ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/oopsePaper/EmpericalEnergy.tex
(Generate patch)

Comparing trunk/oopsePaper/EmpericalEnergy.tex (file contents):
Revision 899 by mmeineke, Tue Jan 6 18:53:58 2004 UTC vs.
Revision 964 by mmeineke, Mon Jan 19 21:03:47 2004 UTC

# Line 1 | Line 1
1  
2   \section{\label{sec:empericalEnergy}The Emperical Energy Functions}
3  
4 < \subsection{\label{sec:atomsMolecules}Atoms and Molecules}
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
# Line 9 | Line 9 | are not currently suporrted by {\sc oopse}.
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 suporrted by {\sc oopse}.
12 > are not currently suported by {\sc oopse}.
13  
14 + \begin{lstlisting}[caption={[Specifier for molecules and atoms] An example specifing 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   The second most basic building block of a simulation is the
26 < molecule. The molecule is a way for {\sc oopse} to keep track of the atoms
27 < in a simulation in logical manner. This particular unit will store the
28 < identities of all the atoms associated with itself and is responsible
29 < for the evaluation of its own bonded interaction (i.e.~bonds, bends,
30 < and torsions).
26 > molecule. The molecule is a way for {\sc oopse} to keep track of the
27 > atoms in a simulation in logical manner. This particular unit will
28 > store the identities of all the atoms associated with itself and is
29 > responsible for the evaluation of its own bonded interaction
30 > (i.e.~bonds, bends, and torsions).
31  
32 + As stated previously, one of the features that sets {\sc oopse} apart
33 + from most of the current molecular simulation packages is the ability
34 + to handle rigid body dynamics. Rigid bodies are non-spherical
35 + particles or collections of particles that have a constant internal
36 + potential and move collectively.\cite{Goldstein01} They are not
37 + included in most simulation packages because of the requirement to
38 + propagate the orientational degrees of freedom. Until recently,
39 + integrators which propagate orientational motion have been lacking.
40 +
41 + Moving a rigid body involves determination of both the force and
42 + torque applied by the surroundings, which directly affect the
43 + translational and rotational motion in turn. In order to accumulate
44 + the total force on a rigid body, the external forces and torques must
45 + first be calculated for all the internal particles. The total force on
46 + the rigid body is simply the sum of these external forces.
47 + Accumulation of the total torque on the rigid body is more complex
48 + than the force in that it is the torque applied on the center of mass
49 + that dictates rotational motion. The torque on rigid body {\it i} is
50 + \begin{equation}
51 + \boldsymbol{\tau}_i=
52 +        \sum_{a}(\mathbf{r}_{ia}-\mathbf{r}_i)\times \mathbf{f}_{ia}
53 +        + \boldsymbol{\tau}_{ia},
54 + \label{eq:torqueAccumulate}
55 + \end{equation}
56 + where $\boldsymbol{\tau}_i$ and $\mathbf{r}_i$ are the torque on and
57 + position of the center of mass respectively, while $\mathbf{f}_{ia}$,
58 + $\mathbf{r}_{ia}$, and $\boldsymbol{\tau}_{ia}$ are the force on,
59 + position of, and torque on the component particles of the rigid body.
60 +
61 + The summation of the total torque is done in the body fixed axis of
62 + the rigid body. In order to move between the space fixed and body
63 + fixed coordinate axes, parameters describing the orientation must be
64 + maintained for each rigid body. At a minimum, the rotation matrix
65 + (\textbf{A}) can be described by the three Euler angles ($\phi,
66 + \theta,$ and $\psi$), where the elements of \textbf{A} are composed of
67 + trigonometric operations involving $\phi, \theta,$ and
68 + $\psi$.\cite{Goldstein01} In order to avoid numerical instabilities
69 + inherent in using the Euler angles, the four parameter ``quaternion''
70 + scheme is often used. The elements of \textbf{A} can be expressed as
71 + arithmetic operations involving the four quaternions ($q_0, q_1, q_2,$
72 + and $q_3$).\cite{allen87:csl} Use of quaternions also leads to
73 + performance enhancements, particularly for very small
74 + systems.\cite{Evans77}
75 +
76 + {\sc oopse} utilizes a relatively new scheme that propagates the
77 + entire nine parameter rotation matrix internally. Further discussion
78 + on this choice can be found in Sec.~\ref{sec:integrate}.
