1 |
|
2 |
\section{\label{sec:mechanics}Mechanics} |
3 |
|
4 |
\subsection{\label{integrate}Integrating the Equations of Motion: the |
5 |
DLM method} |
6 |
|
7 |
The default method for integrating the equations of motion in {\sc |
8 |
oopse} is a velocity-Verlet version of the symplectic splitting method |
9 |
proposed by Dullweber, Leimkuhler and McLachlan |
10 |
(DLM).\cite{Dullweber1997} When there are no directional atoms or |
11 |
rigid bodies present in the simulation, this integrator becomes the |
12 |
standard velocity-Verlet integrator which is known to sample the |
13 |
microcanonical (NVE) ensemble.\cite{} |
14 |
|
15 |
Previous integration methods for orientational motion have problems |
16 |
that are avoided in the DLM method. Direct propagation of the Euler |
17 |
angles has a known $1/\sin\theta$ divergence in the equations of |
18 |
motion for $\phi$ and $\psi$,\cite{AllenTildesley} leading to |
19 |
numerical instabilities any time one of the directional atoms or rigid |
20 |
bodies has an orientation near $\theta=0$ or $\theta=\pi$. More |
21 |
modern quaternion-based integration methods have relatively poor |
22 |
energy conservation. While quaternions work well for orientational |
23 |
motion in other ensembles, the microcanonical ensemble has a |
24 |
constant energy requirement that is quite sensitive to errors in the |
25 |
equations of motion. An earlier implementation of {\sc oopse} |
26 |
utilized quaternions for propagation of rotational motion; however, a |
27 |
detailed investigation showed that they resulted in a steady drift in |
28 |
the total energy, something that has been observed by |
29 |
others.\cite{Laird97} |
30 |
|
31 |
The key difference in the integration method proposed by Dullweber |
32 |
\emph{et al.} is that the entire rotation matrix is propagated from |
33 |
one time step to the next. In the past, this would not have been |
34 |
feasible, since that the rotation matrix for a single body has nine |
35 |
elements compared to 3 or 4 elements for Euler angles and quaternions |
36 |
respectively. System memory has become less costly in recent times, |
37 |
and this can be translated into substantial benefits in energy |
38 |
conservation. |
39 |
|
40 |
The basic equations of motion being integrated are derived from the |
41 |
Hamiltonian for conservative systems containing rigid bodies, |
42 |
\begin{equation} |
43 |
H = \sum_{i} \left( \frac{1}{2} m_i {\bf v}_i^T \cdot {\bf v}_i + |
44 |
\frac{1}{2} {\bf j}_i^T \cdot \overleftrightarrow{\mathsf{I}}_i^{-1} \cdot |
45 |
{\bf j}_i \right) + |
46 |
V\left(\left\{{\bf r}\right\}, \left\{\mathsf{A}\right\}\right) |
47 |
\end{equation} |
48 |
where ${\bf r}_i$ and ${\bf v}_i$ are the cartesian position vector |
49 |
and velocity of the center of mass of particle $i$, and ${\bf j}_i$ |
50 |
and $\overleftrightarrow{\mathsf{I}}_i$ are the body-fixed angular |
51 |
momentum and moment of inertia tensor, respectively. $\mathsf{A}_i$ |
52 |
is the $3 \times 3$ rotation matrix describing the instantaneous |
53 |
orientation of the particle. $V$ is the potential energy function |
54 |
which may depend on both the positions $\left\{{\bf r}\right\}$ and |
55 |
orientations $\left\{\mathsf{A}\right\}$ of all particles. The |
56 |
equations of motion for the particle centers of mass are derived from |
57 |
Hamilton's equations and are quite simple, |
58 |
\begin{eqnarray} |
59 |
\dot{{\bf r}} & = & {\bf v} \\ |
60 |
\dot{{\bf v}} & = & \frac{{\bf f}}{m} |
61 |
\end{eqnarray} |
62 |
where ${\bf f}$ is the instantaneous force on the center of mass |
63 |
of the particle, |
64 |
\begin{equation} |
65 |
{\bf f} = - \frac{\partial}{\partial |
66 |
{\bf r}} V(\left\{{\bf r}(t)\right\}, \left\{\mathsf{A}(t)\right\}). |
67 |
\end{equation} |
68 |
|
69 |
The equations of motion for the orientational degrees of freedom are |
70 |
\begin{eqnarray} |
71 |
\dot{\mathsf{A}} & = & \mathsf{A} \cdot |
72 |
\mbox{ skew}\left(\overleftrightarrow{I}^{-1} \cdot {\bf j}\right) \\ |
73 |
\dot{{\bf j}} & = & {\bf j} \times \left( \overleftrightarrow{I}^{-1} |
74 |
\cdot {\bf j} \right) - \mbox{ rot}\left(\mathsf{A}^{T} \cdot \frac{\partial |
75 |
V}{\partial \mathsf{A}} \right) |
76 |
\end{eqnarray} |
77 |
In these equations of motion, the $\mbox{skew}$ matrix of a vector |
78 |
${\bf v} = \left( v_1, v_2, v_3 \right)$ is defined: |
79 |
\begin{equation} |
80 |
\mbox{skew}\left( {\bf v} \right) := \left( |
81 |
\begin{array}{ccc} |
82 |
0 & v_3 & - v_2 \\ |
