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

Comparing trunk/tengDissertation/Methodology.tex (file contents):
Revision 2685 by tim, Mon Apr 3 18:07:54 2006 UTC vs.
Revision 2804 by tim, Tue Jun 6 19:47:27 2006 UTC

# Line 2 | Line 2
2  
3   \section{\label{methodSection:rigidBodyIntegrators}Integrators for Rigid Body Motion in Molecular Dynamics}
4  
5 < \section{\label{methodSection:conservedQuantities}Integrators to Conserve Properties in Special Ensembles}
5 > In order to mimic the experiments, which are usually performed under
6 > constant temperature and/or pressure, extended Hamiltonian system
7 > methods have been developed to generate statistical ensembles, such
8 > as canonical ensemble and isobaric-isothermal ensemble \textit{etc}.
9 > In addition to the standard ensemble, specific ensembles have been
10 > developed to account for the anisotropy between the lateral and
11 > normal directions of membranes. The $NPAT$ ensemble, in which the
12 > normal pressure and the lateral surface area of the membrane are
13 > kept constant, and the $NP\gamma T$ ensemble, in which the normal
14 > pressure and the lateral surface tension are kept constant were
15 > proposed to address this issue.
16  
17 < \section{\label{methodSection:hydrodynamics}Hydrodynamics}
17 > Integration schemes for rotational motion of the rigid molecules in
18 > microcanonical ensemble have been extensively studied in the last
19 > two decades. Matubayasi and Nakahara developed a time-reversible
20 > integrator for rigid bodies in quaternion representation. Although
21 > it is not symplectic, this integrator still demonstrates a better
22 > long-time energy conservation than traditional methods because of
23 > the time-reversible nature. Extending Trotter-Suzuki to general
24 > system with a flat phase space, Miller and his colleagues devised an
25 > novel symplectic, time-reversible and volume-preserving integrator
26 > in quaternion representation, which was shown to be superior to the
27 > time-reversible integrator of Matubayasi and Nakahara. However, all
28 > of the integrators in quaternion representation suffer from the
29 > computational penalty of constructing a rotation matrix from
30 > quaternions to evolve coordinates and velocities at every time step.
31 > An alternative integration scheme utilizing rotation matrix directly
32 > proposed by Dullweber, Leimkuhler and McLachlan (DLM) also preserved
33 > the same structural properties of the Hamiltonian flow. In this
34 > section, the integration scheme of DLM method will be reviewed and
35 > extended to other ensembles.
36  
37 < \section{\label{methodSection:langevin}Integrators for Langevin Dynamics of Rigid Bodies}
37 > \subsection{\label{methodSection:DLM}Integrating the Equations of Motion: the
38 > DLM method}
39  
40 < \section{\label{methodSection:coarseGrained}Coarse-Grained Modeling}
40 > The DLM method uses a Trotter factorization of the orientational
41 > propagator.  This has three effects:
42 > \begin{enumerate}
43 > \item the integrator is area-preserving in phase space (i.e. it is
44 > {\it symplectic}),
45 > \item the integrator is time-{\it reversible}, making it suitable for Hybrid
46 > Monte Carlo applications, and
47 > \item the error for a single time step is of order $\mathcal{O}\left(h^4\right)$
48 > for timesteps of length $h$.
49 > \end{enumerate}
50  
51 < \section{\label{methodSection:moleculeScale}Molecular-Scale Modeling}
51 > The integration of the equations of motion is carried out in a
52 > velocity-Verlet style 2-part algorithm, where $h= \delta t$:
53 >
54 > {\tt moveA:}
55 > \begin{align*}
56 > {\bf v}\left(t + h / 2\right)  &\leftarrow  {\bf v}(t)
57 >    + \frac{h}{2} \left( {\bf f}(t) / m \right), \\
58 > %
59 > {\bf r}(t + h) &\leftarrow {\bf r}(t)
60 >    + h  {\bf v}\left(t + h / 2 \right), \\
61 > %
62 > {\bf j}\left(t + h / 2 \right)  &\leftarrow {\bf j}(t)
63 >    + \frac{h}{2} {\bf \tau}^b(t), \\
64 > %
65 > \mathsf{A}(t + h) &\leftarrow \mathrm{rotate}\left( h {\bf j}
66 >    (t + h / 2) \cdot \overleftrightarrow{\mathsf{I}}^{-1} \right).
67 > \end{align*}
68 >
69 > In this context, the $\mathrm{rotate}$ function is the reversible
70 > product of the three body-fixed rotations,
71 > \begin{equation}
72 > \mathrm{rotate}({\bf a}) = \mathsf{G}_x(a_x / 2) \cdot
73 > \mathsf{G}_y(a_y / 2) \cdot \mathsf{G}_z(a_z) \cdot \mathsf{G}_y(a_y
74 > / 2) \cdot \mathsf{G}_x(a_x /2),
75 > \end{equation}
76 > where each rotational propagator, $\mathsf{G}_\alpha(\theta)$,
77 > rotates both the rotation matrix ($\mathsf{A}$) and the body-fixed
78 > angular momentum (${\bf j}$) by an angle $\theta$ around body-fixed
79 > axis $\alpha$,
80 > \begin{equation}
81 > \mathsf{G}_\alpha( \theta ) = \left\{
82 > \begin{array}{lcl}
83 > \mathsf{A}(t) & \leftarrow & \mathsf{A}(0) \cdot \mathsf{R}_\alpha(\theta)^T, \\
84 > {\bf j}(t) & \leftarrow & \mathsf{R}_\alpha(\theta) \cdot {\bf
85 > j}(0).
86 > \end{array}
87 > \right.
88 > \end{equation}
89 > $\mathsf{R}_\alpha$ is a quadratic approximation to the single-axis
90 > rotation matrix.  For example, in the small-angle limit, the
91 > rotation matrix around the body-fixed x-axis can be approximated as
92 > \begin{equation}
93 > \mathsf{R}_x(\theta) \approx \left(
94 > \begin{array}{ccc}
95 > 1 & 0 & 0 \\
96 > 0 & \frac{1-\theta^2 / 4}{1 + \theta^2 / 4}  & -\frac{\theta}{1+
97 > \theta^2 / 4} \\
98 > 0 & \frac{\theta}{1+ \theta^2 / 4} & \frac{1-\theta^2 / 4}{1 +
99 > \theta^2 / 4}
100 > \end{array}
101 > \right).
102 > \end{equation}
103 > All other rotations follow in a straightforward manner.
104 >
105 > After the first part of the propagation, the forces and body-fixed
106 > torques are calculated at the new positions and orientations
107 >
108 > {\tt doForces:}
109 > \begin{align*}
110 > {\bf f}(t + h) &\leftarrow
111 >    - \left(\frac{\partial V}{\partial {\bf r}}\right)_{{\bf r}(t + h)}, \\
112 > %
113 > {\bf \tau}^{s}(t + h) &\leftarrow {\bf u}(t + h)
114 >    \times \frac{\partial V}{\partial {\bf u}}, \\
115 > %
116 > {\bf \tau}^{b}(t + h) &\leftarrow \mathsf{A}(t + h)
117 >    \cdot {\bf \tau}^s(t + h).
118 > \end{align*}
119 >
120 > {\sc oopse} automatically updates ${\bf u}$ when the rotation matrix
121 > $\mathsf{A}$ is calculated in {\tt moveA}.  Once the forces and
122 > torques have been obtained at the new time step, the velocities can
123 > be advanced to the same time value.
124 >
125 > {\tt moveB:}
126 > \begin{align*}
127 > {\bf v}\left(t + h \right)  &\leftarrow  {\bf v}\left(t + h / 2
128 > \right)
129 >    + \frac{h}{2} \left( {\bf f}(t + h) / m \right), \\
130 > %
131 > {\bf j}\left(t + h \right)  &\leftarrow {\bf j}\left(t + h / 2
132 > \right)
133 >    + \frac{h}{2} {\bf \tau}^b(t + h) .
134 > \end{align*}
135 >
136 > The matrix rotations used in the DLM method end up being more costly
137 > computationally than the simpler arithmetic quaternion propagation.
138 > With the same time step, a 1000-molecule water simulation shows an
139 > average 7\% increase in computation time using the DLM method in
140 > place of quaternions. This cost is more than justified when
141 > comparing the energy conservation of the two methods as illustrated
142 > in Fig.~\ref{timestep}.
143 >
144 > \begin{figure}
145 > \centering
146 > \includegraphics[width=\linewidth]{timeStep.eps}
147 > \caption[Energy conservation for quaternion versus DLM
148 > dynamics]{Energy conservation using quaternion based integration
149 > versus the method proposed by Dullweber \emph{et al.} with
150 > increasing time step. For each time step, the dotted line is total
151 > energy using the DLM integrator, and the solid line comes from the
152 > quaternion integrator. The larger time step plots are shifted up
153 > from the true energy baseline for clarity.}
154 > \label{methodFig:timestep}
155 > \end{figure}
156 >
157 > In Fig.~\ref{methodFig:timestep}, the resulting energy drift at
158 > various time steps for both the DLM and quaternion integration
159 > schemes is compared. All of the 1000 molecule water simulations
160 > started with the same configuration, and the only difference was the
161 > method for handling rotational motion. At time steps of 0.1 and 0.5
162 > fs, both methods for propagating molecule rotation conserve energy
163 > fairly well, with the quaternion method showing a slight energy
164 > drift over time in the 0.5 fs time step simulation. At time steps of
165 > 1 and 2 fs, the energy conservation benefits of the DLM method are
166 > clearly demonstrated. Thus, while maintaining the same degree of
167 > energy conservation, one can take considerably longer time steps,
168 > leading to an overall reduction in computation time.