79 +
80 + \subsection{\label{sec:LJPot}The Lennard Jones Potential}
81 +
82 + The most basic force field implemented in OOPSE is the Lennard-Jones
83 + potential. The Lennard-Jones potential. Which mimics the Van der Waals
84 + interaction at long distances, and uses an emperical repulsion at
85 + short distances. The Lennard-Jones potential is given by:
86 + \begin{equation}
87 + V_{\text{LJ}}(r_{ij}) =
88 +        4\epsilon_{ij} \biggl[
89 +        \biggl(\frac{\sigma_{ij}}{r_{ij}}\biggr)^{12}
90 +        - \biggl(\frac{\sigma_{ij}}{r_{ij}}\biggr)^{6}
91 +        \biggr]
92 + \label{eq:lennardJonesPot}
93 + \end{equation}
94 + Where $r_{ij}$ is the distance between particle $i$ and $j$,
95 + $\sigma_{ij}$ scales the length of the interaction, and
96 + $\epsilon_{ij}$ scales the well depth of the potential.
97 +
98 + Because this potential is calculated between all pairs, the force
99 + evaluation can become computationally expensive for large systems. To
100 + keep the pair evaluation to a manegable number, OOPSE employs a
101 + cut-off radius.\cite{allen87:csl} The cutoff radius is set to be
102 + $2.5\sigma_{ii}$, where $\sigma_{ii}$ is the largest Lennard-Jones length
103 + parameter in the system. Truncating the calculation at
104 + $r_{\text{cut}}$ introduces a discontinuity into the potential
105 + energy. To offset this discontinuity, the energy value at
106 + $r_{\text{cut}}$ is subtracted from the entire potential. This causes
107 + the potential to go to zero at the cut-off radius.
108 +
109 + Interactions between dissimilar particles requires the generation of
110 + cross term parameters for $\sigma$ and $\epsilon$. These are
111 + calculated through the Lorentz-Berthelot mixing
112 + rules:\cite{allen87:csl}
113 + \begin{equation}
114 + \sigma_{ij} = \frac{1}{2}[\sigma_{ii} + \sigma_{jj}]
115 + \label{eq:sigmaMix}
116 + \end{equation}
117 + and
118 + \begin{equation}
119 + \epsilon_{ij} = \sqrt{\epsilon_{ii} \epsilon_{jj}}
120 + \label{eq:epsilonMix}
121 + \end{equation}
122 +
123 +
124   \subsection{\label{sec:DUFF}Dipolar Unified-Atom Force Field}
125  
126 < The \underline{D}ipolar \underline{U}nified-Atom
127 < \underline{F}orce \underline{F}ield ({\sc duff}) was developed to
25 < simulate lipid bilayers. We needed a model capable of forming
126 > The Dipolar Unified-atom Force Field ({\sc duff}) was developed to
127 > simulate lipid bilayers. The systems require a model capable of forming
128   bilayers, while still being sufficiently computationally efficient to
129   allow simulations of large systems ($\approx$100's of phospholipids,
130   $\approx$1000's of waters) for long times ($\approx$10's of
131   nanoseconds).
132  
133 < With this goal in mind, we have eliminated all point charges. Charge
134 < distributions were replaced with dipoles, and charge-neutral
135 < distributions were reduced to Lennard-Jones interaction sites. This
133 > With this goal in mind, {\sc duff} has no point charges. Charge
134 > neutral distributions were replaced with dipoles, while most atoms and
135 > groups of atoms were reduced to Lennard-Jones interaction sites. This
136   simplification cuts the length scale of long range interactions from
137   $\frac{1}{r}$ to $\frac{1}{r^3}$, allowing us to avoid the
138 < computationally expensive Ewald-Sum. Instead, we can use
139 < neighbor-lists and cutoff radii for the dipolar interactions.
138 > computationally expensive Ewald sum. Instead, we can use
139 > neighbor-lists, reaction field, and cutoff radii for the dipolar
140 > interactions.
141  
142 < As an example, lipid head groups in {\sc duff} are represented as point
143 < dipole interaction sites.PC and PE Lipid head groups are typically
144 < zwitterionic in nature, with charges separated by as much as
145 < 6~$\mbox{\AA}$. By placing a dipole of 20.6~Debye at the head group
146 < center of mass, our model mimics the head group of PC.\cite{Cevc87}
147 < Additionally, a Lennard-Jones site is located at the
148 < pseudoatom's center of mass. The model is illustrated by the dark grey
149 < atom in Fig.~\ref{fig:lipidModel}.