83 |
-v_3 & 0 & v_1 \\ |
84 |
v_2 & -v_1 & 0 |
85 |
\end{array} |
86 |
\right) |
87 |
\end{equation} |
88 |
The $\mbox{rot}$ notation refers to the mapping of the $3 \times 3$ |
89 |
rotation matrix to a vector of orientations by first computing the |
90 |
skew-symmetric part $\left(\mathsf{A} - \mathsf{A}^{T}\right)$ and |
91 |
then associating this with a length 3 vector by inverting the |
92 |
$\mbox{skew}$ function above: |
93 |
\begin{equation} |
94 |
\mbox{rot}\left(\mathsf{A}\right) := \mbox{ skew}^{-1}\left(\mathsf{A} |
95 |
- \mathsf{A}^{T} \right) |
96 |
\end{equation} |
97 |
Written this way, the $\mbox{rot}$ operation creates a set of |
98 |
conjugate angle coordinates to the body-fixed angular momenta |
99 |
represented by ${\bf j}$. This equation of motion for angular momenta |
100 |
is equivalent to the more familiar body-fixed forms, |
101 |
\begin{eqnarray} |
102 |
\dot{j_{x}} & = & \tau^b_x(t) + |
103 |
\left(\overleftrightarrow{\mathsf{I}}_{yy} - \overleftrightarrow{\mathsf{I}}_{zz} \right) j_y j_z \\ |
104 |
\dot{j_{y}} & = & \tau^b_y(t) + |
105 |
\left(\overleftrightarrow{\mathsf{I}}_{zz} - \overleftrightarrow{\mathsf{I}}_{xx} \right) j_z j_x \\ |
106 |
\dot{j_{z}} & = & \tau^b_z(t) + |
107 |
\left(\overleftrightarrow{\mathsf{I}}_{xx} - \overleftrightarrow{\mathsf{I}}_{yy} \right) j_x j_y |
108 |
\end{eqnarray} |
109 |
which utilize the body-fixed torques, ${\bf \tau}^b$. Torques are |
110 |
most easily derived in the space-fixed frame, |
111 |
\begin{equation} |
112 |
{\bf \tau}^b(t) = \mathsf{A}(t) \cdot {\bf \tau}^s(t) |
113 |
\end{equation} |
114 |
where the torques are either derived from the forces on the |
115 |
constituent atoms of the rigid body, or for directional atoms, |
116 |
directly from derivatives of the potential energy, |
117 |
\begin{equation} |
118 |
{\bf \tau}^s(t) = - \hat{\bf u}(t) \times \left( \frac{\partial} |
119 |
{\partial \hat{\bf u}} V\left(\left\{ {\bf r}(t) \right\}, \left\{ |
120 |
\mathsf{A}(t) \right\}\right) \right). |
121 |
\end{equation} |
122 |
Here $\hat{\bf u}$ is a unit vector pointing along the principal axis |
123 |
of the particle in the space-fixed frame. |
124 |
|
125 |
The DLM method uses a Trotter factorization of the orientational |
126 |
propagator. This has three effects: |
127 |
\begin{enumerate} |
128 |
\item the integrator is area-preserving in phase space (i.e. it is |
129 |
{\it symplectic}), |
130 |
\item the integrator is time-{\it reversible}, making it suitable for Hybrid |
131 |
Monte Carlo applications, and |
132 |
\item the error for a single time step is of order $O\left(h^3\right)$ |
133 |
for timesteps of length $h$. |
134 |
\end{enumerate} |
135 |
|
136 |
The integration of the equations of motion is carried out in a |
137 |
velocity-Verlet style 2 part algorithm: |
138 |
|
139 |
{\tt moveA:} |
140 |
\begin{eqnarray} |
141 |
{\bf v}\left(t + \delta t / 2\right) & \leftarrow & {\bf |
142 |
v}(t) + \frac{\delta t}{2} \left( {\bf f}(t) / m \right) \\ |
143 |
{\bf r}(t + \delta t) & \leftarrow & {\bf r}(t) + \delta t {\bf |
144 |
v}\left(t + \delta t / 2 \right) \\ |
145 |
{\bf j}\left(t + \delta t / 2 \right) & \leftarrow & {\bf |
146 |
j}(t) + \frac{\delta t}{2} {\bf \tau}^b(t) \\ |
147 |
\mathsf{A}(t + \delta t) & \leftarrow & \mathrm{rot}\left( \delta t |
148 |
{\bf j}(t + \delta t / 2) \cdot \overleftrightarrow{\mathsf{I}}^{-1} |
149 |
\right) |
150 |
\end{eqnarray} |
151 |
|
152 |
In this context, the $\mathrm{rot}$ function is the reversible product |
153 |
of the three body-fixed rotations, |
154 |
\begin{equation} |
155 |
\mathrm{rot}({\bf a}) = \mathsf{G}_x(a_x / 2) \cdot |
156 |
\mathsf{G}_y(a_y / 2) \cdot \mathsf{G}_z(a_z) \cdot \mathsf{G}_y(a_y / |
157 |
2) \cdot \mathsf{G}_x(a_x /2) |
158 |
\end{equation} |
159 |
where each rotational propagator, $\mathsf{G}_\alpha(\theta)$, rotates |
160 |
both the rotation matrix ($\mathsf{A}$) and the body-fixed angular |
161 |
momentum (${\bf j}$) by an angle $\theta$ around body-fixed axis |
162 |
$\alpha$, |
163 |
\begin{equation} |
164 |
\mathsf{G}_\alpha( \theta ) = \left\{ |
165 |
\begin{array}{lcl} |
166 |
\mathsf{A}(t) & \leftarrow & \mathsf{A}(0) \cdot \mathsf{R}_\alpha(\theta)^T \\ |
167 |
{\bf j}(t) & \leftarrow & \mathsf{R}_\alpha(\theta) \cdot {\bf j}(0) |
168 |
\end{array} |
169 |
\right. |
170 |
\end{equation} |
171 |
$\mathsf{R}_\alpha$ is a quadratic approximation to |
172 |
the single-axis rotation matrix. For example, in the small-angle |
173 |
