},
43 \label{eq:torqueAccumulate}
44 \end{equation}
45 where $\boldsymbol{\tau}_i$ and $\mathbf{r}_i$ are the torque on and
46 position of the center of mass respectively, while $\mathbf{f}_{ia}$,
47 $\mathbf{r}_{ia}$, and $\boldsymbol{\tau}_{ia}$ are the force on,
48 position of, and torque on the component particles of the rigid body.
49
50 The summation of the total torque is done in the body fixed axis of
51 the rigid body. In order to move between the space fixed and body
52 fixed coordinate axes, parameters describing the orientation must be
53 maintained for each rigid body. At a minimum, the rotation matrix
54 (\textbf{A}) can be described by the three Euler angles ($\phi,
55 \theta,$ and $\psi$), where the elements of \textbf{A} are composed of
56 trigonometric operations involving $\phi, \theta,$ and
57 $\psi$.\cite{Goldstein01} In order to avoid numerical instabilities
58 inherent in using the Euler angles, the four parameter ``quaternion''
59 scheme is often used. The elements of \textbf{A} can be expressed as
60 arithmetic operations involving the four quaternions ($q_0, q_1, q_2,$
61 and $q_3$).\cite{allen87:csl} Use of quaternions also leads to
62 performance enhancements, particularly for very small
63 systems.\cite{Evans77}
64
65 {\sc oopse} utilizes a relatively new scheme that propagates the
66 entire nine parameter rotation matrix internally. Further discussion
67 on this choice can be found in Sec.~\ref{sec:integrate}.
68
69 \subsection{\label{sec:LJPot}The Lennard Jones Potential}
70
71 The most basic force field implemented in OOPSE is the Lennard-Jones
72 potential. The Lennard-Jones potential. Which mimics the Van der Waals
73 interaction at long distances, and uses an emperical repulsion at
74 short distances. The Lennard-Jones potential is given by:
75 \begin{equation}
76 V_{\text{LJ}}(r_{ij}) =
77 4\epsilon_{ij} \biggl[
78 \biggl(\frac{\sigma_{ij}}{r_{ij}}\biggr)^{12}
79 - \biggl(\frac{\sigma_{ij}}{r_{ij}}\biggr)^{6}
80 \biggr]
81 \label{eq:lennardJonesPot}
82 \end{equation}
83 Where $r_{ij}$ is the distance between particle $i$ and $j$,
84 $\sigma_{ij}$ scales the length of the interaction, and
85 $\epsilon_{ij}$ scales the well depth of the potential.
86
87 Because this potential is calculated between all pairs, the force
88 evaluation can become computationally expensive for large systems. To
89 keep the pair evaluation to a manegable number, OOPSE employs a
90 cut-off radius.\cite{allen87:csl} The cutoff radius is set to be
91 $2.5\sigma_{ii}$, where $\sigma_{ii}$ is the largest Lennard-Jones length
92 parameter in the system. Truncating the calculation at
93 $r_{\text{cut}}$ introduces a discontinuity into the potential
94 energy. To offset this discontinuity, the energy value at
95 $r_{\text{cut}}$ is subtracted from the entire potential. This causes
96 the potential to go to zero at the cut-off radius.
97
98 Interactions between dissimilar particles requires the generation of
99 cross term parameters for $\sigma$ and $\epsilon$. These are
100 calculated through the Lorentz-Berthelot mixing
101 rules:\cite{allen87:csl}
102 \begin{equation}
103 \sigma_{ij} = \frac{1}{2}[\sigma_{ii} + \sigma_{jj}]
104 \label{eq:sigmaMix}
105 \end{equation}
106 and
107 \begin{equation}
108 \epsilon_{ij} = \sqrt{\epsilon_{ii} \epsilon_{jj}}
109 \label{eq:epsilonMix}
110 \end{equation}
111
112
113 \subsection{\label{sec:DUFF}Dipolar Unified-Atom Force Field}
114
115 The Dipolar Unified-atom Force Field ({\sc duff}) was developed to
116 simulate lipid bilayers. The systems require a model capable of forming
117 bilayers, while still being sufficiently computationally efficient to
118 allow simulations of large systems ($\approx$100's of phospholipids,
119 $\approx$1000's of waters) for long times ($\approx$10's of
120 nanoseconds).