169 >
170 > \subsection{\label{methodSection:NVT}Nos\'{e}-Hoover Thermostatting}
171 >
172 > The Nos\'e-Hoover equations of motion are given by\cite{Hoover1985}
173 > \begin{eqnarray}
174 > \dot{{\bf r}} & = & {\bf v}, \\
175 > \dot{{\bf v}} & = & \frac{{\bf f}}{m} - \chi {\bf v} ,\\
176 > \dot{\mathsf{A}} & = & \mathsf{A} \cdot
177 > \mbox{ skew}\left(\overleftrightarrow{\mathsf{I}}^{-1} \cdot {\bf j}\right), \\
178 > \dot{{\bf j}} & = & {\bf j} \times \left(
179 > \overleftrightarrow{\mathsf{I}}^{-1} \cdot {\bf j} \right) - \mbox{
180 > rot}\left(\mathsf{A}^{T} \cdot \frac{\partial V}{\partial
181 > \mathsf{A}} \right) - \chi {\bf j}. \label{eq:nosehoovereom}
182 > \end{eqnarray}
183 >
184 > $\chi$ is an ``extra'' variable included in the extended system, and
185 > it is propagated using the first order equation of motion
186 > \begin{equation}
187 > \dot{\chi} = \frac{1}{\tau_{T}^2} \left(
188 > \frac{T}{T_{\mathrm{target}}} - 1 \right). \label{eq:nosehooverext}
189 > \end{equation}
190 >
191 > The instantaneous temperature $T$ is proportional to the total
192 > kinetic energy (both translational and orientational) and is given
193 > by
194 > \begin{equation}
195 > T = \frac{2 K}{f k_B}
196 > \end{equation}
197 > Here, $f$ is the total number of degrees of freedom in the system,
198 > \begin{equation}
199 > f = 3 N + 3 N_{\mathrm{orient}} - N_{\mathrm{constraints}},
200 > \end{equation}
201 > and $K$ is the total kinetic energy,
202 > \begin{equation}
203 > K = \sum_{i=1}^{N} \frac{1}{2} m_i {\bf v}_i^T \cdot {\bf v}_i +
204 > \sum_{i=1}^{N_{\mathrm{orient}}}  \frac{1}{2} {\bf j}_i^T \cdot
205 > \overleftrightarrow{\mathsf{I}}_i^{-1} \cdot {\bf j}_i.
206 > \end{equation}
207 >
208 > In eq.(\ref{eq:nosehooverext}), $\tau_T$ is the time constant for
209 > relaxation of the temperature to the target value.  To set values
210 > for $\tau_T$ or $T_{\mathrm{target}}$ in a simulation, one would use
211 > the {\tt tauThermostat} and {\tt targetTemperature} keywords in the
212 > {\tt .bass} file.  The units for {\tt tauThermostat} are fs, and the
213 > units for the {\tt targetTemperature} are degrees K.   The
214 > integration of the equations of motion is carried out in a
215 > velocity-Verlet style 2 part algorithm:
216 >
217 > {\tt moveA:}
218 > \begin{align*}
219 > T(t) &\leftarrow \left\{{\bf v}(t)\right\}, \left\{{\bf j}(t)\right\} ,\\
220 > %
221 > {\bf v}\left(t + h / 2\right)  &\leftarrow {\bf v}(t)
222 >    + \frac{h}{2} \left( \frac{{\bf f}(t)}{m} - {\bf v}(t)
223 >    \chi(t)\right), \\
224 > %
225 > {\bf r}(t + h) &\leftarrow {\bf r}(t)
226 >    + h {\bf v}\left(t + h / 2 \right) ,\\
227 > %
228 > {\bf j}\left(t + h / 2 \right)  &\leftarrow {\bf j}(t)
229 >    + \frac{h}{2} \left( {\bf \tau}^b(t) - {\bf j}(t)
230 >    \chi(t) \right) ,\\
231 > %
232 > \mathsf{A}(t + h) &\leftarrow \mathrm{rotate}
233 >    \left(h * {\bf j}(t + h / 2)
234 >    \overleftrightarrow{\mathsf{I}}^{-1} \right) ,\\
235 > %
236 > \chi\left(t + h / 2 \right) &\leftarrow \chi(t)
237 >    + \frac{h}{2 \tau_T^2} \left( \frac{T(t)}
238 >    {T_{\mathrm{target}}} - 1 \right) .
239 > \end{align*}
240 >
241 > Here $\mathrm{rotate}(h * {\bf j}
242 > \overleftrightarrow{\mathsf{I}}^{-1})$ is the same symplectic
243 > Trotter factorization of the three rotation operations that was
244 > discussed in the section on the DLM integrator.  Note that this
245 > operation modifies both the rotation matrix $\mathsf{A}$ and the
246 > angular momentum ${\bf j}$.  {\tt moveA} propagates velocities by a
247 > half time step, and positional degrees of freedom by a full time
248 > step.  The new positions (and orientations) are then used to
249 > calculate a new set of forces and torques in exactly the same way
250 > they are calculated in the {\tt doForces} portion of the DLM
251 > integrator.
252 >
253 > Once the forces and torques have been obtained at the new time step,
254 > the temperature, velocities, and the extended system variable can be
255 > advanced to the same time value.
256 >
257 > {\tt moveB:}
258 > \begin{align*}
259 > T(t + h) &\leftarrow \left\{{\bf v}(t + h)\right\},
260 >    \left\{{\bf j}(t + h)\right\}, \\
261 > %
262 > \chi\left(t + h \right) &\leftarrow \chi\left(t + h /
263 >    2 \right) + \frac{h}{2 \tau_T^2} \left( \frac{T(t+h)}
264 >    {T_{\mathrm{target}}} - 1 \right), \\
265 > %
266 > {\bf v}\left(t + h \right)  &\leftarrow {\bf v}\left(t
267 >    + h / 2 \right) + \frac{h}{2} \left(
268 >    \frac{{\bf f}(t + h)}{m} - {\bf v}(t + h)
269 >    \chi(t h)\right) ,\\
270 > %
271 > {\bf j}\left(t + h \right) &\leftarrow {\bf j}\left(t
272 >    + h / 2 \right) + \frac{h}{2}
273 >    \left( {\bf \tau}^b(t + h) - {\bf j}(t + h)
274 >    \chi(t + h) \right) .
275 > \end{align*}
276 >
277 > Since ${\bf v}(t + h)$ and ${\bf j}(t + h)$ are required to
278 > caclculate $T(t + h)$ as well as $\chi(t + h)$, they indirectly
279 > depend on their own values at time $t + h$.  {\tt moveB} is
280 > therefore done in an iterative fashion until $\chi(t + h)$ becomes
281 > self-consistent.  The relative tolerance for the self-consistency
282 > check defaults to a value of $\mbox{10}^{-6}$, but {\sc oopse} will
283 > terminate the iteration after 4 loops even if the consistency check
284 > has not been satisfied.
285 >
286 > The Nos\'e-Hoover algorithm is known to conserve a Hamiltonian for
287 > the extended system that is, to within a constant, identical to the
288 > Helmholtz free energy,\cite{Melchionna1993}
289 > \begin{equation}
290 > H_{\mathrm{NVT}} = V + K + f k_B T_{\mathrm{target}} \left(
291 > \frac{\tau_{T}^2 \chi^2(t)}{2} + \int_{0}^{t} \chi(t^\prime)
292 > dt^\prime \right).
293 > \end{equation}
294 > Poor choices of $h$ or $\tau_T$ can result in non-conservation of
295 > $H_{\mathrm{NVT}}$, so the conserved quantity is maintained in the
296 > last column of the {\tt .stat} file to allow checks on the quality
297 > of the integration.