limit, the rotation matrix around the body-fixed x-axis can be |
174 |
approximated as |
175 |
\begin{equation} |
176 |
\mathsf{R}_x(\theta) = \left( |
177 |
\begin{array}{ccc} |
178 |
1 & 0 & 0 \\ |
179 |
0 & \frac{1-\theta^2 / 4}{1 + \theta^2 / 4} & -\frac{\theta}{1+ |
180 |
\theta^2 / 4} \\ |
181 |
0 & \frac{\theta}{1+ |
182 |
\theta^2 / 4} & \frac{1-\theta^2 / 4}{1 + \theta^2 / 4} |
183 |
\end{array} |
184 |
\right). |
185 |
\end{equation} |
186 |
All other rotations follow in a straightforward manner. |
187 |
|
188 |
After the first part of the propagation, the forces and body-fixed |
189 |
torques are calculated at the new positions and orientations |
190 |
|
191 |
{\tt doForces:} |
192 |
\begin{eqnarray} |
193 |
{\bf f}(t + \delta t) & \leftarrow & - \frac{\partial V}{\partial {\bf |
194 |
r}(t + \delta t)} \\ |
195 |
{\bf \tau}^{s}(t + \delta t) & \leftarrow & {\bf u}(t + \delta t) |
196 |
\times \frac{\partial V}{\partial {\bf u}} \\ |
197 |
{\bf \tau}^{b}(t + \delta t) & \leftarrow & \mathsf{A}(t + \delta t) |
198 |
\cdot {\bf \tau}^s(t + \delta t) |
199 |
\end{eqnarray} |
200 |
|
201 |
{\sc oopse} automatically updates ${\bf u}$ when the rotation matrix |
202 |
$\mathsf{A}$ is calculated in {\tt moveA}. Once the forces and |
203 |
torques have been obtained at the new time step, the velocities can be |
204 |
advanced to the same time value. |
205 |
|
206 |
{\tt moveB:} |
207 |
\begin{eqnarray} |
208 |
{\bf v}\left(t + \delta t \right) & \leftarrow & {\bf |
209 |
v}\left(t + \delta t / 2 \right) + \frac{\delta t}{2} \left( |
210 |
{\bf f}(t + \delta t) / m \right) \\ |
211 |
{\bf j}\left(t + \delta t \right) & \leftarrow & {\bf |
212 |
j}\left(t + \delta t / 2 \right) + \frac{\delta t}{2} {\bf |
213 |
\tau}^b(t + \delta t) |
214 |
\end{eqnarray} |
215 |
|
216 |
The matrix rotations used in the DLM method end up being more costly |
217 |
computationally than the simpler arithmetic quaternion |
218 |
propagation. With the same time step, a 1000-molecule water simulation |
219 |
shows an average 7\% increase in computation time using the DLM method |
220 |
in place of quaternions. This cost is more than justified when |
221 |
comparing the energy conservation of the two methods as illustrated in |
222 |
figure \ref{timestep}. |
223 |
|
224 |
\begin{figure} |
225 |
\epsfxsize=6in |
226 |
\epsfbox{timeStep.epsi} |
227 |
\caption{Energy conservation using quaternion based integration versus |
228 |
the method proposed by Dullweber \emph{et al.} with increasing time |
229 |
step. For each time step, the dotted line is total energy using the |
230 |
symplectic step integrator, and the solid line comes from the |
231 |
quaternion integrator. The larger time step plots are shifted up from |
232 |
the true energy baseline for clarity.} |
233 |
\label{timestep} |
234 |
\end{figure} |
235 |
|
236 |
In figure \ref{timestep}, the resulting energy drift at various time |
237 |
steps for both the DLM and quaternion integration schemes is |
238 |
compared. All of the 1000 molecule water simulations started with the |
239 |
same configuration, and the only difference was the method for |
240 |
handling rotational motion. At time steps of 0.1 and 0.5 fs, both |
241 |
methods for propagating molecule rotation conserve energy fairly well, |
242 |
with the quaternion method showing a slight energy drift over time in |
243 |
the 0.5 fs time step simulation. At time steps of 1 and 2 fs, the |
244 |
energy conservation benefits of the DLM method are clearly |
245 |
demonstrated. Thus, while maintaining the same degree of energy |
246 |
conservation, one can take considerably longer time steps, leading to |
247 |
an overall reduction in computation time. |
248 |
|
249 |
|
250 |
There is only one specific keyword relevant to the default integrator, |
251 |
and that is the time step for integrating the equations of motion. |
252 |
|
253 |
\begin{tabular}{llll} |
254 |
{\bf variable} & {\bf {\tt .bass} keyword} & {\bf units} & {\bf |
255 |
default value} \\ |
256 |
$\delta t$ & {\tt dt = 2.0;} & fs & none |
257 |
\end{tabular} |
258 |
|
259 |
\subsection{\label{sec:extended}Extended Systems for other Ensembles} |
260 |
|
261 |
{\sc oopse} implements a number of extended system integrators for |
262 |
sampling from other ensembles relevant to chemical physics. The |
263 |
integrator can selected with the {\tt ensemble} keyword in the |
264 |
{\tt .bass} file: |
265 |
|
266 |
\begin{center} |
267 |
\begin{tabular}{lll} |
268 |