142 > As an example, lipid head-groups in {\sc duff} are represented as
143 > point dipole interaction sites. By placing a dipole of 20.6~Debye at
144 > the head group center of mass, our model mimics the head group of
145 > phosphatidylcholine.\cite{Cevc87} Additionally, a Lennard-Jones site
146 > is located at the pseudoatom's center of mass. The model is
147 > 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
148 > repaarameterization of the soft sticky dipole (SSD) model of Ichiye
149 > \emph{et al.}\cite{liu96:new_model}
150  
151   \begin{figure}
152 < \epsfxsize=6in
153 < \epsfbox{lipidModel.epsi}
152 > \epsfxsize=\linewidth
153 > \epsfbox{lipidModel.eps}
154   \caption{A representation of the lipid model. $\phi$ is the torsion angle, $\theta$ %
155 < is the bend angle, $\mu$ is the dipole moment of the head group, and n is the chain length.}
155 > is the bend angle, $\mu$ is the dipole moment of the head group, and n
156 > is the chain length.}
157   \label{fig:lipidModel}
158   \end{figure}
159  
56 The water model we use to complement the dipoles of the lipids is
57 the soft sticky dipole (SSD) model of Ichiye \emph{et
58 al.}\cite{liu96:new_model} This model is discussed in greater detail
59 in Sec.~\ref{sec:SSD}. In all cases we reduce water to a single
60 Lennard-Jones interaction site. The site also contains a dipole to
61 mimic the partial charges on the hydrogens and the oxygen. However,
62 what makes the SSD model unique is the inclusion of a tetrahedral
63 short range potential to recover the hydrogen bonding of water, an
64 important factor when modeling bilayers, as it has been shown that
65 hydrogen bond network formation is a leading contribution to the
66 entropic driving force towards lipid bilayer formation.\cite{Cevc87}
67
68
160   Turning to the tails of the phospholipids, we have used a set of
161   scalable parameters to model the alkyl groups with Lennard-Jones
162   sites. For this, we have used the TraPPE force field of Siepmann
# Line 91 | Line 182 | V_{\text{Total}} = \sum^{N}_{I=1} V^{I}_{\text{Interna
182   The total energy of function in {\sc duff} is given by the following:
183   \begin{equation}
184   V_{\text{Total}} = \sum^{N}_{I=1} V^{I}_{\text{Internal}}
185 <        + \sum^{N}_{I=1} \sum^{N}_{J=1} V^{IJ}_{\text{Cross}}
185 >        + \sum^{N}_{I=1} \sum_{J>I} V^{IJ}_{\text{Cross}}
186   \label{eq:totalPotential}
187   \end{equation}
188   Where $V^{I}_{\text{Internal}}$ is the internal potential of a molecule:
189   \begin{equation}
190   V^{I}_{\text{Internal}} =
191          \sum_{\theta_{ijk} \in I} V_{\text{bend}}(\theta_{ijk})
192 <        + \sum_{\phi_{ijkl} \in I} V_{\text{torsion}}(\theta_{ijkl})
192 >        + \sum_{\phi_{ijkl} \in I} V_{\text{torsion}}(\phi_{ijkl})
193          + \sum_{i \in I} \sum_{(j>i+4) \in I}
194          \biggl[ V_{\text{LJ}}(r_{ij}) +  V_{\text{dipole}}
195          (\mathbf{r}_{ij},\boldsymbol{\Omega}_{i},\boldsymbol{\Omega}_{j})
# Line 106 | Line 197 | within in the molecule. $V_{\text{torsion}}$ is the to
197   \label{eq:internalPotential}
198   \end{equation}
199   Here $V_{\text{bend}}$ is the bend potential for all 1, 3 bonded pairs
200 < within in the molecule. $V_{\text{torsion}}$ is the torsion The
201 < pairwise portions of the internal potential are excluded for pairs
202 < that ar closer than three bonds, i.e.~atom pairs farther away than a
203 < torsion are included in the pair-wise loop.
200 > within the molecule, and $V_{\text{torsion}}$ is the torsion potential
201 > for all 1, 4 bonded pairs. The pairwise portions of the internal
202 > potential are excluded for pairs that are closer than three bonds,
203 > i.e.~atom pairs farther away than a torsion are included in the
204 > pair-wise loop.