298 >
299 > \subsection{\label{methodSection:NPTi}Constant-pressure integration with
300 > isotropic box deformations (NPTi)}
301 >
302 > To carry out isobaric-isothermal ensemble calculations {\sc oopse}
303 > implements the Melchionna modifications to the
304 > Nos\'e-Hoover-Andersen equations of motion,\cite{Melchionna1993}
305 >
306 > \begin{eqnarray}
307 > \dot{{\bf r}} & = & {\bf v} + \eta \left( {\bf r} - {\bf R}_0 \right), \\
308 > \dot{{\bf v}} & = & \frac{{\bf f}}{m} - (\eta + \chi) {\bf v}, \\
309 > \dot{\mathsf{A}} & = & \mathsf{A} \cdot
310 > \mbox{ skew}\left(\overleftrightarrow{I}^{-1} \cdot {\bf j}\right),\\
311 > \dot{{\bf j}} & = & {\bf j} \times \left(
312 > \overleftrightarrow{I}^{-1} \cdot {\bf j} \right) - \mbox{
313 > rot}\left(\mathsf{A}^{T} \cdot \frac{\partial
314 > V}{\partial \mathsf{A}} \right) - \chi {\bf j}, \\
315 > \dot{\chi} & = & \frac{1}{\tau_{T}^2} \left(
316 > \frac{T}{T_{\mathrm{target}}} - 1 \right) ,\\
317 > \dot{\eta} & = & \frac{1}{\tau_{B}^2 f k_B T_{\mathrm{target}}} V
318 > \left( P -
319 > P_{\mathrm{target}} \right), \\
320 > \dot{\mathcal{V}} & = & 3 \mathcal{V} \eta . \label{eq:melchionna1}
321 > \end{eqnarray}
322 >
323 > $\chi$ and $\eta$ are the ``extra'' degrees of freedom in the
324 > extended system.  $\chi$ is a thermostat, and it has the same
325 > function as it does in the Nos\'e-Hoover NVT integrator.  $\eta$ is
326 > a barostat which controls changes to the volume of the simulation
327 > box.  ${\bf R}_0$ is the location of the center of mass for the
328 > entire system, and $\mathcal{V}$ is the volume of the simulation
329 > box.  At any time, the volume can be calculated from the determinant
330 > of the matrix which describes the box shape:
331 > \begin{equation}
332 > \mathcal{V} = \det(\mathsf{H}).
333 > \end{equation}
334 >
335 > The NPTi integrator requires an instantaneous pressure. This
336 > quantity is calculated via the pressure tensor,
337 > \begin{equation}
338 > \overleftrightarrow{\mathsf{P}}(t) = \frac{1}{\mathcal{V}(t)} \left(
339 > \sum_{i=1}^{N} m_i {\bf v}_i(t) \otimes {\bf v}_i(t) \right) +
340 > \overleftrightarrow{\mathsf{W}}(t).
341 > \end{equation}
342 > The kinetic contribution to the pressure tensor utilizes the {\it
343 > outer} product of the velocities denoted by the $\otimes$ symbol.
344 > The stress tensor is calculated from another outer product of the
345 > inter-atomic separation vectors (${\bf r}_{ij} = {\bf r}_j - {\bf
346 > r}_i$) with the forces between the same two atoms,
347 > \begin{equation}
348 > \overleftrightarrow{\mathsf{W}}(t) = \sum_{i} \sum_{j>i} {\bf
349 > r}_{ij}(t) \otimes {\bf f}_{ij}(t).
350 > \end{equation}
351 > The instantaneous pressure is then simply obtained from the trace of
352 > the Pressure tensor,
353 > \begin{equation}
354 > P(t) = \frac{1}{3} \mathrm{Tr} \left(
355 > \overleftrightarrow{\mathsf{P}}(t). \right)
356 > \end{equation}
357 >
358 > In eq.(\ref{eq:melchionna1}), $\tau_B$ is the time constant for
359 > relaxation of the pressure to the target value.  To set values for
360 > $\tau_B$ or $P_{\mathrm{target}}$ in a simulation, one would use the
361 > {\tt tauBarostat} and {\tt targetPressure} keywords in the {\tt
362 > .bass} file.  The units for {\tt tauBarostat} are fs, and the units
363 > for the {\tt targetPressure} are atmospheres.  Like in the NVT
364 > integrator, the integration of the equations of motion is carried
365 > out in a velocity-Verlet style 2 part algorithm:
366 >
367 > {\tt moveA:}
368 > \begin{align*}
369 > T(t) &\leftarrow \left\{{\bf v}(t)\right\}, \left\{{\bf j}(t)\right\} ,\\
370 > %
371 > P(t) &\leftarrow \left\{{\bf r}(t)\right\}, \left\{{\bf v}(t)\right\} ,\\
372 > %
373 > {\bf v}\left(t + h / 2\right)  &\leftarrow {\bf v}(t)
374 >    + \frac{h}{2} \left( \frac{{\bf f}(t)}{m} - {\bf v}(t)
375 >    \left(\chi(t) + \eta(t) \right) \right), \\
376 > %
377 > {\bf j}\left(t + h / 2 \right)  &\leftarrow {\bf j}(t)
378 >    + \frac{h}{2} \left( {\bf \tau}^b(t) - {\bf j}(t)
379 >    \chi(t) \right), \\
380 > %
381 > \mathsf{A}(t + h) &\leftarrow \mathrm{rotate}\left(h *
382 >    {\bf j}(t + h / 2) \overleftrightarrow{\mathsf{I}}^{-1}
383 >    \right) ,\\
384 > %
385 > \chi\left(t + h / 2 \right) &\leftarrow \chi(t) +
386 >    \frac{h}{2 \tau_T^2} \left( \frac{T(t)}{T_{\mathrm{target}}} - 1
387 >    \right) ,\\
388 > %
389 > \eta(t + h / 2) &\leftarrow \eta(t) + \frac{h
390 >    \mathcal{V}(t)}{2 N k_B T(t) \tau_B^2} \left( P(t)
391 >    - P_{\mathrm{target}} \right), \\
392 > %
393 > {\bf r}(t + h) &\leftarrow {\bf r}(t) + h
394 >    \left\{ {\bf v}\left(t + h / 2 \right)
395 >    + \eta(t + h / 2)\left[ {\bf r}(t + h)
396 >    - {\bf R}_0 \right] \right\} ,\\
397 > %
398 > \mathsf{H}(t + h) &\leftarrow e^{-h \eta(t + h / 2)}
399 >    \mathsf{H}(t).
400 > \end{align*}
401 >
402 > Most of these equations are identical to their counterparts in the
403 > NVT integrator, but the propagation of positions to time $t + h$
404 > depends on the positions at the same time.  {\sc oopse} carries out
405 > this step iteratively (with a limit of 5 passes through the
406 > iterative loop).  Also, the simulation box $\mathsf{H}$ is scaled
407 > uniformly for one full time step by an exponential factor that
408 > depends on the value of $\eta$ at time $t + h / 2$.  Reshaping the
409 > box uniformly also scales the volume of the box by
410 > \begin{equation}
411 > \mathcal{V}(t + h) \leftarrow e^{ - 3 h \eta(t + h /2)}.
412 > \mathcal{V}(t)
413 > \end{equation}
414 >
415 > The {\tt doForces} step for the NPTi integrator is exactly the same
416 > as in both the DLM and NVT integrators.  Once the forces and torques
417 > have been obtained at the new time step, the velocities can be
418 > advanced to the same time value.
419 >
420 > {\tt moveB:}
421 > \begin{align*}
422 > T(t + h) &\leftarrow \left\{{\bf v}(t + h)\right\},
423 >    \left\{{\bf j}(t + h)\right\} ,\\
424 > %
425 > P(t + h) &\leftarrow  \left\{{\bf r}(t + h)\right\},
426 >    \left\{{\bf v}(t + h)\right\}, \\
427 > %
428 > \chi\left(t + h \right) &\leftarrow \chi\left(t + h /
429 >    2 \right) + \frac{h}{2 \tau_T^2} \left( \frac{T(t+h)}
430 >    {T_{\mathrm{target}}} - 1 \right), \\
431 > %
432 > \eta(t + h) &\leftarrow \eta(t + h / 2) +
433 >    \frac{h \mathcal{V}(t + h)}{2 N k_B T(t + h)
434 >    \tau_B^2} \left( P(t + h) - P_{\mathrm{target}} \right), \\
435 > %
436 > {\bf v}\left(t + h \right)  &\leftarrow {\bf v}\left(t
437 >    + h / 2 \right) + \frac{h}{2} \left(
438 >    \frac{{\bf f}(t + h)}{m} - {\bf v}(t + h)
439 >    (\chi(t + h) + \eta(t + h)) \right) ,\\
440 > %
441 > {\bf j}\left(t + h \right)  &\leftarrow {\bf j}\left(t
442 >    + h / 2 \right) + \frac{h}{2} \left( {\bf
443 >    \tau}^b(t + h) - {\bf j}(t + h)
444 >    \chi(t + h) \right) .
445 > \end{align*}
446 >
447 > Once again, since ${\bf v}(t + h)$ and ${\bf j}(t + h)$ are required
448 > to caclculate $T(t + h)$, $P(t + h)$, $\chi(t + h)$, and $\eta(t +
449 > h)$, they indirectly depend on their own values at time $t + h$.
450 > {\tt moveB} is therefore done in an iterative fashion until $\chi(t
451 > + h)$ and $\eta(t + h)$ become self-consistent.  The relative
452 > tolerance for the self-consistency check defaults to a value of
453 > $\mbox{10}^{-6}$, but {\sc oopse} will terminate the iteration after
454 > 4 loops even if the consistency check has not been satisfied.
455 >
456 > The Melchionna modification of the Nos\'e-Hoover-Andersen algorithm
457 > is known to conserve a Hamiltonian for the extended system that is,
458 > to within a constant, identical to the Gibbs free energy,
459 > \begin{equation}
460 > H_{\mathrm{NPTi}} = V + K + f k_B T_{\mathrm{target}} \left(
461 > \frac{\tau_{T}^2 \chi^2(t)}{2} + \int_{0}^{t} \chi(t^\prime)
462 > dt^\prime \right) + P_{\mathrm{target}} \mathcal{V}(t).