121
122 With this goal in mind, {\sc duff} has no point charges. Charge
123 neutral distributions were replaced with dipoles, while most atoms and
124 groups of atoms were reduced to Lennard-Jones interaction sites. This
125 simplification cuts the length scale of long range interactions from
126 $\frac{1}{r}$ to $\frac{1}{r^3}$, allowing us to avoid the
127 computationally expensive Ewald sum. Instead, we can use
128 neighbor-lists, reaction field, and cutoff radii for the dipolar
129 interactions.
130
131 As an example, lipid head-groups in {\sc duff} are represented as
132 point dipole interaction sites. By placing a dipole of 20.6~Debye at
133 the head group center of mass, our model mimics the head group of
134 phosphatidylcholine.\cite{Cevc87} Additionally, a Lennard-Jones site
135 is located at the pseudoatom's center of mass. The model is
136 illustrated by the dark grey atom in Fig.~\ref{fig:lipidModel}. The water model we use to complement the dipoles of the lipids is out
137 repaarameterization of the soft sticky dipole (SSD) model of Ichiye
138 \emph{et al.}\cite{liu96:new_model}
139
140 \begin{figure}
141 \epsfxsize=\linewidth
142 \epsfbox{lipidModel.eps}
143 \caption{A representation of the lipid model. $\phi$ is the torsion angle, $\theta$ %
144 is the bend angle, $\mu$ is the dipole moment of the head group, and n
145 is the chain length.}
146 \label{fig:lipidModel}
147 \end{figure}
148
149 Turning to the tails of the phospholipids, we have used a set of
150 scalable parameters to model the alkyl groups with Lennard-Jones
151 sites. For this, we have used the TraPPE force field of Siepmann
152 \emph{et al}.\cite{Siepmann1998} TraPPE is a unified-atom
153 representation of n-alkanes, which is parametrized against phase
154 equilibria using Gibbs Monte Carlo simulation
155 techniques.\cite{Siepmann1998} One of the advantages of TraPPE is that
156 it generalizes the types of atoms in an alkyl chain to keep the number
157 of pseudoatoms to a minimum; the parameters for an atom such as
158 $\text{CH}_2$ do not change depending on what species are bonded to
159 it.
160
161 TraPPE also constrains of all bonds to be of fixed length. Typically,
162 bond vibrations are the fastest motions in a molecular dynamic
163 simulation. Small time steps between force evaluations must be used to
164 ensure adequate sampling of the bond potential conservation of
165 energy. By constraining the bond lengths, larger time steps may be
166 used when integrating the equations of motion.
167
168
169 \subsubsection{\label{subSec:energyFunctions}{\sc duff} Energy Functions}
170
171 The total energy of function in {\sc duff} is given by the following:
172 \begin{equation}
173 V_{\text{Total}} = \sum^{N}_{I=1} V^{I}_{\text{Internal}}
174 + \sum^{N}_{I=1} \sum_{J>I} V^{IJ}_{\text{Cross}}
175 \label{eq:totalPotential}
176 \end{equation}
177 Where $V^{I}_{\text{Internal}}$ is the internal potential of a molecule:
178 \begin{equation}
179 V^{I}_{\text{Internal}} =
180 \sum_{\theta_{ijk} \in I} V_{\text{bend}}(\theta_{ijk})
181 + \sum_{\phi_{ijkl} \in I} V_{\text{torsion}}(\phi_{ijkl})
182 + \sum_{i \in I} \sum_{(j>i+4) \in I}
183 \biggl[ V_{\text{LJ}}(r_{ij}) + V_{\text{dipole}}
184 (\mathbf{r}_{ij},\boldsymbol{\Omega}_{i},\boldsymbol{\Omega}_{j})
185 \biggr]
186 \label{eq:internalPotential}
187 \end{equation}
188 Here $V_{\text{bend}}$ is the bend potential for all 1, 3 bonded pairs
189 within the molecule, and $V_{\text{torsion}}$ is the torsion potential
190 for all 1, 4 bonded pairs. The pairwise portions of the internal
191 potential are excluded for pairs that are closer than three bonds,
192 i.e.~atom pairs farther away than a torsion are included in the
193 pair-wise loop.
