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 2854 by tim, Sun Jun 11 02:53:15 2006 UTC vs.
Revision 2911 by tim, Thu Jun 29 23:56:11 2006 UTC

# Line 2 | Line 2 | In order to mimic the experiments, which are usually p
2  
3   \section{\label{methodSection:rigidBodyIntegrators}Integrators for Rigid Body Motion in Molecular Dynamics}
4  
5 < In order to mimic the experiments, which are usually performed under
5 > In order to mimic 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.
8 > as the canonical and isobaric-isothermal ensembles. In addition to
9 > the standard ensemble, specific ensembles have been developed to
10 > account for the anisotropy between the lateral and normal directions
11 > of membranes. The $NPAT$ ensemble, in which the normal pressure and
12 > the lateral surface area of the membrane are kept constant, and the
13 > $NP\gamma T$ ensemble, in which the normal pressure and the lateral
14 > surface tension are kept constant were proposed to address the
15 > issues.
16  
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 developed a time-reversible integrator for
20 < rigid bodies in quaternion representation. Although it is not
21 < symplectic, this integrator still demonstrates a better long-time
22 < energy conservation than traditional methods because of the
23 < time-reversible nature. Extending Trotter-Suzuki to general system
24 < with a flat phase space, Miller and his colleagues devised an novel
25 < symplectic, time-reversible and volume-preserving integrator in
17 > Integration schemes for the rotational motion of the rigid molecules
18 > in the microcanonical ensemble have been extensively studied over
19 > the last two decades. Matubayasi developed a
20 > time-reversible integrator for rigid bodies in quaternion
21 > representation\cite{Matubayasi1999}. Although it is not symplectic, this integrator still
22 > demonstrates a better long-time energy conservation than Euler angle
23 > methods because of the time-reversible nature. Extending the
24 > Trotter-Suzuki factorization to general system with a flat phase
25 > space, Miller\cite{Miller2002} and his colleagues devised a novel
26 > symplectic, time-reversible and volume-preserving integrator in the
27   quaternion representation, which was shown to be superior to the
28   Matubayasi's time-reversible integrator. However, all of the
29 < integrators in quaternion representation suffer from the
29 > integrators in the quaternion representation suffer from the
30   computational penalty of constructing a rotation matrix from
31   quaternions to evolve coordinates and velocities at every time step.
32 < An alternative integration scheme utilizing rotation matrix directly
33 < proposed by Dullweber, Leimkuhler and McLachlan (DLM) also preserved
34 < the same structural properties of the Hamiltonian flow. In this
35 < section, the integration scheme of DLM method will be reviewed and
36 < extended to other ensembles.
32 > An alternative integration scheme utilizing the rotation matrix
33 > directly proposed by Dullweber, Leimkuhler and McLachlan (DLM) also
34 > preserved the same structural properties of the Hamiltonian
35 > propagator\cite{Dullweber1997}. In this section, the integration
36 > scheme of DLM method will be reviewed and extended to other
37 > ensembles.
38  
39   \subsection{\label{methodSection:DLM}Integrating the Equations of Motion: the
40   DLM method}
# Line 47 | Line 49 | for timesteps of length $h$.
49   \item the error for a single time step is of order $\mathcal{O}\left(h^4\right)$
50   for timesteps of length $h$.
51   \end{enumerate}
50
52   The integration of the equations of motion is carried out in a
53   velocity-Verlet style 2-part algorithm, where $h= \delta t$:
54  
# Line 62 | Line 63 | velocity-Verlet style 2-part algorithm, where $h= \del
63   {\bf j}\left(t + h / 2 \right)  &\leftarrow {\bf j}(t)
64      + \frac{h}{2} {\bf \tau}^b(t), \\
65   %
66 < \mathsf{A}(t + h) &\leftarrow \mathrm{rotate}\left( h {\bf j}
66 > \mathsf{Q}(t + h) &\leftarrow \mathrm{rotate}\left( h {\bf j}
67      (t + h / 2) \cdot \overleftrightarrow{\mathsf{I}}^{-1} \right).