{\bf Integrator} & {\bf Ensemble} & {\bf {\tt .bass} line} \\ |
269 |
NVE & microcanonical & {\tt ensemble = ``NVE''; } \\ |
270 |
NVT & canonical & {\tt ensemble = ``NVT''; } \\ |
271 |
NPTi & isobaric-isothermal (with isotropic volume changes) & {\tt |
272 |
ensemble = ``NPTi'';} \\ |
273 |
NPTf & isobaric-isothermal (with changes to box shape) & {\tt |
274 |
ensemble = ``NPTf'';} \\ |
275 |
NPTxyz & approximate isobaric-isothermal & {\tt ensemble = |
276 |
``NPTxyz'';} \\ |
277 |
& (with separate barostats on each box dimension) & |
278 |
\end{tabular} |
279 |
\end{center} |
280 |
|
281 |
The relatively well-known Nos\'e-Hoover thermostat is implemented in |
282 |
{\sc oopse}'s NVT integrator. This method couples an extra degree of |
283 |
freedom (the thermostat) to the kinetic energy of the system, and has |
284 |
been shown to sample the canonical distribution in the system degrees |
285 |
of freedom while conserving a quantity that is, to within a constant, |
286 |
the Helmholtz free energy. |
287 |
|
288 |
NPT algorithms attempt to maintain constant pressure in the system by |
289 |
coupling the volume of the system to a barostat. {\sc oopse} contains |
290 |
three different constant pressure algorithms. The first two, NPTi and |
291 |
NPTf have been shown to conserve a quantity that is, to within a |
292 |
constant, the Gibbs free energy. The Melchionna modification to the |
293 |
Hoover barostat is implemented in both NPTi and NPTf. NPTi allows |
294 |
only isotropic changes in the simulation box, while box {\it shape} |
295 |
variations are allowed in NPTf. The NPTxyz integrator has {\it not} |
296 |
been shown to sample from the isobaric-isothermal ensemble. It is |
297 |
useful, however, in that it maintains orthogonality for the axes of |
298 |
the simulation box while attempting to equalize pressure along the |
299 |
three perpendicular directions in the box. |
300 |
|
301 |
Each of the extended system integrators requires additional keywords |
302 |
to set target values for the thermodynamic state variables that are |
303 |
being held constant. Keywords are also required to set the |
304 |
characteristic decay times for the dynamics of the extended |
305 |
variables. |
306 |
|
307 |
\begin{tabular}{llll} |
308 |
{\bf variable} & {\bf {\tt .bass} keyword} & {\bf units} & {\bf |
309 |
default value} \\ |
310 |
$T_{\mathrm{target}}$ & {\tt targetTemperature = 300;} & K & none \\ |
311 |
$P_{\mathrm{target}}$ & {\tt targetPressure = 1;} & atm & none \\ |
312 |
$\tau_T$ & {\tt tauThermostat = 1e3;} & fs & none \\ |
313 |
$\tau_B$ & {\tt tauBarostat = 5e3;} & fs & none \\ |
314 |
& {\tt resetTime = 200;} & fs & none \\ |
315 |
& {\tt useInitialExtendedSystemState = ``true'';} & logical & |
316 |
false |
317 |
\end{tabular} |
318 |
|
319 |
Two additional keywords can be used to either clear the extended |
320 |
system variables periodically ({\tt resetTime}), or to maintain the |
321 |
state of the extended system variables between simulations ({\tt |
322 |
useInitialExtendedSystemState}). More details on these variables |
323 |
and their use in the integrators follows below. |
324 |
|
325 |
\subsubsection{\label{sec:noseHooverThermo}Nos\'{e}-Hoover Thermostatting} |
326 |
|
327 |
The Nos\'e-Hoover equations of motion are given by\cite{Hoover85} |
328 |
\begin{eqnarray} |
329 |
\dot{{\bf r}} & = & {\bf v} \\ |
330 |
\dot{{\bf v}} & = & \frac{{\bf f}}{m} - \chi {\bf v} \\ |
331 |
\dot{\mathsf{A}} & = & \mathsf{A} \cdot |
332 |
\mbox{ skew}\left(\overleftrightarrow{I}^{-1} \cdot {\bf j}\right) \\ |
333 |
\dot{{\bf j}} & = & {\bf j} \times \left( \overleftrightarrow{I}^{-1} |
334 |
\cdot {\bf j} \right) - \mbox{ rot}\left(\mathsf{A}^{T} \cdot \frac{\partial |
335 |
V}{\partial \mathsf{A}} \right) - \chi {\bf j} |
336 |
\label{eq:nosehoovereom} |
337 |
\end{eqnarray} |
338 |
|
339 |
$\chi$ is an ``extra'' variable included in the extended system, and |
340 |
it is propagated using the first order equation of motion |
341 |
\begin{equation} |
342 |
\dot{\chi} = \frac{1}{\tau_{T}^2} \left( \frac{T}{T_{\mathrm{target}}} - 1 \right). |
343 |
\label{eq:nosehooverext} |
344 |
\end{equation} |
345 |
|
346 |
The instantaneous temperature $T$ is proportional to the total kinetic |
347 |
energy (both translational and orientational) and is given by |
348 |
\begin{equation} |
349 |
T = \frac{2 K}{f k_B} |
350 |
\end{equation} |
351 |
Here, $f$ is the total number of degrees of freedom in the system, |
352 |
\begin{equation} |