194
195
196 The bend potential of a molecule is represented by the following function:
197 \begin{equation}
198 V_{\theta_{ijk}} = k_{\theta}( \theta_{ijk} - \theta_0 )^2 \label{eq:bendPot}
199 \end{equation}
200 Where $\theta_{ijk}$ is the angle defined by atoms $i$, $j$, and $k$
201 (see Fig.~\ref{fig:lipidModel}), and $\theta_0$ is the equilibrium
202 bond angle. $k_{\theta}$ is the force constant which determines the
203 strength of the harmonic bend. The parameters for $k_{\theta}$ and
204 $\theta_0$ are based off of those in TraPPE.\cite{Siepmann1998}
205
206 The torsion potential and parameters are also taken from TraPPE. It is
207 of the form:
208 \begin{equation}
209 V_{\text{torsion}}(\phi) = c_1[1 + \cos \phi]
210 + c_2[1 + \cos(2\phi)]
211 + c_3[1 + \cos(3\phi)]
212 \label{eq:origTorsionPot}
213 \end{equation}
214 Here $\phi_{ijkl}$ is the angle defined by four bonded neighbors $i$,
215 $j$, $k$, and $l$ (again, see Fig.~\ref{fig:lipidModel}). For
216 computaional efficency, the torsion potential has been recast after
217 the method of CHARMM\cite{charmm1983} whereby the angle series is
218 converted to a power series of the form:
219 \begin{equation}
220 V_{\text{torsion}}(\phi_{ijkl}) =
221 k_3 \cos^3 \phi + k_2 \cos^2 \phi + k_1 \cos \phi + k_0
222 \label{eq:torsionPot}
223 \end{equation}
224 Where:
225 \begin{align*}
226 k_0 &= c_1 + c_3 \\
227 k_1 &= c_1 - 3c_3 \\
228 k_2 &= 2 c_2 \\
229 k_3 &= 4c_3
230 \end{align*}
231 By recasting the equation to a power series, repeated trigonometric
232 evaluations are avoided during the calculation of the potential.
233
234
235 The cross portion of the total potential, $V^{IJ}_{\text{Cross}}$, is
236 as follows:
237 \begin{equation}
238 V^{IJ}_{\text{Cross}} =
239 \sum_{i \in I} \sum_{j \in J}
240 \biggl[ V_{\text{LJ}}(r_{ij}) + V_{\text{dipole}}
241 (\mathbf{r}_{ij},\boldsymbol{\Omega}_{i},\boldsymbol{\Omega}_{j})
242 + V_{\text{sticky}}
243 (\mathbf{r}_{ij},\boldsymbol{\Omega}_{i},\boldsymbol{\Omega}_{j})
244 \biggr]
245 \label{eq:crossPotentail}
246 \end{equation}
247 Where $V_{\text{LJ}}$ is the Lennard Jones potential,
248 $V_{\text{dipole}}$ is the dipole dipole potential, and
249 $V_{\text{sticky}}$ is the sticky potential defined by the SSD
250 model. Note that not all atom types include all interactions.