463 > \end{equation}
464 > Poor choices of $\delta t$, $\tau_T$, or $\tau_B$ can result in
465 > non-conservation of $H_{\mathrm{NPTi}}$, so the conserved quantity
466 > is maintained in the last column of the {\tt .stat} file to allow
467 > checks on the quality of the integration.  It is also known that
468 > this algorithm samples the equilibrium distribution for the enthalpy
469 > (including contributions for the thermostat and barostat),
470 > \begin{equation}
471 > H_{\mathrm{NPTi}} = V + K + \frac{f k_B T_{\mathrm{target}}}{2}
472 > \left( \chi^2 \tau_T^2 + \eta^2 \tau_B^2 \right) +
473 > P_{\mathrm{target}} \mathcal{V}(t).
474 > \end{equation}
475 >
476 > Bond constraints are applied at the end of both the {\tt moveA} and
477 > {\tt moveB} portions of the algorithm.  Details on the constraint
478 > algorithms are given in section \ref{oopseSec:rattle}.
479 >
480 > \subsection{\label{methodSection:NPTf}Constant-pressure integration with a
481 > flexible box (NPTf)}
482 >
483 > There is a relatively simple generalization of the
484 > Nos\'e-Hoover-Andersen method to include changes in the simulation
485 > box {\it shape} as well as in the volume of the box.  This method
486 > utilizes the full $3 \times 3$ pressure tensor and introduces a
487 > tensor of extended variables ($\overleftrightarrow{\eta}$) to
488 > control changes to the box shape.  The equations of motion for this
489 > method are
490 > \begin{eqnarray}
491 > \dot{{\bf r}} & = & {\bf v} + \overleftrightarrow{\eta} \cdot \left( {\bf r} - {\bf R}_0 \right), \\
492 > \dot{{\bf v}} & = & \frac{{\bf f}}{m} - (\overleftrightarrow{\eta} +
493 > \chi \cdot \mathsf{1}) {\bf v}, \\
494 > \dot{\mathsf{A}} & = & \mathsf{A} \cdot
495 > \mbox{ skew}\left(\overleftrightarrow{I}^{-1} \cdot {\bf j}\right) ,\\
496 > \dot{{\bf j}} & = & {\bf j} \times \left(
497 > \overleftrightarrow{I}^{-1} \cdot {\bf j} \right) - \mbox{
498 > rot}\left(\mathsf{A}^{T} \cdot \frac{\partial
499 > V}{\partial \mathsf{A}} \right) - \chi {\bf j} ,\\
500 > \dot{\chi} & = & \frac{1}{\tau_{T}^2} \left(
501 > \frac{T}{T_{\mathrm{target}}} - 1 \right) ,\\
502 > \dot{\overleftrightarrow{\eta}} & = & \frac{1}{\tau_{B}^2 f k_B
503 > T_{\mathrm{target}}} V \left( \overleftrightarrow{\mathsf{P}} - P_{\mathrm{target}}\mathsf{1} \right) ,\\
504 > \dot{\mathsf{H}} & = &  \overleftrightarrow{\eta} \cdot \mathsf{H} .
505 > \label{eq:melchionna2}
506 > \end{eqnarray}
507 >
508 > Here, $\mathsf{1}$ is the unit matrix and
509 > $\overleftrightarrow{\mathsf{P}}$ is the pressure tensor.  Again,
510 > the volume, $\mathcal{V} = \det \mathsf{H}$.
511 >
512 > The propagation of the equations of motion is nearly identical to
513 > the NPTi integration:
514 >
515 > {\tt moveA:}
516 > \begin{align*}
517 > T(t) &\leftarrow \left\{{\bf v}(t)\right\}, \left\{{\bf j}(t)\right\} ,\\
518 > %
519 > \overleftrightarrow{\mathsf{P}}(t) &\leftarrow \left\{{\bf
520 > r}(t)\right\},
521 >    \left\{{\bf v}(t)\right\} ,\\
522 > %
523 > {\bf v}\left(t + h / 2\right)  &\leftarrow {\bf v}(t)
524 >    + \frac{h}{2} \left( \frac{{\bf f}(t)}{m} -
525 >    \left(\chi(t)\mathsf{1} + \overleftrightarrow{\eta}(t) \right) \cdot
526 >    {\bf v}(t) \right), \\
527 > %
528 > {\bf j}\left(t + h / 2 \right)  &\leftarrow {\bf j}(t)
529 >    + \frac{h}{2} \left( {\bf \tau}^b(t) - {\bf j}(t)
530 >    \chi(t) \right), \\
531 > %
532 > \mathsf{A}(t + h) &\leftarrow \mathrm{rotate}\left(h *
533 >    {\bf j}(t + h / 2) \overleftrightarrow{\mathsf{I}}^{-1}
534 >    \right), \\
535 > %
536 > \chi\left(t + h / 2 \right) &\leftarrow \chi(t) +
537 >    \frac{h}{2 \tau_T^2} \left( \frac{T(t)}{T_{\mathrm{target}}}
538 >    - 1 \right), \\
539 > %
540 > \overleftrightarrow{\eta}(t + h / 2) &\leftarrow
541 >    \overleftrightarrow{\eta}(t) + \frac{h \mathcal{V}(t)}{2 N k_B
542 >    T(t) \tau_B^2} \left( \overleftrightarrow{\mathsf{P}}(t)
543 >    - P_{\mathrm{target}}\mathsf{1} \right), \\
544 > %
545 > {\bf r}(t + h) &\leftarrow {\bf r}(t) + h \left\{ {\bf v}
546 >    \left(t + h / 2 \right) + \overleftrightarrow{\eta}(t +
547 >    h / 2) \cdot \left[ {\bf r}(t + h)
548 >    - {\bf R}_0 \right] \right\}, \\
549 > %
550 > \mathsf{H}(t + h) &\leftarrow \mathsf{H}(t) \cdot e^{-h
551 >    \overleftrightarrow{\eta}(t + h / 2)} .
552 > \end{align*}
553 > {\sc oopse} uses a power series expansion truncated at second order
554 > for the exponential operation which scales the simulation box.
555 >
556 > The {\tt moveB} portion of the algorithm is largely unchanged from
557 > the NPTi integrator:
558 >
559 > {\tt moveB:}
560 > \begin{align*}
561 > T(t + h) &\leftarrow \left\{{\bf v}(t + h)\right\},
562 >    \left\{{\bf j}(t + h)\right\}, \\
563 > %
564 > \overleftrightarrow{\mathsf{P}}(t + h) &\leftarrow \left\{{\bf r}
565 >    (t + h)\right\}, \left\{{\bf v}(t
566 >    + h)\right\}, \left\{{\bf f}(t + h)\right\} ,\\
567 > %
568 > \chi\left(t + h \right) &\leftarrow \chi\left(t + h /
569 >    2 \right) + \frac{h}{2 \tau_T^2} \left( \frac{T(t+
570 >    h)}{T_{\mathrm{target}}} - 1 \right), \\
571 > %
572 > \overleftrightarrow{\eta}(t + h) &\leftarrow
573 >    \overleftrightarrow{\eta}(t + h / 2) +
574 >    \frac{h \mathcal{V}(t + h)}{2 N k_B T(t + h)
575 >    \tau_B^2} \left( \overleftrightarrow{P}(t + h)
576 >    - P_{\mathrm{target}}\mathsf{1} \right) ,\\
577 > %
578 > {\bf v}\left(t + h \right)  &\leftarrow {\bf v}\left(t
579 >    + h / 2 \right) + \frac{h}{2} \left(
580 >    \frac{{\bf f}(t + h)}{m} -
581 >    (\chi(t + h)\mathsf{1} + \overleftrightarrow{\eta}(t
582 >    + h)) \right) \cdot {\bf v}(t + h), \\
583 > %
584 > {\bf j}\left(t + h \right)  &\leftarrow {\bf j}\left(t
585 >    + h / 2 \right) + \frac{h}{2} \left( {\bf \tau}^b(t
586 >    + h) - {\bf j}(t + h) \chi(t + h) \right) .
587 > \end{align*}
588 >
589 > The iterative schemes for both {\tt moveA} and {\tt moveB} are
590 > identical to those described for the NPTi integrator.
591 >
592 > The NPTf integrator is known to conserve the following Hamiltonian:
593 > \begin{equation}
594 > H_{\mathrm{NPTf}} = V + K + f k_B T_{\mathrm{target}} \left(
595 > \frac{\tau_{T}^2 \chi^2(t)}{2} + \int_{0}^{t} \chi(t^\prime)
596 > dt^\prime \right) + P_{\mathrm{target}} \mathcal{V}(t) + \frac{f k_B
597 > T_{\mathrm{target}}}{2}
598 > \mathrm{Tr}\left[\overleftrightarrow{\eta}(t)\right]^2 \tau_B^2.
599 > \end{equation}
600 >
601 > This integrator must be used with care, particularly in liquid
602 > simulations.  Liquids have very small restoring forces in the
603 > off-diagonal directions, and the simulation box can very quickly
604 > form elongated and sheared geometries which become smaller than the
605 > electrostatic or Lennard-Jones cutoff radii.  The NPTf integrator
606 > finds most use in simulating crystals or liquid crystals which
607 > assume non-orthorhombic geometries.