205  
114 The cross portion of the total potential, $V^{IJ}_{\text{Cross}}$, is
115 as follows:
116 \begin{equation}
117 V^{IJ}_{\text{Cross}} =
118        \sum_{i \in I} \sum_{j \in J}
119        \biggl[ V_{\text{LJ}}(r_{ij}) +  V_{\text{dipole}}
120        (\mathbf{r}_{ij},\boldsymbol{\Omega}_{i},\boldsymbol{\Omega}_{j})
121        + V_{\text{sticky}}
122        (\mathbf{r}_{ij},\boldsymbol{\Omega}_{i},\boldsymbol{\Omega}_{j})
123        \biggr]
124 \label{eq:crossPotentail}
125 \end{equation}
126 Where $V_{\text{LJ}}$ is the Lennard Jones potential,
127 $V_{\text{dipole}}$ is the dipole dipole potential, and
128 $V_{\text{sticky}}$ is the sticky potential defined by the SSD model.
206  
207   The bend potential of a molecule is represented by the following function:
208   \begin{equation}
# Line 140 | Line 217 | V_{\text{torsion}}(\phi_{ijkl}) = c_1[1 + \cos \phi]
217   The torsion potential and parameters are also taken from TraPPE. It is
218   of the form:
219   \begin{equation}
220 < V_{\text{torsion}}(\phi_{ijkl}) = c_1[1 + \cos \phi]
220 > V_{\text{torsion}}(\phi) = c_1[1 + \cos \phi]
221          + c_2[1 + \cos(2\phi)]
222          + c_3[1 + \cos(3\phi)]
223   \label{eq:origTorsionPot}
224   \end{equation}
225   Here $\phi_{ijkl}$ is the angle defined by four bonded neighbors $i$,
226 < $j$, $k$, and $l$ (again, see Fig.~\ref{fig:lipidModel}).  However,
227 < for computaional efficency, the torsion potentail has been recast
228 < after the method of CHARMM\cite{charmm1983} whereby the angle series
229 < is converted to a power series of the form:
226 > $j$, $k$, and $l$ (again, see Fig.~\ref{fig:lipidModel}). For
227 > computaional efficency, the torsion potential has been recast after
228 > the method of CHARMM\cite{charmm1983} whereby the angle series is
229 > converted to a power series of the form:
230   \begin{equation}
231   V_{\text{torsion}}(\phi_{ijkl}) =  
232          k_3 \cos^3 \phi + k_2 \cos^2 \phi + k_1 \cos \phi + k_0
# Line 165 | Line 242 | The Lennard-Jones potential is given by:
242   By recasting the equation to a power series, repeated trigonometric
243   evaluations are avoided during the calculation of the potential.
244  
245 < The Lennard-Jones potential is given by:
245 >
246 > The cross portion of the total potential, $V^{IJ}_{\text{Cross}}$, is
247 > as follows:
248   \begin{equation}
249 < V_{\text{LJ}}(r_{ij}) =
250 <        4\epsilon_{ij} \biggl[
251 <        \biggl(\frac{\sigma_{ij}}{r_{ij}}\biggr)^{12}
252 <        - \biggl(\frac{\sigma_{ij}}{r_{ij}}\biggr)^{6}
249 > V^{IJ}_{\text{Cross}} =
250 >        \sum_{i \in I} \sum_{j \in J}
251 >        \biggl[ V_{\text{LJ}}(r_{ij}) +  V_{\text{dipole}}
252 >        (\mathbf{r}_{ij},\boldsymbol{\Omega}_{i},\boldsymbol{\Omega}_{j})
253 >        + V_{\text{sticky}}
254 >        (\mathbf{r}_{ij},\boldsymbol{\Omega}_{i},\boldsymbol{\Omega}_{j})
255          \biggr]
256 < \label{eq:lennardJonesPot}
256 > \label{eq:crossPotentail}
257   \end{equation}
258 < Where $r_ij$ is the distance between atoms $i$ and $j$, $\sigma_{ij}$
259 < scales the length of the interaction, and $\epsilon_{ij}$ scales the
260 < energy of the potential.
258 > Where $V_{\text{LJ}}$ is the Lennard Jones potential,
259 > $V_{\text{dipole}}$ is the dipole dipole potential, and
260 > $V_{\text{sticky}}$ is the sticky potential defined by the SSD
261 > model. Note that not all atom types include all interactions.