68   \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}
# Line 74 | Line 74 | rotates both the rotation matrix ($\mathsf{A}$) and th
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
77 > rotates both the rotation matrix $\mathsf{Q}$ 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, \\
83 > \mathsf{Q}(t) & \leftarrow & \mathsf{Q}(0) \cdot \mathsf{R}_\alpha(\theta)^T, \\
84   {\bf j}(t) & \leftarrow & \mathsf{R}_\alpha(\theta) \cdot {\bf
85   j}(0).
86   \end{array}
# Line 100 | Line 100 | All other rotations follow in a straightforward manner
100   \end{array}
101   \right).
102   \end{equation}
103 < All other rotations follow in a straightforward manner.
103 > All other rotations follow in a straightforward manner. After the
104 > first part of the propagation, the forces and body-fixed torques are
105 > calculated at the new positions and orientations
106  
105 After the first part of the propagation, the forces and body-fixed
106 torques are calculated at the new positions and orientations
107
107   {\tt doForces:}
108   \begin{align*}
109   {\bf f}(t + h) &\leftarrow
110      - \left(\frac{\partial V}{\partial {\bf r}}\right)_{{\bf r}(t + h)}, \\
111   %
112   {\bf \tau}^{s}(t + h) &\leftarrow {\bf u}(t + h)
113 <    \times \frac{\partial V}{\partial {\bf u}}, \\
113 >    \times (\frac{\partial V}{\partial {\bf u}})_{u(t+h)}, \\
114   %
115 < {\bf \tau}^{b}(t + h) &\leftarrow \mathsf{A}(t + h)
115 > {\bf \tau}^{b}(t + h) &\leftarrow \mathsf{Q}(t + h)
116      \cdot {\bf \tau}^s(t + h).
117   \end{align*}
118 <
119 < ${\bf u}$ will be automatically updated when the rotation matrix
121 < $\mathsf{A}$ is calculated in {\tt moveA}.  Once the forces and
118 > ${\bf u}$ is automatically updated when the rotation matrix
119 > $\mathsf{Q}$ is calculated in {\tt moveA}.  Once the forces and
120   torques have been obtained at the new time step, the velocities can
121   be advanced to the same time value.
122  
# Line 132 | Line 130 | be advanced to the same time value.
130   \right)
131      + \frac{h}{2} {\bf \tau}^b(t + h) .
132   \end{align*}
135
133   The matrix rotations used in the DLM method end up being more costly
134   computationally than the simpler arithmetic quaternion propagation.
135   With the same time step, a 1000-molecule water simulation shows an
136   average 7\% increase in computation time using the DLM method in
137   place of quaternions. This cost is more than justified when
138   comparing the energy conservation of the two methods as illustrated
139 < in Fig.~\ref{methodFig:timestep}.
139 > in Fig.~\ref{methodFig:timestep} where the resulting energy drift at
140 > various time steps for both the DLM and quaternion integration
141 > schemes is compared. All of the 1000 molecule water simulations
142 > started with the same configuration, and the only difference was the
143 > method for handling rotational motion. At time steps of 0.1 and 0.5
144 > fs, both methods for propagating molecule rotation conserve energy
145 > fairly well, with the quaternion method showing a slight energy
146 > drift over time in the 0.5 fs time step simulation. At time steps of
147 > 1 and 2 fs, the energy conservation benefits of the DLM method are
148 > clearly demonstrated. Thus, while maintaining the same degree of
149 > energy conservation, one can take considerably longer time steps,
150 > leading to an overall reduction in computation time.