251
252 The dipole-dipole potential has the following form:
253 \begin{equation}
254 V_{\text{dipole}}(\mathbf{r}_{ij},\boldsymbol{\Omega}_{i},
255 \boldsymbol{\Omega}_{j}) = \frac{|\mu_i||\mu_j|}{4\pi\epsilon_{0}r_{ij}^{3}} \biggl[
256 \boldsymbol{\hat{u}}_{i} \cdot \boldsymbol{\hat{u}}_{j}
257 -
258 \frac{3(\boldsymbol{\hat{u}}_i \cdot \mathbf{r}_{ij}) %
259 (\boldsymbol{\hat{u}}_j \cdot \mathbf{r}_{ij}) }
260 {r^{2}_{ij}} \biggr]
261 \label{eq:dipolePot}
262 \end{equation}
263 Here $\mathbf{r}_{ij}$ is the vector starting at atom $i$ pointing
264 towards $j$, and $\boldsymbol{\Omega}_i$ and $\boldsymbol{\Omega}_j$
265 are the Euler angles of atom $i$ and $j$ respectively. $|\mu_i|$ is
266 the magnitude of the dipole moment of atom $i$ and
267 $\boldsymbol{\hat{u}}_i$ is the standard unit orientation vector of
268 $\boldsymbol{\Omega}_i$.
269
270
271 \subsection{\label{sec:SSD}The {\sc DUFF} Water Models: SSD/E and SSD/RF}
272
273 In the interest of computational efficiency, the default solvent used
274 by {\sc oopse} is the extended Soft Sticky Dipole (SSD/E) water
275 model.\cite{Gezelter04} The original SSD was developed by Ichiye
276 \emph{et al.}\cite{Ichiye96} as a modified form of the hard-sphere
277 water model proposed by Bratko, Blum, and
278 Luzar.\cite{Bratko85,Bratko95} It consists of a single point dipole
279 with a Lennard-Jones core and a sticky potential that directs the
280 particles to assume the proper hydrogen bond orientation in the first
281 solvation shell. Thus, the interaction between two SSD water molecules
282 \emph{i} and \emph{j} is given by the potential
283 \begin{equation}
284 V_{ij} =
285 V_{ij}^{LJ} (r_{ij})\ + V_{ij}^{dp}
286 (\mathbf{r}_{ij},\boldsymbol{\Omega}_i,\boldsymbol{\Omega}_j)\ +
287 V_{ij}^{sp}
288 (\mathbf{r}_{ij},\boldsymbol{\Omega}_i,\boldsymbol{\Omega}_j),
289 \label{eq:ssdPot}
290 \end{equation}
291 where the $\mathbf{r}_{ij}$ is the position vector between molecules
292 \emph{i} and \emph{j} with magnitude equal to the distance $r_{ij}$, and
293 $\boldsymbol{\Omega}_i$ and $\boldsymbol{\Omega}_j$ represent the
294 orientations of the respective molecules. The Lennard-Jones and dipole
295 parts of the potential are given by equations \ref{eq:lennardJonesPot}
296 and \ref{eq:dipolePot} respectively. The sticky part is described by
297 the following,
298 \begin{equation}
299 u_{ij}^{sp}(\mathbf{r}_{ij},\boldsymbol{\Omega}_i,\boldsymbol{\Omega}_j)=
300 \frac{\nu_0}{2}[s(r_{ij})w(\mathbf{r}_{ij},
301 \boldsymbol{\Omega}_i,\boldsymbol{\Omega}_j) +
302 s^\prime(r_{ij})w^\prime(\mathbf{r}_{ij},
303 \boldsymbol{\Omega}_i,\boldsymbol{\Omega}_j)]\ ,
304 \label{eq:stickyPot}
305 \end{equation}
306 where $\nu_0$ is a strength parameter for the sticky potential, and
307 $s$ and $s^\prime$ are cubic switching functions which turn off the
308 sticky interaction beyond the first solvation shell. The $w$ function