262  
263   The dipole-dipole potential has the following form:
264   \begin{equation}
265   V_{\text{dipole}}(\mathbf{r}_{ij},\boldsymbol{\Omega}_{i},
266 <        \boldsymbol{\Omega}_{j}) = \frac{1}{4\pi\epsilon_{0}} \biggl[
267 <        \frac{\boldsymbol{\mu}_{i} \cdot \boldsymbol{\mu}_{j}}{r^{3}_{ij}}
266 >        \boldsymbol{\Omega}_{j}) = \frac{|\mu_i||\mu_j|}{4\pi\epsilon_{0}r_{ij}^{3}} \biggl[
267 >        \boldsymbol{\hat{u}}_{i} \cdot \boldsymbol{\hat{u}}_{j}
268          -
269 <        \frac{3(\boldsymbol{\mu}_i \cdot \mathbf{r}_{ij}) %
270 <                (\boldsymbol{\mu}_j \cdot \mathbf{r}_{ij}) }
271 <                {r^{5}_{ij}} \biggr]
269 >        \frac{3(\boldsymbol{\hat{u}}_i \cdot \mathbf{r}_{ij}) %
270 >                (\boldsymbol{\hat{u}}_j \cdot \mathbf{r}_{ij}) }
271 >                {r^{2}_{ij}} \biggr]
272   \label{eq:dipolePot}
273   \end{equation}
274   Here $\mathbf{r}_{ij}$ is the vector starting at atom $i$ pointing
275   towards $j$, and $\boldsymbol{\Omega}_i$ and $\boldsymbol{\Omega}_j$
276 < are the Euler angles of atom $i$ and $j$
277 < respectively. $\boldsymbol{\mu}_i$ is the dipole vector of atom
278 < $i$ it takes its orientation from $\boldsymbol{\Omega}_i$.
276 > are the Euler angles of atom $i$ and $j$ respectively. $|\mu_i|$ is
277 > the magnitude of the dipole moment of atom $i$ and
278 > $\boldsymbol{\hat{u}}_i$ is the standard unit orientation vector of
279 > $\boldsymbol{\Omega}_i$.
280  
281  
282 < \subsection{\label{sec:SSD}Water Model: SSD and Derivatives}
282 > \subsection{\label{sec:SSD}The {\sc DUFF} Water Models: SSD/E and SSD/RF}
283  
284 < In the interest of computational efficiency, the native solvent used
285 < in {\sc oopse} is the Soft Sticky Dipole (SSD) water model. SSD was
286 < developed by Ichiye \emph{et al.} as a modified form of the
287 < hard-sphere water model proposed by Bratko, Blum, and
284 > In the interest of computational efficiency, the default solvent used
285 > by {\sc oopse} is the extended Soft Sticky Dipole (SSD/E) water
286 > model.\cite{Gezelter04} The original SSD was developed by Ichiye
287 > \emph{et al.}\cite{Ichiye96} as a modified form of the hard-sphere
288 > water model proposed by Bratko, Blum, and
289   Luzar.\cite{Bratko85,Bratko95} It consists of a single point dipole
290   with a Lennard-Jones core and a sticky potential that directs the
291   particles to assume the proper hydrogen bond orientation in the first
292   solvation shell. Thus, the interaction between two SSD water molecules
293   \emph{i} and \emph{j} is given by the potential
294   \begin{equation}
295 < u_{ij} = u_{ij}^{LJ} (r_{ij})\ + u_{ij}^{dp}
296 < (\mathbf{r}_{ij},\boldsymbol{\Omega}_i,\boldsymbol{\Omega}_j)\ +
297 < u_{ij}^{sp}
298 < (\mathbf{r}_{ij},\boldsymbol{\Omega}_i,\boldsymbol{\Omega}_j),
295 > V_{ij} =
296 >        V_{ij}^{LJ} (r_{ij})\ + V_{ij}^{dp}
297 >        (\mathbf{r}_{ij},\boldsymbol{\Omega}_i,\boldsymbol{\Omega}_j)\ +
298 >        V_{ij}^{sp}
299 >        (\mathbf{r}_{ij},\boldsymbol{\Omega}_i,\boldsymbol{\Omega}_j),
300 > \label{eq:ssdPot}
301   \end{equation}
302   where the $\mathbf{r}_{ij}$ is the position vector between molecules
303 < \emph{i} and \emph{j} with magnitude equal to the distance $r_ij$, and
303 > \emph{i} and \emph{j} with magnitude equal to the distance $r_{ij}$, and
304   $\boldsymbol{\Omega}_i$ and $\boldsymbol{\Omega}_j$ represent the
305 < orientations of the respective molecules. The Lennard-Jones, dipole,
306 < and sticky parts of the potential are giving by the following
307 < equations,
305 > orientations of the respective molecules. The Lennard-Jones and dipole
306 > parts of the potential are given by equations \ref{eq:lennardJonesPot}
307 > and \ref{eq:dipolePot} respectively. The sticky part is described by