151  
152   \begin{figure}
153   \centering
# Line 154 | Line 162 | In Fig.~\ref{methodFig:timestep}, the resulting energy
162   \label{methodFig:timestep}
163   \end{figure}
164  
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
165   \subsection{\label{methodSection:NVT}Nos\'{e}-Hoover Thermostatting}
166  
167   The Nos\'e-Hoover equations of motion are given by\cite{Hoover1985}
168   \begin{eqnarray}
169   \dot{{\bf r}} & = & {\bf v}, \\
170   \dot{{\bf v}} & = & \frac{{\bf f}}{m} - \chi {\bf v} ,\\
171 < \dot{\mathsf{A}} & = & \mathsf{A} \cdot
171 > \dot{\mathsf{Q}} & = & \mathsf{Q} \cdot
172   \mbox{ skew}\left(\overleftrightarrow{\mathsf{I}}^{-1} \cdot {\bf j}\right), \\
173   \dot{{\bf j}} & = & {\bf j} \times \left(
174   \overleftrightarrow{\mathsf{I}}^{-1} \cdot {\bf j} \right) - \mbox{
175 < rot}\left(\mathsf{A}^{T} \cdot \frac{\partial V}{\partial
176 < \mathsf{A}} \right) - \chi {\bf j}. \label{eq:nosehoovereom}
175 > rot}\left(\mathsf{Q}^{T} \cdot \frac{\partial V}{\partial
176 > \mathsf{Q}} \right) - \chi {\bf j}. \label{eq:nosehoovereom}
177   \end{eqnarray}
183
178   $\chi$ is an ``extra'' variable included in the extended system, and
179   it is propagated using the first order equation of motion
180   \begin{equation}
181   \dot{\chi} = \frac{1}{\tau_{T}^2} \left(
182   \frac{T}{T_{\mathrm{target}}} - 1 \right). \label{eq:nosehooverext}
183   \end{equation}
184 <
185 < The instantaneous temperature $T$ is proportional to the total
186 < kinetic energy (both translational and orientational) and is given
193 < by
184 > where $\tau_T$ is the time constant for relaxation of the
185 > temperature to the target value, and the instantaneous temperature
186 > $T$ is given by
187   \begin{equation}
188 < T = \frac{2 K}{f k_B}
188 > T = \frac{2 K}{f k_B}.
189   \end{equation}
190   Here, $f$ is the total number of degrees of freedom in the system,
191   \begin{equation}
192   f = 3 N + 3 N_{\mathrm{orient}} - N_{\mathrm{constraints}},
193   \end{equation}
194 < and $K$ is the total kinetic energy,
195 < \begin{equation}
196 < 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}
194 > where $N_{\mathrm{orient}}$ is the number of molecules with
195 > orientational degrees of freedom. The integration of the equations of motion
196 > is carried out in a velocity-Verlet style 2 part algorithm:
197  
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
198   {\tt moveA:}
199   \begin{align*}
200   T(t) &\leftarrow \left\{{\bf v}(t)\right\}, \left\{{\bf j}(t)\right\} ,\\
# Line 229 | Line 210 | T(t) &\leftarrow \left\{{\bf v}(t)\right\}, \left\{{\b
210      + \frac{h}{2} \left( {\bf \tau}^b(t) - {\bf j}(t)
211      \chi(t) \right) ,\\
212   %
213 < \mathsf{A}(t + h) &\leftarrow \mathrm{rotate}
213 > \mathsf{Q}(t + h) &\leftarrow \mathrm{rotate}
214      \left(h * {\bf j}(t + h / 2)
215      \overleftrightarrow{\mathsf{I}}^{-1} \right) ,\\
216   %
# Line 237 | Line 218 | T(t) &\leftarrow \left\{{\bf v}(t)\right\}, \left\{{\b
218      + \frac{h}{2 \tau_T^2} \left( \frac{T(t)}
219      {T_{\mathrm{target}}} - 1 \right) .
220   \end{align*}
240
221   Here $\mathrm{rotate}(h * {\bf j}
222 < \overleftrightarrow{\mathsf{I}}^{-1})$ is the same symplectic
223 < Trotter factorization of the three rotation operations that was
224 < discussed in the section on the DLM integrator.  Note that this
225 < operation modifies both the rotation matrix $\mathsf{A}$ and the
226 < angular momentum ${\bf j}$.  {\tt moveA} propagates velocities by a
227 < half time step, and positional degrees of freedom by a full time
228 < step.  The new positions (and orientations) are then used to
229 < calculate a new set of forces and torques in exactly the same way
230 < they are calculated in the {\tt doForces} portion of the DLM
231 < integrator.