309 can be thought of as an attractive potential with tetrahedral
310 geometry:
311 \begin{equation}
312 w({\bf r}_{ij},{\bf \Omega}_i,{\bf \Omega}_j)=
313 \sin\theta_{ij}\sin2\theta_{ij}\cos2\phi_{ij},
314 \label{eq:stickyW}
315 \end{equation}
316 while the $w^\prime$ function counters the normal aligned and
317 anti-aligned structures favored by point dipoles:
318 \begin{equation}
319 w^\prime({\bf r}_{ij},{\bf \Omega}_i,{\bf \Omega}_j)=
320 (\cos\theta_{ij}-0.6)^2(\cos\theta_{ij}+0.8)^2-w^0,
321 \label{eq:stickyWprime}
322 \end{equation}
323 It should be noted that $w$ is proportional to the sum of the $Y_3^2$
324 and $Y_3^{-2}$ spherical harmonics (a linear combination which
325 enhances the tetrahedral geometry for hydrogen bonded structures),
326 while $w^\prime$ is a purely empirical function. A more detailed
327 description of the functional parts and variables in this potential
328 can be found in the original SSD
329 articles.\cite{Ichiye96,Ichiye96b,Ichiye99,Ichiye03}
330
331 Since SSD is a single-point {\it dipolar} model, the force
332 calculations are simplified significantly relative to the standard
333 {\it charged} multi-point models. In the original Monte Carlo
334 simulations using this model, Ichiye {\it et al.} reported that using
335 SSD decreased computer time by a factor of 6-7 compared to other
336 models.\cite{Ichiye96} What is most impressive is that these savings
337 did not come at the expense of accurate depiction of the liquid state
338 properties. Indeed, SSD maintains reasonable agreement with the Soper
339 diffraction data for the structural features of liquid
340 water.\cite{Soper86,Ichiye96} Additionally, the dynamical properties
341 exhibited by SSD agree with experiment better than those of more
342 computationally expensive models (like TIP3P and
343 SPC/E).\cite{Ichiye99} The combination of speed and accurate depiction
344 of solvent properties makes SSD a very attractive model for the
345 simulation of large scale biochemical simulations.
346
347 Recent constant pressure simulations revealed issues in the original
348 SSD model that led to lower than expected densities at all target
349 pressures.\cite{Ichiye03,Gezelter04} The default model in {\sc oopse}
350 is therefore SSD/E, a density corrected derivative of SSD that
351 exhibits improved liquid structure and transport behavior. If the use
352 of a reaction field long-range interaction correction is desired, it
353 is recommended that the parameters be modified to those of the SSD/RF
354 model. Solvent parameters can be easily modified in an accompanying
355 {\sc BASS} file as illustrated in the scheme below. A table of the
356 parameter values and the drawbacks and benefits of the different
357 density corrected SSD models can be found in reference
358 \ref{Gezelter04}.
359
360 !!!Place a {\sc BASS} scheme showing SSD parameters around here!!!