353 |
f = 3 N + 3 N_{\mathrm{orient}} - N_{\mathrm{constraints}} |
354 |
\end{equation} |
355 |
and $K$ is the total kinetic energy, |
356 |
\begin{equation} |
357 |
K = \sum_{i=1}^{N} \frac{1}{2} m_i {\bf v}_i^T \cdot {\bf v}_i + |
358 |
\sum_{i=1}^{N_{\mathrm{orient}}} \sum_{\alpha=x,y,z} \frac{{\bf |
359 |
j}_{i\alpha}^T \cdot {\bf j}_{i\alpha}}{2 |
360 |
\overleftrightarrow{\mathsf{I}}_{i,\alpha \alpha}} |
361 |
\end{equation} |
362 |
|
363 |
In eq.(\ref{eq:nosehooverext}), $\tau_T$ is the time constant for |
364 |
relaxation of the temperature to the target value. To set values for |
365 |
$\tau_T$ or $T_{\mathrm{target}}$ in a simulation, one would use the |
366 |
{\tt tauThermostat} and {\tt targetTemperature} keywords in the {\tt |
367 |
.bass} file. The units for {\tt tauThermostat} are fs, and the units |
368 |
for the {\tt targetTemperature} are degrees K. The integration of |
369 |
the equations of motion is carried out in a velocity-Verlet style 2 |
370 |
part algorithm: |
371 |
|
372 |
{\tt moveA:} |
373 |
\begin{eqnarray} |
374 |
T(t) & \leftarrow & \left\{{\bf v}(t)\right\}, \left\{{\bf j}(t)\right\} \\ |
375 |
{\bf v}\left(t + \delta t / 2\right) & \leftarrow & {\bf |
376 |
v}(t) + \frac{\delta t}{2} \left( \frac{{\bf f}(t)}{m} - {\bf v}(t) |
377 |
\chi(t)\right) \\ |
378 |
{\bf r}(t + \delta t) & \leftarrow & {\bf r}(t) + \delta t {\bf |
379 |
v}\left(t + \delta t / 2 \right) \\ |
380 |
{\bf j}\left(t + \delta t / 2 \right) & \leftarrow & {\bf |
381 |
j}(t) + \frac{\delta t}{2} \left( {\bf \tau}^b(t) - {\bf j}(t) |
382 |
\chi(t) \right) \\ |
383 |
\mathsf{A}(t + \delta t) & \leftarrow & \mathrm{rot}\left(\delta t * |
384 |
{\bf j}(t + \delta t / 2) \overleftrightarrow{\mathsf{I}}^{b} \right) \\ |
385 |
\chi\left(t + \delta t / 2 \right) & \leftarrow & \chi(t) + |
386 |
\frac{\delta t}{2 \tau_T^2} \left( \frac{T(t)}{T_{\mathrm{target}}} - 1 |
387 |
\right) |
388 |
\end{eqnarray} |
389 |
|
390 |
Here $\mathrm{rot}( {\bf j} / {\bf I} )$ is the same symplectic |
391 |
Trotter factorization of the three rotation operations that was |
392 |
discussed in the section on the DLM integrator. Note that this |
393 |
operation modifies both the rotation matrix $\mathsf{A}$ and the |
394 |
angular momentum ${\bf j}$. {\tt moveA} propagates velocities by a |
395 |
half time step, and positional degrees of freedom by a full time step. |
396 |
The new positions (and orientations) are then used to calculate a new |
397 |
set of forces and torques. |
398 |
|
399 |
{\tt doForces:} |
400 |
\begin{eqnarray} |
401 |
{\bf f}(t + \delta t) & \leftarrow & - \frac{\partial V}{\partial {\bf |
402 |
r}(t + \delta t)} \\ |
403 |
{\bf \tau}^{s}(t + \delta t) & \leftarrow & {\bf u}(t + \delta t) |
404 |
\times \frac{\partial V}{\partial {\bf u}} \\ |
405 |
{\bf \tau}^{b}(t + \delta t) & \leftarrow & \mathsf{A}(t + \delta t) |
406 |
\cdot {\bf \tau}^s(t + \delta t) |
407 |
\end{eqnarray} |
408 |
|
409 |
Here ${\bf u}$ is a unit vector pointing along the principal axis of |
410 |
the rigid body being propagated, ${\bf \tau}^s$ is the torque in the |
411 |
space-fixed (laboratory) frame, and ${\bf \tau}^b$ is the torque in |
412 |
the body-fixed frame. ${\bf u}$ is automatically calculated when the |
413 |
rotation matrix $\mathsf{A}$ is calculated in {\tt moveA}. |
414 |
|
415 |
Once the forces and torques have been obtained at the new time step, |
416 |
the velocities can be advanced to the same time value. |
417 |
|
418 |
{\tt moveB:} |
419 |
\begin{eqnarray} |
420 |
T(t + \delta t) & \leftarrow & \left\{{\bf v}(t + \delta t)\right\}, |
421 |
\left\{{\bf j}(t + \delta t)\right\} \\ |
422 |
\chi\left(t + \delta t \right) & \leftarrow & \chi\left(t + \delta t / |
423 |
2 \right) + \frac{\delta t}{2 \tau_T^2} \left( \frac{T(t+\delta |
424 |
t)}{T_{\mathrm{target}}} - 1 \right) \\ |
425 |
{\bf v}\left(t + \delta t \right) & \leftarrow & {\bf |
426 |
v}\left(t + \delta t / 2 \right) + \frac{\delta t}{2} \left( |
427 |
\frac{{\bf f}(t + \delta t)}{m} - {\bf v}(t + \delta t) |
428 |
\chi(t \delta t)\right) \\ |
429 |
{\bf j}\left(t + \delta t \right) & \leftarrow & {\bf |
430 |
j}\left(t + \delta t / 2 \right) + \frac{\delta t}{2} \left( {\bf |
431 |
\tau}^b(t + \delta t) - {\bf j}(t + \delta t) |
432 |
\chi(t + \delta t) \right) |
433 |
\end{eqnarray} |
434 |
|
435 |
Since ${\bf v}(t + \delta t)$ and ${\bf j}(t + \delta t)$ are required |
436 |
to caclculate $T(t + \delta t)$ as well as $\chi(t + \delta t)$, they |
437 |