232 <
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
222 > \overleftrightarrow{\mathsf{I}}^{-1})$ is the same symplectic Strang
223 > factorization of the three rotation operations that was discussed in
224 > the section on the DLM integrator.  Note that this operation
225 > modifies both the rotation matrix $\mathsf{Q}$ and the angular
226 > momentum ${\bf j}$.  {\tt moveA} propagates velocities by a half
227 > time step, and positional degrees of freedom by a full time step.
228 > The new positions (and orientations) are then used to calculate a
229 > new set of forces and torques in exactly the same way they are
230 > calculated in the {\tt doForces} portion of the DLM integrator. Once
231 > the forces and torques have been obtained at the new time step, the
232 > temperature, velocities, and the extended system variable can be
233   advanced to the same time value.
234  
235   {\tt moveB:}
# Line 273 | Line 251 | T(t + h) &\leftarrow \left\{{\bf v}(t + h)\right\},
251      \left( {\bf \tau}^b(t + h) - {\bf j}(t + h)
252      \chi(t + h) \right) .
253   \end{align*}
276
254   Since ${\bf v}(t + h)$ and ${\bf j}(t + h)$ are required to
255   caclculate $T(t + h)$ as well as $\chi(t + h)$, they indirectly
256   depend on their own values at time $t + h$.  {\tt moveB} is
257   therefore done in an iterative fashion until $\chi(t + h)$ becomes
258 < self-consistent.
259 <
260 < The Nos\'e-Hoover algorithm is known to conserve a Hamiltonian for
284 < the extended system that is, to within a constant, identical to the
285 < Helmholtz free energy,\cite{Melchionna1993}
258 > self-consistent. The Nos\'e-Hoover algorithm is known to conserve a
259 > Hamiltonian for the extended system that is, to within a constant,
260 > identical to the Helmholtz free energy,\cite{Melchionna1993}
261   \begin{equation}
262   H_{\mathrm{NVT}} = V + K + f k_B T_{\mathrm{target}} \left(
263   \frac{\tau_{T}^2 \chi^2(t)}{2} + \int_{0}^{t} \chi(t^\prime)
264   dt^\prime \right).
265   \end{equation}
266   Poor choices of $h$ or $\tau_T$ can result in non-conservation of
267 < $H_{\mathrm{NVT}}$, so the conserved quantity is maintained in the
268 < last column of the {\tt .stat} file to allow checks on the quality
294 < of the integration.
267 > $H_{\mathrm{NVT}}$, so the conserved quantity should be checked
268 > periodically to verify the quality of the integration.