608 >
609 > \subsection{\label{methodSection:otherSpecialEnsembles}Other Special Ensembles}
610 >
611 > \subsubsection{\label{methodSection:NPAT}NPAT Ensemble}
612 >
613 > A comprehensive understanding of structure¨Cfunction relations of
614 > biological membrane system ultimately relies on structure and
615 > dynamics of lipid bilayer, which are strongly affected by the
616 > interfacial interaction between lipid molecules and surrounding
617 > media. One quantity to describe the interfacial interaction is so
618 > called the average surface area per lipid. Constat area and constant
619 > lateral pressure simulation can be achieved by extending the
620 > standard NPT ensemble with a different pressure control strategy
621 >
622 > \begin{equation}
623 > \dot {\overleftrightarrow{\eta}} _{\alpha \beta}=\left\{\begin{array}{ll}
624 >                  \frac{{V(P_{\alpha \beta }  - P_{{\rm{target}}} )}}{{\tau_{\rm{B}}^{\rm{2}} fk_B T_{{\rm{target}}} }}
625 >                  & \mbox{if $ \alpha = \beta  = z$}\\
626 >                  0 & \mbox{otherwise}\\
627 >           \end{array}
628 >    \right.
629 > \end{equation}
630 >
631 > Note that the iterative schemes for NPAT are identical to those
632 > described for the NPTi integrator.
633 >
634 > \subsubsection{\label{methodSection:NPrT}NP$\gamma$T Ensemble}
635 >
636 > Theoretically, the surface tension $\gamma$ of a stress free
637 > membrane system should be zero since its surface free energy $G$ is
638 > minimum with respect to surface area $A$
639 > \[
640 > \gamma  = \frac{{\partial G}}{{\partial A}}.
641 > \]
642 > However, a surface tension of zero is not appropriate for relatively
643 > small patches of membrane. In order to eliminate the edge effect of
644 > the membrane simulation, a special ensemble, NP$\gamma$T, is
645 > proposed to maintain the lateral surface tension and normal
646 > pressure. The equation of motion for cell size control tensor,
647 > $\eta$, in $NP\gamma T$ is
648 > \begin{equation}
649 > \dot {\overleftrightarrow{\eta}} _{\alpha \beta}=\left\{\begin{array}{ll}
650 >    - A_{xy} (\gamma _\alpha   - \gamma _{{\rm{target}}} ) & \mbox{$\alpha  = \beta  = x$ or $\alpha  = \beta  = y$}\\
651 >    \frac{{V(P_{\alpha \beta }  - P_{{\rm{target}}} )}}{{\tau _{\rm{B}}^{\rm{2}} fk_B T_{{\rm{target}}}}} & \mbox{$\alpha  = \beta  = z$} \\
652 >    0 & \mbox{$\alpha  \ne \beta$} \\
653 >       \end{array}
654 >    \right.
655 > \end{equation}
656 > where $ \gamma _{{\rm{target}}}$ is the external surface tension and
657 > the instantaneous surface tensor $\gamma _\alpha$ is given by
658 > \begin{equation}
659 > \gamma _\alpha   =  - h_z( \overleftrightarrow{P} _{\alpha \alpha }
660 > - P_{{\rm{target}}} )
661 > \label{methodEquation:instantaneousSurfaceTensor}
662 > \end{equation}
663 >
664 > There is one additional extended system integrator (NPTxyz), in
665 > which each attempt to preserve the target pressure along the box
666 > walls perpendicular to that particular axis.  The lengths of the box
667 > axes are allowed to fluctuate independently, but the angle between
668 > the box axes does not change. It should be noted that the NPTxyz
669 > integrator is a special case of $NP\gamma T$ if the surface tension
670 > $\gamma$ is set to zero.
671 >
672 > \section{\label{methodSection:zcons}Z-Constraint Method}
673 >
674 > Based on the fluctuation-dissipation theorem, a force
675 > auto-correlation method was developed by Roux and Karplus to
676 > investigate the dynamics of ions inside ion channels\cite{Roux1991}.
677 > The time-dependent friction coefficient can be calculated from the
678 > deviation of the instantaneous force from its mean force.
679 > \begin{equation}
680 > \xi(z,t)=\langle\delta F(z,t)\delta F(z,0)\rangle/k_{B}T,
681 > \end{equation}
682 > where%
683 > \begin{equation}
684 > \delta F(z,t)=F(z,t)-\langle F(z,t)\rangle.
685 > \end{equation}
686 >
687 > If the time-dependent friction decays rapidly, the static friction
688 > coefficient can be approximated by
689 > \begin{equation}
690 > \xi_{\text{static}}(z)=\int_{0}^{\infty}\langle\delta F(z,t)\delta
691 > F(z,0)\rangle dt.
692 > \end{equation}
693 > Allowing diffusion constant to then be calculated through the
694 > Einstein relation:\cite{Marrink1994}
695 > \begin{equation}
696 > D(z)=\frac{k_{B}T}{\xi_{\text{static}}(z)}=\frac{(k_{B}T)^{2}}{\int_{0}^{\infty
697 > }\langle\delta F(z,t)\delta F(z,0)\rangle dt}.%
698 > \end{equation}
699 >
700 > The Z-Constraint method, which fixes the z coordinates of the
701 > molecules with respect to the center of the mass of the system, has
702 > been a method suggested to obtain the forces required for the force
703 > auto-correlation calculation.\cite{Marrink1994} However, simply
704 > resetting the coordinate will move the center of the mass of the
705 > whole system. To avoid this problem, we reset the forces of
706 > z-constrained molecules as well as subtract the total constraint
707 > forces from the rest of the system after the force calculation at
708 > each time step instead of resetting the coordinate.
709 >
710 > After the force calculation, define $G_\alpha$ as
711 > \begin{equation}
712 > G_{\alpha} = \sum_i F_{\alpha i}, \label{oopseEq:zc1}
713 > \end{equation}
714 > where $F_{\alpha i}$ is the force in the z direction of atom $i$ in
715 > z-constrained molecule $\alpha$. The forces of the z constrained
716 > molecule are then set to:
717 > \begin{equation}
718 > F_{\alpha i} = F_{\alpha i} -
719 >    \frac{m_{\alpha i} G_{\alpha}}{\sum_i m_{\alpha i}}.
720 > \end{equation}
721 > Here, $m_{\alpha i}$ is the mass of atom $i$ in the z-constrained
722 > molecule. Having rescaled the forces, the velocities must also be
723 > rescaled to subtract out any center of mass velocity in the z
724 > direction.
725 > \begin{equation}
726 > v_{\alpha i} = v_{\alpha i} -
727 >    \frac{\sum_i m_{\alpha i} v_{\alpha i}}{\sum_i m_{\alpha i}},
728 > \end{equation}
729 > where $v_{\alpha i}$ is the velocity of atom $i$ in the z direction.
730 > Lastly, all of the accumulated z constrained forces must be
731 > subtracted from the system to keep the system center of mass from
732 > drifting.
733 > \begin{equation}
734 > F_{\beta i} = F_{\beta i} - \frac{m_{\beta i} \sum_{\alpha}
735 > G_{\alpha}}
736 >    {\sum_{\beta}\sum_i m_{\beta i}},
737 > \end{equation}
738 > where $\beta$ are all of the unconstrained molecules in the system.
739 > Similarly, the velocities of the unconstrained molecules must also
740 > be scaled.
741 > \begin{equation}
742 > v_{\beta i} = v_{\beta i} + \sum_{\alpha}
743 >    \frac{\sum_i m_{\alpha i} v_{\alpha i}}{\sum_i m_{\alpha i}}.
744 > \end{equation}
745 >
746 > At the very beginning of the simulation, the molecules may not be at
747 > their constrained positions. To move a z-constrained molecule to its
748 > specified position, a simple harmonic potential is used
749 > \begin{equation}
750 > U(t)=\frac{1}{2}k_{\text{Harmonic}}(z(t)-z_{\text{cons}})^{2},%
751 > \end{equation}
752 > where $k_{\text{Harmonic}}$ is the harmonic force constant, $z(t)$
753 > is the current $z$ coordinate of the center of mass of the
754 > constrained molecule, and $z_{\text{cons}}$ is the constrained
755 > position. The harmonic force operating on the z-constrained molecule
756 > at time $t$ can be calculated by
757 > \begin{equation}
758 > F_{z_{\text{Harmonic}}}(t)=-\frac{\partial U(t)}{\partial z(t)}=
759 >    -k_{\text{Harmonic}}(z(t)-z_{\text{cons}}).