308 > the following,
309   \begin{equation}
310 < u_{ij}^{LJ}(r_{ij}) = 4\epsilon \left[\left(\frac{\sigma}{r_{ij}}\right)^{12}-\left(\frac{\sigma}{r_{ij}}\right)^{6}\right],
310 > u_{ij}^{sp}(\mathbf{r}_{ij},\boldsymbol{\Omega}_i,\boldsymbol{\Omega}_j)=
311 >        \frac{\nu_0}{2}[s(r_{ij})w(\mathbf{r}_{ij},
312 >        \boldsymbol{\Omega}_i,\boldsymbol{\Omega}_j) +
313 >        s^\prime(r_{ij})w^\prime(\mathbf{r}_{ij},
314 >        \boldsymbol{\Omega}_i,\boldsymbol{\Omega}_j)]\ ,
315 > \label{eq:stickyPot}
316   \end{equation}
317 + where $\nu_0$ is a strength parameter for the sticky potential, and
318 + $s$ and $s^\prime$ are cubic switching functions which turn off the
319 + sticky interaction beyond the first solvation shell. The $w$ function
320 + can be thought of as an attractive potential with tetrahedral
321 + geometry:
322   \begin{equation}
323 < u_{ij}^{dp} = \frac{\boldsymbol{\mu}_i\cdot\boldsymbol{\mu}_j}{r_{ij}^3}-\frac{3(\boldsymbol{\mu}_i\cdot\mathbf{r}_{ij})(\boldsymbol{\mu}_j\cdot\mathbf{r}_{ij})}{r_{ij}^5}\ ,
323 > w({\bf r}_{ij},{\bf \Omega}_i,{\bf \Omega}_j)=
324 >        \sin\theta_{ij}\sin2\theta_{ij}\cos2\phi_{ij},
325 > \label{eq:stickyW}
326   \end{equation}
327 + while the $w^\prime$ function counters the normal aligned and
328 + anti-aligned structures favored by point dipoles:
329   \begin{equation}
330 < \begin{split}
331 < u_{ij}^{sp}
332 < (\mathbf{r}_{ij},\boldsymbol{\Omega}_i,\boldsymbol{\Omega}_j)
232 < &=
233 < \frac{\nu_0}{2}[s(r_{ij})w(\mathbf{r}_{ij},\boldsymbol{\Omega}_i,\boldsymbol{\Omega}_j)\\
234 < & \quad \ +
235 < s^\prime(r_{ij})w^\prime(\mathbf{r}_{ij},\boldsymbol{\Omega}_i,\boldsymbol{\Omega}_j)]\ ,
236 < \end{split}
330 > w^\prime({\bf r}_{ij},{\bf \Omega}_i,{\bf \Omega}_j)=
331 >        (\cos\theta_{ij}-0.6)^2(\cos\theta_{ij}+0.8)^2-w^0,
332 > \label{eq:stickyWprime}
333   \end{equation}
334 < where $\boldsymbol{\mu}_i$ and $\boldsymbol{\mu}_j$ are the dipole
335 < unit vectors of particles \emph{i} and \emph{j} with magnitude 2.35 D,
336 < $\nu_0$ scales the strength of the overall sticky potential, $s$ and
337 < $s^\prime$ are cubic switching functions. The $w$ and $w^\prime$
338 < functions take the following forms,
334 > It should be noted that $w$ is proportional to the sum of the $Y_3^2$
335 > and $Y_3^{-2}$ spherical harmonics (a linear combination which
336 > enhances the tetrahedral geometry for hydrogen bonded structures),
337 > while $w^\prime$ is a purely empirical function.  A more detailed
338 > description of the functional parts and variables in this potential
339 > can be found in the original SSD
340 > articles.\cite{Ichiye96,Ichiye96b,Ichiye99,Ichiye03}
341 >
342 > Since SSD is a single-point {\it dipolar} model, the force
343 > calculations are simplified significantly relative to the standard
344 > {\it charged} multi-point models. In the original Monte Carlo
345 > simulations using this model, Ichiye {\it et al.} reported that using
346 > SSD decreased computer time by a factor of 6-7 compared to other
347 > models.\cite{Ichiye96} What is most impressive is that these savings
348 > did not come at the expense of accurate depiction of the liquid state
349 > properties.  Indeed, SSD maintains reasonable agreement with the Soper
350 > diffraction data for the structural features of liquid
351 > water.\cite{Soper86,Ichiye96} Additionally, the dynamical properties
352 > exhibited by SSD agree with experiment better than those of more
353 > computationally expensive models (like TIP3P and
354 > SPC/E).\cite{Ichiye99} The combination of speed and accurate depiction
355 > of solvent properties makes SSD a very attractive model for the
356 > simulation of large scale biochemical simulations.