indirectly depend on their own values at time $t + \delta t$. {\tt |
438 |
moveB} is therefore done in an iterative fashion until $\chi(t + |
439 |
\delta t)$ becomes self-consistent. The relative tolerance for the |
440 |
self-consistency check defaults to a value of $\mbox{10}^{-6}$, but |
441 |
{\sc oopse} will terminate the iteration after 4 loops even if the |
442 |
consistency check has not been satisfied. |
443 |
|
444 |
The Nos\'e-Hoover algorithm is known to conserve a Hamiltonian for the |
445 |
extended system that is, to within a constant, identical to the |
446 |
Helmholtz free energy, |
447 |
\begin{equation} |
448 |
H_{\mathrm{NVT}} = V + K + f k_B T_{\mathrm{target}} \left( |
449 |
\frac{\tau_{T}^2 \chi^2(t)}{2} + \int_{0}^{t} \chi(t\prime) dt\prime |
450 |
\right) |
451 |
\end{equation} |
452 |
Poor choices of $\delta t$ or $\tau_T$ can result in non-conservation |
453 |
of $H_{\mathrm{NVT}}$, so the conserved quantity is maintained in the |
454 |
last column of the {\tt .stat} file to allow checks on the quality of |
455 |
the integration. |
456 |
|
457 |
Bond constraints are applied at the end of both the {\tt moveA} and |
458 |
{\tt moveB} portions of the algorithm. Details on the constraint |
459 |
algorithms are given in section \ref{sec:rattle}. |
460 |
|
461 |
\subsubsection{\label{sec:NPTi}Constant-pressure integration (isotropic box)} |
462 |
|
463 |
To carry out isobaric-isothermal ensemble calculations {\sc oopse} |
464 |
implements the Melchionna modifications to the Nos\'e-Hoover-Andersen |
465 |
equations of motion,\cite{Melchionna93} |
466 |
|
467 |
\begin{eqnarray} |
468 |
\dot{{\bf r}} & = & {\bf v} + \eta \left( {\bf r} - {\bf R}_0 \right) \\ |
469 |
\dot{{\bf v}} & = & \frac{{\bf f}}{m} - (\eta + \chi) {\bf v} \\ |
470 |
\dot{\mathsf{A}} & = & \mathsf{A} \cdot |
471 |
\mbox{ skew}\left(\overleftrightarrow{I}^{-1} \cdot {\bf j}\right) \\ |
472 |
\dot{{\bf j}} & = & {\bf j} \times \left( \overleftrightarrow{I}^{-1} |
473 |
\cdot {\bf j} \right) - \mbox{ rot}\left(\mathsf{A}^{T} \cdot \frac{\partial |
474 |
V}{\partial \mathsf{A}} \right) - \chi {\bf j} \\ |
475 |
\dot{\chi} & = & \frac{1}{\tau_{T}^2} \left( |
476 |
\frac{T}{T_{\mathrm{target}}} - 1 \right) \\ |
477 |
\dot{\eta} & = & \frac{1}{\tau_{B}^2 f k_B T_{\mathrm{target}}} V \left( P - |
478 |
P_{\mathrm{target}} \right) \\ |
479 |
\dot{V} & = & 3 V \eta |
480 |
\label{eq:melchionna1} |
481 |
\end{eqnarray} |
482 |
|
483 |
$\chi$ and $\eta$ are the ``extra'' degrees of freedom in the extended |
484 |
system. $\chi$ is a thermostat, and it has the same function as it |
485 |
does in the Nos\'e-Hoover NVT integrator. $\eta$ is a barostat which |
486 |
controls changes to the volume of the simulation box. ${\bf R}_0$ is |
487 |
the location of the center of mass for the entire system. |
488 |
|
489 |
The NPTi integrator requires an instantaneous pressure. This quantity |
490 |
is calculated via the pressure tensor, |
491 |
\begin{equation} |
492 |
\overleftrightarrow{\mathsf{P}}(t) = \frac{1}{V(t)} \left( \sum_{i=1}^{N} |
493 |
m_i {\bf v}_i(t) \otimes {\bf v}_i(t) \right) + |
494 |
\overleftrightarrow{\mathsf{W}}(t) |
495 |
\end{equation} |
496 |
The kinetic contribution to the pressure tensor utilizes the {\it |
497 |
outer} product of the velocities denoted by the $\otimes$ symbol. The |
498 |
stress tensor is calculated from another outer product of the |
499 |
inter-atomic separation vectors (${\bf r}_{ij} = {\bf r}_j - {\bf |
500 |
r}_i$) with the forces between the same two atoms, |
501 |
\begin{equation} |
502 |
\overleftrightarrow{\mathsf{W}}(t) = \sum_{i} \sum_{j>i} {\bf r}_{ij}(t) |
503 |
\otimes {\bf f}_{ij}(t) |
504 |
\end{equation} |
505 |
The instantaneous pressure is then simply obtained from the trace of |
506 |
the Pressure tensor, |
507 |
\begin{equation} |
508 |
P(t) = \frac{1}{3} \mathrm{Tr} \left( \overleftrightarrow{\mathsf{P}}(t) |
509 |
\right) |
510 |
\end{equation} |
511 |
|
512 |
In eq.(\ref{eq:melchionna1}), $\tau_B$ is the time constant for |
513 |
relaxation of the pressure to the target value. To set values for |
514 |
$\tau_B$ or $P_{\mathrm{target}}$ in a simulation, one would use the |
515 |
{\tt tauBarostat} and {\tt targetPressure} keywords in the {\tt .bass} |
516 |
file. The units for {\tt tauBarostat} are fs, and the units for the |
517 |
{\tt targetPressure} are atmospheres. Like in the NVT integrator, the |
518 |
integration of the equations of motion is carried out in a |
519 |