760 > \end{equation}
761 >
762 >
763 > \section{\label{methodSection:langevin}Integrators for Langevin Dynamics of Rigid Bodies}
764 >
765 > %\subsection{\label{methodSection:temperature}Temperature Control}
766 >
767 > %\subsection{\label{methodSection:pressureControl}Pressure Control}
768 >
769 > %\section{\label{methodSection:hydrodynamics}Hydrodynamics}
770 >
771 > %applications of langevin dynamics
772 > As an excellent alternative to newtonian dynamics, Langevin
773 > dynamics, which mimics a simple heat bath with stochastic and
774 > dissipative forces, has been applied in a variety of studies. The
775 > stochastic treatment of the solvent enables us to carry out
776 > substantially longer time simulation. Implicit solvent Langevin
777 > dynamics simulation of met-enkephalin not only outperforms explicit
778 > solvent simulation on computation efficiency, but also agrees very
779 > well with explicit solvent simulation on dynamics
780 > properties\cite{Shen2002}. Recently, applying Langevin dynamics with
781 > UNRES model, Liow and his coworkers suggest that protein folding
782 > pathways can be possibly exploited within a reasonable amount of
783 > time\cite{Liwo2005}. The stochastic nature of the Langevin dynamics
784 > also enhances the sampling of the system and increases the
785 > probability of crossing energy barrier\cite{Banerjee2004, Cui2003}.
786 > Combining Langevin dynamics with Kramers's theory, Klimov and
787 > Thirumalai identified the free-energy barrier by studying the
788 > viscosity dependence of the protein folding rates\cite{Klimov1997}.
789 > In order to account for solvent induced interactions missing from
790 > implicit solvent model, Kaya incorporated desolvation free energy
791 > barrier into implicit coarse-grained solvent model in protein
792 > folding/unfolding study and discovered a higher free energy barrier
793 > between the native and denatured states. Because of its stability
794 > against noise, Langevin dynamics is very suitable for studying
795 > remagnetization processes in various
796 > systems\cite{Garcia-Palacios1998,Berkov2002,Denisov2003}. For
797 > instance, the oscillation power spectrum of nanoparticles from
798 > Langevin dynamics simulation has the same peak frequencies for
799 > different wave vectors,which recovers the property of magnetic
800 > excitations in small finite structures\cite{Berkov2005a}. In an
801 > attempt to reduce the computational cost of simulation, multiple
802 > time stepping (MTS) methods have been introduced and have been of
803 > great interest to macromolecule and protein
804 > community\cite{Tuckerman1992}. Relying on the observation that
805 > forces between distant atoms generally demonstrate slower
806 > fluctuations than forces between close atoms, MTS method are
807 > generally implemented by evaluating the slowly fluctuating forces
808 > less frequently than the fast ones. Unfortunately, nonlinear
809 > instability resulting from increasing timestep in MTS simulation
810 > have became a critical obstruction preventing the long time
811 > simulation. Due to the coupling to the heat bath, Langevin dynamics
812 > has been shown to be able to damp out the resonance artifact more
813 > efficiently\cite{Sandu1999}.
814 >
815 > %review rigid body dynamics
816 > Rigid bodies are frequently involved in the modeling of different
817 > areas, from engineering, physics, to chemistry. For example,
818 > missiles and vehicle are usually modeled by rigid bodies.  The
819 > movement of the objects in 3D gaming engine or other physics
820 > simulator is governed by the rigid body dynamics. In molecular
821 > simulation, rigid body is used to simplify the model in
822 > protein-protein docking study\cite{Gray2003}.
823 >
824 > It is very important to develop stable and efficient methods to
825 > integrate the equations of motion of orientational degrees of
826 > freedom. Euler angles are the nature choice to describe the
827 > rotational degrees of freedom. However, due to its singularity, the
828 > numerical integration of corresponding equations of motion is very
829 > inefficient and inaccurate. Although an alternative integrator using
830 > different sets of Euler angles can overcome this
831 > difficulty\cite{Ryckaert1977, Andersen1983}, the computational
832 > penalty and the lost of angular momentum conservation still remain.
833 > In 1977, a singularity free representation utilizing quaternions was
834 > developed by Evans\cite{Evans1977}. Unfortunately, this approach
835 > suffer from the nonseparable Hamiltonian resulted from quaternion
836 > representation, which prevents the symplectic algorithm to be
837 > utilized. Another different approach is to apply holonomic
838 > constraints to the atoms belonging to the rigid
839 > body\cite{Barojas1973}. Each atom moves independently under the
840 > normal forces deriving from potential energy and constraint forces
841 > which are used to guarantee the rigidness. However, due to their
842 > iterative nature, SHAKE and Rattle algorithm converge very slowly
843 > when the number of constraint increases.
844 >
845 > The break through in geometric literature suggests that, in order to
846 > develop a long-term integration scheme, one should preserve the
847 > geometric structure of the flow. Matubayasi and Nakahara developed a
848 > time-reversible integrator for rigid bodies in quaternion
849 > representation. Although it is not symplectic, this integrator still
850 > demonstrates a better long-time energy conservation than traditional
851 > methods because of the time-reversible nature. Extending
852 > Trotter-Suzuki to general system with a flat phase space, Miller and
853 > his colleagues devised an novel symplectic, time-reversible and
854 > volume-preserving integrator in quaternion representation. However,
855 > all of the integrators in quaternion representation suffer from the
856 > computational penalty of constructing a rotation matrix from
857 > quaternions to evolve coordinates and velocities at every time step.
858 > An alternative integration scheme utilizing rotation matrix directly
859 > is RSHAKE\cite{Kol1997}, in which a conjugate momentum to rotation
860 > matrix is introduced to re-formulate the Hamiltonian's equation and
861 > the Hamiltonian is evolved in a constraint manifold by iteratively
862 > satisfying the orthogonality constraint. However, RSHAKE is
863 > inefficient because of the iterative procedure. An extremely
864 > efficient integration scheme in rotation matrix representation,
865 > which also preserves the same structural properties of the
866 > Hamiltonian flow as Miller's integrator, is proposed by Dullweber,
867 > Leimkuhler and McLachlan (DLM)\cite{Dullweber1997}.
868 >
869 > %review langevin/browninan dynamics for arbitrarily shaped rigid body
870 > Combining Langevin or Brownian dynamics with rigid body dynamics,
871 > one can study the slow processes in biomolecular systems. Modeling
872 > the DNA as a chain of rigid spheres beads, which subject to harmonic
873 > potentials as well as excluded volume potentials, Mielke and his
874 > coworkers discover rapid superhelical stress generations from the
875 > stochastic simulation of twin supercoiling DNA with response to
876 > induced torques\cite{Mielke2004}. Membrane fusion is another key
877 > biological process which controls a variety of physiological
878 > functions, such as release of neurotransmitters \textit{etc}. A
879 > typical fusion event happens on the time scale of millisecond, which
880 > is impracticable to study using all atomistic model with newtonian
881 > mechanics. With the help of coarse-grained rigid body model and
882 > stochastic dynamics, the fusion pathways were exploited by many
883 > researchers\cite{Noguchi2001,Noguchi2002,Shillcock2005}. Due to the
884 > difficulty of numerical integration of anisotropy rotation, most of
885 > the rigid body models are simply modeled by sphere, cylinder,
886 > ellipsoid or other regular shapes in stochastic simulations. In an
887 > effort to account for the diffusion anisotropy of the arbitrary
888 > particles, Fernandes and de la Torre improved the original Brownian
889 > dynamics simulation algorithm\cite{Ermak1978,Allison1991} by
890 > incorporating a generalized $6\times6$ diffusion tensor and
891 > introducing a simple rotation evolution scheme consisting of three
892 > consecutive rotations\cite{Fernandes2002}. Unfortunately, unexpected
893 > error and bias are introduced into the system due to the arbitrary
894 > order of applying the noncommuting rotation
895 > operators\cite{Beard2003}. Based on the observation the momentum
896 > relaxation time is much less than the time step, one may ignore the
897 > inertia in Brownian dynamics. However, assumption of the zero
898 > average acceleration is not always true for cooperative motion which
899 > is common in protein motion. An inertial Brownian dynamics (IBD) was
900 > proposed to address this issue by adding an inertial correction
901 > term\cite{Beard2001}. As a complement to IBD which has a lower bound
902 > in time step because of the inertial relaxation time, long-time-step
903 > inertial dynamics (LTID) can be used to investigate the inertial
904 > behavior of the polymer segments in low friction
905 > regime\cite{Beard2001}. LTID can also deal with the rotational
906 > dynamics for nonskew bodies without translation-rotation coupling by
907 > separating the translation and rotation motion and taking advantage
908 > of the analytical solution of hydrodynamics properties. However,
909 > typical nonskew bodies like cylinder and ellipsoid are inadequate to
910 > represent most complex macromolecule assemblies. These intricate
911 > molecules have been represented by a set of beads and their
912 > hydrodynamics properties can be calculated using variant
913 > hydrodynamic interaction tensors.
914 >
915 > The goal of the present work is to develop a Langevin dynamics
916 > algorithm for arbitrary rigid particles by integrating the accurate
917 > estimation of friction tensor from hydrodynamics theory into the
918 > sophisticated rigid body dynamics.