357 >
358 > Recent constant pressure simulations revealed issues in the original
359 > SSD model that led to lower than expected densities at all target
360 > pressures.\cite{Ichiye03,Gezelter04} The default model in {\sc oopse}
361 > is therefore SSD/E, a density corrected derivative of SSD that
362 > exhibits improved liquid structure and transport behavior. If the use
363 > of a reaction field long-range interaction correction is desired, it
364 > is recommended that the parameters be modified to those of the SSD/RF
365 > model. Solvent parameters can be easily modified in an accompanying
366 > {\sc BASS} file as illustrated in the scheme below. A table of the
367 > parameter values and the drawbacks and benefits of the different
368 > density corrected SSD models can be found in reference
369 > \ref{Gezelter04}.
370 >
371 > !!!Place a {\sc BASS} scheme showing SSD parameters around here!!!
372 >
373 > \subsection{\label{sec:eam}Embedded Atom Method}
374 >
375 > Several other molecular dynamics packages\cite{dynamo86} exist which have the
376 > capacity to simulate metallic systems, including some that have
377 > parallel computational abilities\cite{plimpton93}. Potentials that
378 > describe bonding transition metal
379 > systems\cite{Finnis84,Ercolessi88,Chen90,Qi99,Ercolessi02} have a
380 > attractive interaction which models  ``Embedding''
381 > a positively charged metal ion in the electron density due to the
382 > free valance ``sea'' of electrons created by the surrounding atoms in
383 > the system. A mostly repulsive pairwise part of the potential
384 > describes the interaction of the positively charged metal core ions
385 > with one another. A particular potential description called the
386 > Embedded Atom Method\cite{Daw84,FBD86,johnson89,Lu97}({\sc eam}) that has
387 > particularly wide adoption has been selected for inclusion in OOPSE. A
388 > good review of {\sc eam} and other metallic potential formulations was done
389 > by Voter.\cite{voter}
390 >
391 > The {\sc eam} potential has the form:
392 > \begin{eqnarray}
393 > V & = & \sum_{i} F_{i}\left[\rho_{i}\right] + \sum_{i} \sum_{j \neq i}
394 > \phi_{ij}({\bf r}_{ij})  \\
395 > \rho_{i}  & = & \sum_{j \neq i} f_{j}({\bf r}_{ij})
396 > \end{eqnarray}S
397 >
398 > where $F_{i} $ is the embedding function that equates the energy required to embed a
399 > positively-charged core ion $i$ into a linear superposition of
400 > sperically averaged atomic electron densities given by
401 > $\rho_{i}$.  $\phi_{ij}$ is a primarily repulsive pairwise interaction
402 > between atoms $i$ and $j$. In the original formulation of
403 > {\sc eam} cite{Daw84}, $\phi_{ij}$ was an entirely repulsive term, however
404 > in later refinements to EAM have shown that non-uniqueness between $F$
405 > and $\phi$ allow for more general forms for $\phi$.\cite{Daw89}
406 > There is a cutoff distance, $r_{cut}$, which limits the
407 > summations in the {\sc eam} equation to the few dozen atoms
408 > surrounding atom $i$ for both the density $\rho$ and pairwise $\phi$
409 > 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}.