velocity-Verlet style 2 part algorithm: |
520 |
|
521 |
{\tt moveA:} |
522 |
\begin{eqnarray} |
523 |
T(t) & \leftarrow & \left\{{\bf v}(t)\right\}, \left\{{\bf j}(t)\right\} \\ |
524 |
P(t) & \leftarrow & \left\{{\bf r}(t)\right\}, \left\{{\bf v}(t)\right\}, \left\{{\bf f}(t)\right\} \\ |
525 |
{\bf v}\left(t + \delta t / 2\right) & \leftarrow & {\bf |
526 |
v}(t) + \frac{\delta t}{2} \left( \frac{{\bf f}(t)}{m} - {\bf v}(t) |
527 |
\left(\chi(t) + \eta(t) \right) \right) \\ |
528 |
{\bf j}\left(t + \delta t / 2 \right) & \leftarrow & {\bf |
529 |
j}(t) + \frac{\delta t}{2} \left( {\bf \tau}^b(t) - {\bf j}(t) |
530 |
\chi(t) \right) \\ |
531 |
\mathsf{A}(t + \delta t) & \leftarrow & \mathrm{rot}\left(\delta t * |
532 |
{\bf j}(t + \delta t / 2) \overleftrightarrow{\mathsf{I}}^{b} \right) \\ |
533 |
\chi\left(t + \delta t / 2 \right) & \leftarrow & \chi(t) + |
534 |
\frac{\delta t}{2 \tau_T^2} \left( \frac{T(t)}{T_{\mathrm{target}}} - 1 |
535 |
\right) \\ |
536 |
\eta(t + \delta t / 2) & \leftarrow & \eta(t) + \frac{\delta t V(t)}{2 N k_B |
537 |
T(t) \tau_B^2} \left( P(t) - P_{\mathrm{target}} \right) \\ |
538 |
{\bf r}(t + \delta t) & \leftarrow & {\bf r}(t) + \delta t \left\{ {\bf |
539 |
v}\left(t + \delta t / 2 \right) + \eta(t + \delta t / 2)\left[ {\bf |
540 |
r}(t + \delta t) - {\bf R}_0 \right] \right\} \\ |
541 |
\mathsf{H}(t + \delta t) & \leftarrow & e^{-\delta t \eta(t + \delta t |
542 |
/ 2)} \mathsf{H}(t) |
543 |
\end{eqnarray} |
544 |
|
545 |
Most of these equations are identical to their counterparts in the NVT |
546 |
integrator, but the propagation of positions to time $t + \delta t$ |
547 |
depends on the positions at the same time. {\sc oopse} carries out |
548 |
this step iteratively (with a limit of 5 passes through the iterative |
549 |
loop). Also, the simulation box $\mathsf{H}$ is scaled uniformly for |
550 |
one full time step by an exponential factor that depends on the value |
551 |
of $\eta$ at time $t + |
552 |
\delta t / 2$. Reshaping the box uniformly also scales the volume of |
553 |
the box by |
554 |
\begin{equation} |
555 |
V(t + \delta t) \leftarrow e^{ - 3 \delta t \eta(t + \delta t /2)} |
556 |
V(t) |
557 |
\end{equation} |
558 |
|
559 |
The {\tt doForces} step for the NPTi integrator is exactly the same as |
560 |
in the DLM and NVT integrators. Once the forces and torques have been |
561 |
obtained at the new time step, the velocities can be advanced to the |
562 |
same time value. |
563 |
|
564 |
{\tt moveB:} |
565 |
\begin{eqnarray} |
566 |
T(t + \delta t) & \leftarrow & \left\{{\bf v}(t + \delta t)\right\}, |
567 |
\left\{{\bf j}(t + \delta t)\right\} \\ |
568 |
P(t + \delta t) & \leftarrow & \left\{{\bf r}(t + \delta t)\right\}, |
569 |
\left\{{\bf v}(t + \delta t)\right\}, \left\{{\bf f}(t + \delta t)\right\} \\ |
570 |
\chi\left(t + \delta t \right) & \leftarrow & \chi\left(t + \delta t / |
571 |
2 \right) + \frac{\delta t}{2 \tau_T^2} \left( \frac{T(t+\delta |
572 |
t)}{T_{\mathrm{target}}} - 1 \right) \\ |
573 |
\eta(t + \delta t) & \leftarrow & \eta(t + \delta t / 2) + |
574 |
\frac{\delta t V(t + \delta t)}{2 N k_B T(t + \delta t) \tau_B^2} |
575 |
\left( P(t + \delta t) - P_{\mathrm{target}} |
576 |
\right) \\ |
577 |
{\bf v}\left(t + \delta t \right) & \leftarrow & {\bf |
578 |
v}\left(t + \delta t / 2 \right) + \frac{\delta t}{2} \left( |
579 |
\frac{{\bf f}(t + \delta t)}{m} - {\bf v}(t + \delta t) |
580 |
(\chi(t + \delta t) + \eta(t + \delta t)) \right) \\ |
581 |
{\bf j}\left(t + \delta t \right) & \leftarrow & {\bf |
582 |
j}\left(t + \delta t / 2 \right) + \frac{\delta t}{2} \left( {\bf |
583 |
\tau}^b(t + \delta t) - {\bf j}(t + \delta t) |
584 |
\chi(t + \delta t) \right) |
585 |
\end{eqnarray} |
586 |
|
587 |
Once again, since ${\bf v}(t + \delta t)$ and ${\bf j}(t + \delta t)$ |
588 |
are required to caclculate $T(t + \delta t)$, $P(t + \delta t)$, $\chi(t + |
589 |
\delta t)$, and $\eta(t + \delta t)$, they indirectly depend on their |
590 |
own values at time $t + \delta t$. {\tt moveB} is therefore done in |
591 |
an iterative fashion until $\chi(t + \delta t)$ and $\eta(t + \delta |
592 |
t)$ become self-consistent. The relative tolerance for the |
593 |
self-consistency check defaults to a value of $\mbox{10}^{-6}$, but |
594 |
{\sc oopse} will terminate the iteration after 4 loops even if the |
595 |
consistency check has not been satisfied. |
596 |
|
597 |
The Melchionna modification of the Nos\'e-Hoover-Andersen algorithm is |
598 |