919 >
920 >
921 > \subsection{Friction Tensor}
922 >
923 > For an arbitrary rigid body moves in a fluid, it may experience
924 > friction force $f_r$ or friction torque $\tau _r$ along the opposite
925 > direction of the velocity $v$ or angular velocity $\omega$ at
926 > arbitrary origin $P$,
927 > \begin{equation}
928 > \left( \begin{array}{l}
929 > f_r  \\
930 > \tau _r  \\
931 > \end{array} \right) =  - \left( {\begin{array}{*{20}c}
932 >   {\Xi _{P,t} } & {\Xi _{P,c}^T }  \\
933 >   {\Xi _{P,c} } & {\Xi _{P,r} }  \\
934 > \end{array}} \right)\left( \begin{array}{l}
935 > \nu  \\
936 > \omega  \\
937 > \end{array} \right)
938 > \end{equation}
939 > where $\Xi _{P,t}t$ is the translation friction tensor, $\Xi _{P,r}$
940 > is the rotational friction tensor and $\Xi _{P,c}$ is the
941 > translation-rotation coupling tensor. The procedure of calculating
942 > friction tensor using hydrodynamic tensor and comparison between
943 > bead model and shell model were elaborated by Carrasco \textit{et
944 > al}\cite{Carrasco1999}. An important property of the friction tensor
945 > is that the translational friction tensor is independent of origin
946 > while the rotational and coupling are sensitive to the choice of the
947 > origin \cite{Brenner1967}, which can be described by
948 > \begin{equation}
949 > \begin{array}{c}
950 > \Xi _{P,t}  = \Xi _{O,t}  = \Xi _t  \\
951 > \Xi _{P,c}  = \Xi _{O,c}  - r_{OP}  \times \Xi _t  \\
952 > \Xi _{P,r}  = \Xi _{O,r}  - r_{OP}  \times \Xi _t  \times r_{OP}  + \Xi _{O,c}  \times r_{OP}  - r_{OP}  \times \Xi _{O,c}^T  \\
953 > \end{array}
954 > \end{equation}
955 > Where $O$ is another origin and $r_{OP}$ is the vector joining $O$
956 > and $P$. It is also worthy of mention that both of translational and
957 > rotational frictional tensors are always symmetric. In contrast,
958 > coupling tensor is only symmetric at center of reaction:
959 > \begin{equation}
960 > \Xi _{R,c}  = \Xi _{R,c}^T
961 > \end{equation}
962 > The proper location for applying friction force is the center of
963 > reaction, at which the trace of rotational resistance tensor reaches
964 > minimum.
965 >
966 > \subsection{Rigid body dynamics}
967 >
968 > The Hamiltonian of rigid body can be separated in terms of potential
969 > energy $V(r,A)$ and kinetic energy $T(p,\pi)$,
970 > \[
971 > H = V(r,A) + T(v,\pi )
972 > \]
973 > A second-order symplectic method is now obtained by the composition
974 > of the flow maps,
975 > \[
976 > \varphi _{\Delta t}  = \varphi _{\Delta t/2,V}  \circ \varphi
977 > _{\Delta t,T}  \circ \varphi _{\Delta t/2,V}.
978 > \]
979 > Moreover, $\varphi _{\Delta t/2,V}$ can be divided into two
980 > sub-flows which corresponding to force and torque respectively,
981 > \[
982 > \varphi _{\Delta t/2,V}  = \varphi _{\Delta t/2,F}  \circ \varphi
983 > _{\Delta t/2,\tau }.
984 > \]
985 > Since the associated operators of $\varphi _{\Delta t/2,F} $ and
986 > $\circ \varphi _{\Delta t/2,\tau }$ are commuted, the composition
987 > order inside $\varphi _{\Delta t/2,V}$ does not matter.
988 >
989 > Furthermore, kinetic potential can be separated to translational
990 > kinetic term, $T^t (p)$, and rotational kinetic term, $T^r (\pi )$,
991 > \begin{equation}
992 > T(p,\pi ) =T^t (p) + T^r (\pi ).
993 > \end{equation}
994 > where $ T^t (p) = \frac{1}{2}v^T m v $ and $T^r (\pi )$ is defined
995 > by \ref{introEquation:rotationalKineticRB}. Therefore, the
996 > corresponding flow maps are given by
997 > \[
998 > \varphi _{\Delta t,T}  = \varphi _{\Delta t,T^t }  \circ \varphi
999 > _{\Delta t,T^r }.
1000 > \]
1001 > The free rigid body is an example of Lie-Poisson system with
1002 > Hamiltonian function
1003 > \begin{equation}
1004 > T^r (\pi ) = T_1 ^r (\pi _1 ) + T_2^r (\pi _2 ) + T_3^r (\pi _3 )
1005 > \label{introEquation:rotationalKineticRB}
1006 > \end{equation}
1007 > where $T_i^r (\pi _i ) = \frac{{\pi _i ^2 }}{{2I_i }}$ and
1008 > Lie-Poisson structure matrix,
1009 > \begin{equation}
1010 > J(\pi ) = \left( {\begin{array}{*{20}c}
1011 >   0 & {\pi _3 } & { - \pi _2 }  \\
1012 >   { - \pi _3 } & 0 & {\pi _1 }  \\
1013 >   {\pi _2 } & { - \pi _1 } & 0  \\
1014 > \end{array}} \right)
1015 > \end{equation}
1016 > Thus, the dynamics of free rigid body is governed by
1017 > \begin{equation}
1018 > \frac{d}{{dt}}\pi  = J(\pi )\nabla _\pi  T^r (\pi )
1019 > \end{equation}
1020 > One may notice that each $T_i^r$ in Equation
1021 > \ref{introEquation:rotationalKineticRB} can be solved exactly. For
1022 > instance, the equations of motion due to $T_1^r$ are given by
1023 > \begin{equation}
1024 > \frac{d}{{dt}}\pi  = R_1 \pi ,\frac{d}{{dt}}A = AR_1
1025 > \label{introEqaution:RBMotionSingleTerm}
1026 > \end{equation}
1027 > where
1028 > \[ R_1  = \left( {\begin{array}{*{20}c}
1029 >   0 & 0 & 0  \\
1030 >   0 & 0 & {\pi _1 }  \\
1031 >   0 & { - \pi _1 } & 0  \\
1032 > \end{array}} \right).
1033 > \]
1034 > The solutions of Equation \ref{introEqaution:RBMotionSingleTerm} is
1035 > \[
1036 > \pi (\Delta t) = e^{\Delta tR_1 } \pi (0),A(\Delta t) =
1037 > A(0)e^{\Delta tR_1 }
1038 > \]
1039 > with
1040 > \[
1041 > e^{\Delta tR_1 }  = \left( {\begin{array}{*{20}c}
1042 >   0 & 0 & 0  \\
1043 >   0 & {\cos \theta _1 } & {\sin \theta _1 }  \\
1044 >   0 & { - \sin \theta _1 } & {\cos \theta _1 }  \\
1045 > \end{array}} \right),\theta _1  = \frac{{\pi _1 }}{{I_1 }}\Delta t.
1046 > \]
1047 > To reduce the cost of computing expensive functions in $e^{\Delta
1048 > tR_1 }$, we can use Cayley transformation,
1049 > \[
1050 > e^{\Delta tR_1 }  \approx (1 - \Delta tR_1 )^{ - 1} (1 + \Delta tR_1
1051 > )
1052 > \]
1053 > The flow maps for $T_2^r$ and $T_3^r$ can be found in the same
1054 > manner.
1055 >
1056 > In order to construct a second-order symplectic method, we split the
1057 > angular kinetic Hamiltonian function into five terms
1058 > \[
1059 > T^r (\pi ) = \frac{1}{2}T_1 ^r (\pi _1 ) + \frac{1}{2}T_2^r (\pi _2
1060 > ) + T_3^r (\pi _3 ) + \frac{1}{2}T_2^r (\pi _2 ) + \frac{1}{2}T_1 ^r
1061 > (\pi _1 )
1062 > \].
1063 > Concatenating flows corresponding to these five terms, we can obtain
1064 > the flow map for free rigid body,
1065 > \[
1066 > \varphi _{\Delta t,T^r }  = \varphi _{\Delta t/2,\pi _1 }  \circ
1067 > \varphi _{\Delta t/2,\pi _2 }  \circ \varphi _{\Delta t,\pi _3 }
1068 > \circ \varphi _{\Delta t/2,\pi _2 }  \circ \varphi _{\Delta t/2,\pi
1069 > _1 }.
1070 > \]
1071 >
1072 > The equations of motion corresponding to potential energy and
1073 > kinetic energy are listed in the below table,
1074 > \begin{center}
1075 > \begin{tabular}{|l|l|}
1076 >  \hline
1077 >  % after \\: \hline or \cline{col1-col2} \cline{col3-col4} ...