410 >
411 >
412 > \subsection{\label{Sec:pbc}Periodic Boundary Conditions}
413 >
414 > \newcommand{\roundme}{\operatorname{round}}
415 >
416 > \textit{Periodic boundary conditions} are widely used to simulate truly
417 > macroscopic systems with a relatively small number of particles. The
418 > simulation box is replicated throughout space to form an infinite
419 > lattice.  During the simulation, when a particle moves in the primary
420 > cell, its image in other boxes move in exactly the same direction with
421 > exactly the same orientation.Thus, as a particle leaves the primary
422 > cell, one of its images will enter through the opposite face.If the
423 > simulation box is large enough to avoid "feeling" the symmetries of
424 > the periodic lattice, surface effects can be ignored. Cubic,
425 > orthorhombic and parallelepiped are the available periodic cells In
426 > OOPSE. We use a matrix to describe the property of the simulation
427 > box. Therefore, both the size and shape of the simulation box can be
428 > changed during the simulation. The transformation from box space
429 > vector $\mathbf{s}$ to its corresponding real space vector
430 > $\mathbf{r}$ is defined by
431   \begin{equation}
432 < w(\mathbf{r}_{ij},\boldsymbol{\Omega}_i,\boldsymbol{\Omega}_j)=\sin\theta_{ij}\sin2\theta_{ij}\cos2\phi_{ij},
432 > \mathbf{r}=\underline{\underline{H}}\cdot\mathbf{s}%
433   \end{equation}
434 +
435 +
436 + where $H=(h_{x},h_{y},h_{z})$ is a transformation matrix made up of
437 + the three box axis vectors. $h_{x},h_{y}$ and $h_{z}$ represent the
438 + three sides of the simulation box respectively.
439 +
440 + To find the minimum image, we convert the real vector to its
441 + corresponding vector in box space first, \bigskip%
442   \begin{equation}
443 < w^\prime(\mathbf{r}_{ij},\boldsymbol{\Omega}_i,\boldsymbol{\Omega}_j) = (\cos\theta_{ij}-0.6)^2(\cos\theta_{ij}+0.8)^2-w^0,
443 > \mathbf{s}=\underline{\underline{H}}^{-1}\cdot\mathbf{r}%
444   \end{equation}
445 < where $w^0 = 0.07715$. The $w$ function is the tetrahedral attractive
446 < term that promotes hydrogen bonding orientations within the first
447 < solvation shell, and $w^\prime$ is a dipolar repulsion term that
448 < repels unrealistic dipolar arrangements within the first solvation
449 < shell. A more detailed description of the functional parts and
254 < variables in this potential can be found in other
255 < articles.\cite{liu96:new_model,chandra99:ssd_md}
445 > And then, each element of $\mathbf{s}$ is wrapped to lie between -0.5 to 0.5,
446 > \begin{equation}
447 > s_{i}^{\prime}=s_{i}-\roundme(s_{i})
448 > \end{equation}
449 > where
450  
451 < Since SSD is a one-site point dipole model, the force calculations are
258 < simplified significantly from a computational standpoint, in that the
259 < number of long-range interactions is dramatically reduced. In the
260 < original Monte Carlo simulations using this model, Ichiye \emph{et
261 < al.} reported a calculation speed up of up to an order of magnitude
262 < over other comparable models while maintaining the structural behavior
263 < of water.\cite{liu96:new_model} In the original molecular dynamics studies of
264 < SSD, it was shown that it actually improves upon the prediction of
265 < water's dynamical properties over TIP3P and SPC/E.\cite{chandra99:ssd_md}
451 > %
452  
453 < Recent constant pressure simulations revealed issues in the original
454 < SSD model that led to lower than expected densities at all target
455 < pressures.\cite{Ichiye03,Gezelter04} Reparameterizations of the
456 < original SSD have resulted in improved density behavior, as well as
457 < alterations in the water structure and transport behavior. {\sc oopse} is
458 < easily modified to impliment these new potential parameter sets for
459 < the derivative water models: SSD1, SSD/E, and SSD/RF. All of the
460 < variable parameters are listed in the accompanying BASS file, and
461 < these parameters simply need to be changed to the updated values.
453 > \begin{equation}
454 > \roundme(x)=\left\{
455 > \begin{array}{cc}
456 > \lfloor{x+0.5}\rfloor & \text{if \ }x\geqslant0\\
457 > \lceil{x-0.5}\rceil & \text{otherwise}%
458 > \end{array}
459 > \right.
460 > \end{equation}
461 > For example, $\roundme(3.6)=4$, $\roundme(3.1)=3$, $\roundme(-3.6)=-4$,
462 > $\roundme(-3.1)=-3$.
463  
464 + Finally, we obtain the minimum image coordinates by transforming back
465 + to real space,%
466  
467 < \subsection{\label{sec:eam}Embedded Atom Model}
467 > \begin{equation}
468 > \mathbf{r}^{\prime}=\underline{\underline{H}}^{-1}\cdot\mathbf{s}^{\prime}%
469 > \end{equation}
470  
280 here there be Monsters

Diff Legend

Removed lines
+ Added lines
< Changed lines
> Changed lines