361
362 \subsection{\label{sec:eam}Embedded Atom Method}
363
364 Several other molecular dynamics packages\cite{dynamo86} exist which have the
365 capacity to simulate metallic systems, including some that have
366 parallel computational abilities\cite{plimpton93}. Potentials that
367 describe bonding transition metal
368 systems\cite{Finnis84,Ercolessi88,Chen90,Qi99,Ercolessi02} have a
369 attractive interaction which models ``Embedding''
370 a positively charged metal ion in the electron density due to the
371 free valance ``sea'' of electrons created by the surrounding atoms in
372 the system. A mostly repulsive pairwise part of the potential
373 describes the interaction of the positively charged metal core ions
374 with one another. A particular potential description called the
375 Embedded Atom Method\cite{Daw84,FBD86,johnson89,Lu97}({\sc eam}) that has
376 particularly wide adoption has been selected for inclusion in OOPSE. A
377 good review of {\sc eam} and other metallic potential formulations was done
378 by Voter.\cite{voter}
379
380 The {\sc eam} potential has the form:
381 \begin{eqnarray}
382 V & = & \sum_{i} F_{i}\left[\rho_{i}\right] + \sum_{i} \sum_{j \neq i}
383 \phi_{ij}({\bf r}_{ij}) \\
384 \rho_{i} & = & \sum_{j \neq i} f_{j}({\bf r}_{ij})
385 \end{eqnarray}S
386
387 where $F_{i} $ is the embedding function that equates the energy required to embed a
388 positively-charged core ion $i$ into a linear superposition of
389 sperically averaged atomic electron densities given by
390 $\rho_{i}$. $\phi_{ij}$ is a primarily repulsive pairwise interaction
391 between atoms $i$ and $j$. In the original formulation of
392 {\sc eam} cite{Daw84}, $\phi_{ij}$ was an entirely repulsive term, however
393 in later refinements to EAM have shown that non-uniqueness between $F$
394 and $\phi$ allow for more general forms for $\phi$.\cite{Daw89}
395 There is a cutoff distance, $r_{cut}$, which limits the
396 summations in the {\sc eam} equation to the few dozen atoms
397 surrounding atom $i$ for both the density $\rho$ and pairwise $\phi$
398 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}.
399
400
401 \subsection{\label{Sec:pbc}Periodic Boundary Conditions}
402
403 \newcommand{\roundme}{\operatorname{round}}
404
405 \textit{Periodic boundary conditions} are widely used to simulate truly
406 macroscopic systems with a relatively small number of particles. The
407 simulation box is replicated throughout space to form an infinite
408 lattice. During the simulation, when a particle moves in the primary
409 cell, its image in other boxes move in exactly the same direction with
410 exactly the same orientation.Thus, as a particle leaves the primary
411 cell, one of its images will enter through the opposite face.If the
412 simulation box is large enough to avoid "feeling" the symmetries of
413 the periodic lattice, surface effects can be ignored. Cubic,
414 orthorhombic and parallelepiped are the available periodic cells In
415 OOPSE. We use a matrix to describe the property of the simulation
416 box. Therefore, both the size and shape of the simulation box can be
417 changed during the simulation. The transformation from box space
418 vector $\mathbf{s}$ to its corresponding real space vector
419 $\mathbf{r}$ is defined by
420 \begin{equation}
421 \mathbf{r}=\underline{\underline{H}}\cdot\mathbf{s}%
422 \end{equation}
423
424
425 where $H=(h_{x},h_{y},h_{z})$ is a transformation matrix made up of
426 the three box axis vectors. $h_{x},h_{y}$ and $h_{z}$ represent the
427 three sides of the simulation box respectively.
428
429 To find the minimum image, we convert the real vector to its
430 corresponding vector in box space first, \bigskip%
431 \begin{equation}
432 \mathbf{s}=\underline{\underline{H}}^{-1}\cdot\mathbf{r}%
433 \end{equation}
434 And then, each element of $\mathbf{s}$ is wrapped to lie between -0.5 to 0.5,
435 \begin{equation}
436 s_{i}^{\prime}=s_{i}-\roundme(s_{i})
437 \end{equation}
438 where
439
440 %
441
442 \begin{equation}
443 \roundme(x)=\left\{
444 \begin{array}{cc}
445 \lfloor{x+0.5}\rfloor & \text{if \ }x\geqslant0\\
446 \lceil{x-0.5}\rceil & \text{otherwise}%
447 \end{array}
448 \right.
449 \end{equation}
450 For example, $\roundme(3.6)=4$, $\roundme(3.1)=3$, $\roundme(-3.6)=-4$,
451 $\roundme(-3.1)=-3$.
452
453 Finally, we obtain the minimum image coordinates by transforming back
454 to real space,%
455
456 \begin{equation}
457 \mathbf{r}^{\prime}=\underline{\underline{H}}^{-1}\cdot\mathbf{s}^{\prime}%
458 \end{equation}
459