known to conserve a Hamiltonian for the extended system that is, to |
599 |
within a constant, identical to the Gibbs free energy, |
600 |
\begin{equation} |
601 |
H_{\mathrm{NPTi}} = V + K + f k_B T_{\mathrm{target}} \left( |
602 |
\frac{\tau_{T}^2 \chi^2(t)}{2} + \int_{0}^{t} \chi(t\prime) dt\prime |
603 |
\right) + P_{\mathrm{target}} V(t). |
604 |
\end{equation} |
605 |
Poor choices of $\delta t$, $\tau_T$, or $\tau_B$ can result in |
606 |
non-conservation of $H_{\mathrm{NPTi}}$, so the conserved quantity is |
607 |
maintained in the last column of the {\tt .stat} file to allow checks |
608 |
on the quality of the integration. It is also known that this |
609 |
algorithm samples the equilibrium distribution for the enthalpy |
610 |
(including contributions for the thermostat and barostat), |
611 |
\begin{equation} |
612 |
H_{\mathrm{NPTi}} = V + K + \frac{f k_B T_{\mathrm{target}}}{2} \left( |
613 |
\chi^2 \tau_T^2 + \eta^2 \tau_B^2 \right) + P_{\mathrm{target}} V(t). |
614 |
\end{equation} |
615 |
|
616 |
Bond constraints are applied at the end of both the {\tt moveA} and |
617 |
{\tt moveB} portions of the algorithm. Details on the constraint |
618 |
algorithms are given in section \ref{sec:rattle}. |
619 |
+ |
620 |
\subsection{\label{Sec:zcons}Z-Constraint Method} |
621 |
|
622 |
Based on fluctuatin-dissipation theorem,\bigskip\ force auto-correlation |
623 |
method was developed to investigate the dynamics of ions inside the ion |
624 |
channels.\cite{Roux91} Time-dependent friction coefficient can be calculated |
625 |
from the deviation of the instaneous force from its mean force. |
626 |
|
627 |
% |
628 |
|
629 |
\begin{equation} |
630 |
\xi(z,t)=\langle\delta F(z,t)\delta F(z,0)\rangle/k_{B}T |
631 |
\end{equation} |
632 |
|
633 |
|
634 |
where% |
635 |
\begin{equation} |
636 |
\delta F(z,t)=F(z,t)-\langle F(z,t)\rangle |
637 |
\end{equation} |
638 |
|
639 |
|
640 |
If the time-dependent friction decay rapidly, static friction coefficient can |
641 |
be approximated by% |
642 |
|
643 |
\begin{equation} |
644 |
\xi^{static}(z)=\int_{0}^{\infty}\langle\delta F(z,t)\delta F(z,0)\rangle dt |
645 |
\end{equation} |
646 |
|
647 |
|
648 |
Hence, diffusion constant can be estimated by |
649 |
\begin{equation} |
650 |
D(z)=\frac{k_{B}T}{\xi^{static}(z)}=\frac{(k_{B}T)^{2}}{\int_{0}^{\infty |
651 |
}\langle\delta F(z,t)\delta F(z,0)\rangle dt}% |
652 |
\end{equation} |
653 |
|
654 |
|
655 |
\bigskip Z-Constraint method, which fixed the z coordinates of the molecules |
656 |
with respect to the center of the mass of the system, was proposed to obtain |
657 |
the forces required in force auto-correlation method.\cite{Marrink94} However, |
658 |
simply resetting the coordinate will move the center of the mass of the whole |
659 |
system. To avoid this problem, a new method was used at {\sc oopse}. Instead of |
660 |
resetting the coordinate, we reset the forces of z-constraint molecules as |
661 |
well as subtract the total constraint forces from the rest of the system after |
662 |
force calculation at each time step. |
663 |
\begin{verbatim} |
664 |
$F_{\alpha i}=0$ |
665 |
$V_{\alpha i}=V_{\alpha i}-\frac{\sum\limits_{i}M_{_{\alpha i}}V_{\alpha i}}{\sum\limits_{i}M_{_{\alpha i}}}$ |
666 |
$F_{\alpha i}=F_{\alpha i}-\frac{M_{_{\alpha i}}}{\sum\limits_{\alpha}\sum\limits_{i}M_{_{\alpha i}}}\sum\limits_{\beta}F_{\beta}$ |
667 |
$V_{\alpha i}=V_{\alpha i}-\frac{\sum\limits_{\alpha}\sum\limits_{i}M_{_{\alpha i}}V_{\alpha i}}{\sum\limits_{\alpha}\sum\limits_{i}M_{_{\alpha i}}}$ |
668 |
\end{verbatim} |
669 |
|
670 |
At the very beginning of the simulation, the molecules may not be at its |
671 |
constraint position. To move the z-constraint molecule to the specified |
672 |
position, a simple harmonic potential is used% |
673 |
|
674 |
\begin{equation} |
675 |
U(t)=\frac{1}{2}k_{Harmonic}(z(t)-z_{cons})^{2}% |
676 |
\end{equation} |
677 |
where $k_{Harmonic}$\bigskip\ is the harmonic force constant, $z(t)$ is |
678 |
current z coordinate of the center of mass of the z-constraint molecule, and |
679 |
$z_{cons}$ is the restraint position. Therefore, the harmonic force operated |
680 |
on the z-constraint molecule at time $t$ can be calculated by% |
681 |
\begin{equation} |
682 |
F_{z_{Harmonic}}(t)=-\frac{\partial U(t)}{\partial z(t)}=-k_{Harmonic}% |
683 |
(z(t)-z_{cons}) |
684 |
\end{equation} |
685 |
Worthy of mention, other kinds of potential functions can also be used to |
686 |
drive the z-constraint molecule. |