1078 >  Potential & Kinetic \\
1079 >  $\frac{{dq}}{{dt}} = \frac{p}{m}$ & $\frac{d}{{dt}}q = p$ \\
1080 >  $\frac{d}{{dt}}p =  - \frac{{\partial V}}{{\partial q}}$ & $ \frac{d}{{dt}}p = 0$ \\
1081 >  $\frac{d}{{dt}}Q = 0$ & $ \frac{d}{{dt}}Q = Qskew(I^{ - 1} j)$ \\
1082 >  $ \frac{d}{{dt}}\pi  = \sum\limits_i {\left( {Q^T F_i (r,Q)} \right) \times X_i }$ & $\frac{d}{{dt}}\pi  = \pi  \times I^{ - 1} \pi$\\
1083 >  \hline
1084 > \end{tabular}
1085 > \end{center}
1086 >
1087 > Finally, we obtain the overall symplectic flow maps for free moving
1088 > rigid body
1089 > \begin{align*}
1090 > \varphi _{\Delta t}  = &\varphi _{\Delta t/2,F}  \circ \varphi _{\Delta t/2,\tau } \circ  \\
1091 >  &\varphi _{\Delta t,T^t }  \circ \varphi _{\Delta t/2,\pi _1 }  \circ \varphi _{\Delta t/2,\pi _2 }  \circ \varphi _{\Delta t,\pi _3 }  \circ \varphi _{\Delta t/2,\pi _2 }  \circ \varphi _{\Delta t/2,\pi _1 } \circ  \\
1092 >  &\varphi _{\Delta t/2,\tau }  \circ \varphi _{\Delta t/2,F}  .\\
1093 > \label{introEquation:overallRBFlowMaps}
1094 > \end{align*}
1095 >
1096 > \subsection{Langevin dynamics for rigid particles of arbitrary shape}
1097 >
1098 > Consider a Langevin equation of motions in generalized coordinates
1099 > \begin{equation}
1100 > M_i \dot V_i (t) = F_{s,i} (t) + F_{f,i(t)}  + F_{r,i} (t)
1101 > \label{LDGeneralizedForm}
1102 > \end{equation}
1103 > where $M_i$ is a $6\times6$ generalized diagonal mass (include mass
1104 > and moment of inertial) matrix and $V_i$ is a generalized velocity,
1105 > $V_i = V_i(v_i,\omega _i)$. The right side of Eq.
1106 > (\ref{LDGeneralizedForm}) consists of three generalized forces in
1107 > lab-fixed frame, systematic force $F_{s,i}$, dissipative force
1108 > $F_{f,i}$ and stochastic force $F_{r,i}$. While the evolution of the
1109 > system in Newtownian mechanics typically refers to lab-fixed frame,
1110 > it is also convenient to handle the rotation of rigid body in
1111 > body-fixed frame. Thus the friction and random forces are calculated
1112 > in body-fixed frame and converted back to lab-fixed frame by:
1113 > \[
1114 > \begin{array}{l}
1115 > F_{f,i}^l (t) = A^T F_{f,i}^b (t), \\
1116 > F_{r,i}^l (t) = A^T F_{r,i}^b (t) \\
1117 > \end{array}.
1118 > \]
1119 > Here, the body-fixed friction force $F_{r,i}^b$ is proportional to
1120 > the body-fixed velocity at center of resistance $v_{R,i}^b$ and
1121 > angular velocity $\omega _i$,
1122 > \begin{equation}
1123 > F_{r,i}^b (t) = \left( \begin{array}{l}
1124 > f_{r,i}^b (t) \\
1125 > \tau _{r,i}^b (t) \\
1126 > \end{array} \right) =  - \left( {\begin{array}{*{20}c}
1127 >   {\Xi _{R,t} } & {\Xi _{R,c}^T }  \\
1128 >   {\Xi _{R,c} } & {\Xi _{R,r} }  \\
1129 > \end{array}} \right)\left( \begin{array}{l}
1130 > v_{R,i}^b (t) \\
1131 > \omega _i (t) \\
1132 > \end{array} \right),
1133 > \end{equation}
1134 > while the random force $F_{r,i}^l$ is a Gaussian stochastic variable
1135 > with zero mean and variance
1136 > \begin{equation}
1137 > \left\langle {F_{r,i}^l (t)(F_{r,i}^l (t'))^T } \right\rangle  =
1138 > \left\langle {F_{r,i}^b (t)(F_{r,i}^b (t'))^T } \right\rangle  =
1139 > 2k_B T\Xi _R \delta (t - t').
1140 > \end{equation}
1141 > The equation of motion for $v_i$ can be written as
1142 > \begin{equation}
1143 > m\dot v_i (t) = f_{t,i} (t) = f_{s,i} (t) + f_{f,i}^l (t) +
1144 > f_{r,i}^l (t)
1145 > \end{equation}
1146 > Since the frictional force is applied at the center of resistance
1147 > which generally does not coincide with the center of mass, an extra
1148 > torque is exerted at the center of mass. Thus, the net body-fixed
1149 > frictional torque at the center of mass, $\tau _{n,i}^b (t)$, is
1150 > given by
1151 > \begin{equation}
1152 > \tau _{r,i}^b = \tau _{r,i}^b +r_{MR} \times f_{r,i}^b
1153 > \end{equation}
1154 > where $r_{MR}$ is the vector from the center of mass to the center
1155 > of the resistance. Instead of integrating angular velocity in
1156 > lab-fixed frame, we consider the equation of motion of angular
1157 > momentum in body-fixed frame
1158 > \begin{equation}
1159 > \dot \pi _i (t) = \tau _{t,i} (t) = \tau _{s,i} (t) + \tau _{f,i}^b
1160 > (t) + \tau _{r,i}^b(t)
1161 > \end{equation}
1162 >
1163 > Embedding the friction terms into force and torque, one can
1164 > integrate the langevin equations of motion for rigid body of
1165 > arbitrary shape in a velocity-Verlet style 2-part algorithm, where
1166 > $h= \delta t$:
1167 >
1168 > {\tt part one:}
1169 > \begin{align*}
1170 > v_i (t + h/2) &\leftarrow v_i (t) + \frac{{hf_{t,i}^l (t)}}{{2m_i }} \\
1171 > \pi _i (t + h/2) &\leftarrow \pi _i (t) + \frac{{h\tau _{t,i}^b (t)}}{2} \\
1172 > r_i (t + h) &\leftarrow r_i (t) + hv_i (t + h/2) \\
1173 > A_i (t + h) &\leftarrow rotate\left( {h\pi _i (t + h/2)I^{ - 1} } \right) \\
1174 > \end{align*}
1175 > In this context, the $\mathrm{rotate}$ function is the reversible
1176 > product of five consecutive body-fixed rotations,
1177 > \begin{equation}
1178 > \mathrm{rotate}({\bf a}) = \mathsf{G}_x(a_x / 2) \cdot
1179 > \mathsf{G}_y(a_y / 2) \cdot \mathsf{G}_z(a_z) \cdot \mathsf{G}_y(a_y
1180 > / 2) \cdot \mathsf{G}_x(a_x /2),
1181 > \end{equation}
1182 > where each rotational propagator, $\mathsf{G}_\alpha(\theta)$,
1183 > rotates both the rotation matrix ($\mathsf{A}$) and the body-fixed
1184 > angular momentum ($\pi$) by an angle $\theta$ around body-fixed axis
1185 > $\alpha$,
1186 > \begin{equation}
1187 > \mathsf{G}_\alpha( \theta ) = \left\{
1188 > \begin{array}{lcl}
1189 > \mathsf{A}(t) & \leftarrow & \mathsf{A}(0) \cdot \mathsf{R}_\alpha(\theta)^T, \\
1190 > {\bf j}(t) & \leftarrow & \mathsf{R}_\alpha(\theta) \cdot {\bf
1191 > j}(0).
1192 > \end{array}
1193 > \right.
1194 > \end{equation}
1195 > $\mathsf{R}_\alpha$ is a quadratic approximation to the single-axis
1196 > rotation matrix.  For example, in the small-angle limit, the
1197 > rotation matrix around the body-fixed x-axis can be approximated as
1198 > \begin{equation}
1199 > \mathsf{R}_x(\theta) \approx \left(
1200 > \begin{array}{ccc}
1201 > 1 & 0 & 0 \\
1202 > 0 & \frac{1-\theta^2 / 4}{1 + \theta^2 / 4}  & -\frac{\theta}{1+
1203 > \theta^2 / 4} \\
1204 > 0 & \frac{\theta}{1+ \theta^2 / 4} & \frac{1-\theta^2 / 4}{1 +
1205 > \theta^2 / 4}
1206 > \end{array}
1207 > \right).
1208 > \end{equation}
1209 > All other rotations follow in a straightforward manner.
1210 >
1211 > After the first part of the propagation, the friction and random
1212 > forces are generated at the center of resistance in body-fixed frame
1213 > and converted back into lab-fixed frame
1214 > \[
1215 > f_{t,i}^l (t + h) =  - \left( {\frac{{\partial V}}{{\partial r_i }}}
1216 > \right)_{r_i (t + h)}  + A_i^T (t + h)[F_{f,i}^b (t + h) + F_{r,i}^b
1217 > (t + h)],
1218 > \]
1219 > while the system torque in lab-fixed frame is transformed into
1220 > body-fixed frame,
1221 > \[
1222 > \tau _{t,i}^b (t + h) = A\tau _{s,i}^l (t) + \tau _{n,i}^b (t) +
1223 > \tau _{r,i}^b (t).
1224 > \]
1225 > Once the forces and torques have been obtained at the new time step,
1226 > the velocities can be advanced to the same time value.
1227 >
1228 > {\tt part two:}
1229 > \begin{align*}
1230 > v_i (t) &\leftarrow v_i (t + h/2) + \frac{{hf_{t,i}^l (t + h)}}{{2m_i }} \\
1231 > \pi _i (t) &\leftarrow \pi _i (t + h/2) + \frac{{h\tau _{t,i}^b (t + h)}}{2} \\
1232 > \end{align*}
1233 >
1234 > \subsection{Results and discussion}

Diff Legend

Removed lines
+ Added lines
< Changed lines
> Changed lines