269  
270   \subsection{\label{methodSection:NPTi}Constant-pressure integration with
271 < isotropic box deformations (NPTi)}
271 > isotropic box (NPTi)}
272  
273 < Isobaric-isothermal ensemble integrator is implemented using the
274 < Melchionna modifications to the Nos\'e-Hoover-Andersen equations of
275 < motion,\cite{Melchionna1993}
302 <
273 > We can used an isobaric-isothermal ensemble integrator which is
274 > implemented using the Melchionna modifications to the
275 > Nos\'e-Hoover-Andersen equations of motion\cite{Melchionna1993}
276   \begin{eqnarray}
277   \dot{{\bf r}} & = & {\bf v} + \eta \left( {\bf r} - {\bf R}_0 \right), \\
278   \dot{{\bf v}} & = & \frac{{\bf f}}{m} - (\eta + \chi) {\bf v}, \\
279 < \dot{\mathsf{A}} & = & \mathsf{A} \cdot
279 > \dot{\mathsf{Q}} & = & \mathsf{Q} \cdot
280   \mbox{ skew}\left(\overleftrightarrow{I}^{-1} \cdot {\bf j}\right),\\
281   \dot{{\bf j}} & = & {\bf j} \times \left(
282   \overleftrightarrow{I}^{-1} \cdot {\bf j} \right) - \mbox{
283 < rot}\left(\mathsf{A}^{T} \cdot \frac{\partial
284 < V}{\partial \mathsf{A}} \right) - \chi {\bf j}, \\
283 > rot}\left(\mathsf{Q}^{T} \cdot \frac{\partial
284 > V}{\partial \mathsf{Q}} \right) - \chi {\bf j}, \\
285   \dot{\chi} & = & \frac{1}{\tau_{T}^2} \left(
286   \frac{T}{T_{\mathrm{target}}} - 1 \right) ,\\
287   \dot{\eta} & = & \frac{1}{\tau_{B}^2 f k_B T_{\mathrm{target}}} V
# Line 316 | Line 289 | P_{\mathrm{target}} \right), \\
289   P_{\mathrm{target}} \right), \\
290   \dot{\mathcal{V}} & = & 3 \mathcal{V} \eta . \label{eq:melchionna1}
291   \end{eqnarray}
319
292   $\chi$ and $\eta$ are the ``extra'' degrees of freedom in the
293   extended system.  $\chi$ is a thermostat, and it has the same
294   function as it does in the Nos\'e-Hoover NVT integrator.  $\eta$ is
# Line 349 | Line 321 | P(t) = \frac{1}{3} \mathrm{Tr} \left(
321   the Pressure tensor,
322   \begin{equation}
323   P(t) = \frac{1}{3} \mathrm{Tr} \left(
324 < \overleftrightarrow{\mathsf{P}}(t). \right)
324 > \overleftrightarrow{\mathsf{P}}(t) \right) .
325   \end{equation}
326 <
355 < In eq.(\ref{eq:melchionna1}), $\tau_B$ is the time constant for
326 > In Eq.~\ref{eq:melchionna1}, $\tau_B$ is the time constant for
327   relaxation of the pressure to the target value. Like in the NVT
328   integrator, the integration of the equations of motion is carried
329   out in a velocity-Verlet style 2 part algorithm:
# Line 371 | Line 342 | P(t) &\leftarrow \left\{{\bf r}(t)\right\}, \left\{{\b
342      + \frac{h}{2} \left( {\bf \tau}^b(t) - {\bf j}(t)
343      \chi(t) \right), \\
344   %
345 < \mathsf{A}(t + h) &\leftarrow \mathrm{rotate}\left(h *
345 > \mathsf{Q}(t + h) &\leftarrow \mathrm{rotate}\left(h *
346      {\bf j}(t + h / 2) \overleftrightarrow{\mathsf{I}}^{-1}
347      \right) ,\\
348   %
# Line 391 | Line 362 | P(t) &\leftarrow \left\{{\bf r}(t)\right\}, \left\{{\b
362   \mathsf{H}(t + h) &\leftarrow e^{-h \eta(t + h / 2)}
363      \mathsf{H}(t).
364   \end{align*}
394
365   Most of these equations are identical to their counterparts in the
366   NVT integrator, but the propagation of positions to time $t + h$
367   depends on the positions at the same time. The simulation box
# Line 403 | Line 373 | box by
373   \mathcal{V}(t + h) \leftarrow e^{ - 3 h \eta(t + h /2)}.
374   \mathcal{V}(t)
375   \end{equation}
406
376   The {\tt doForces} step for the NPTi integrator is exactly the same
377   as in both the DLM and NVT integrators.  Once the forces and torques
378   have been obtained at the new time step, the velocities can be
# Line 435 | Line 404 | P(t + h) &\leftarrow  \left\{{\bf r}(t + h)\right\},
404      \tau}^b(t + h) - {\bf j}(t + h)
405      \chi(t + h) \right) .
406   \end{align*}
438
407   Once again, since ${\bf v}(t + h)$ and ${\bf j}(t + h)$ are required
408   to caclculate $T(t + h)$, $P(t + h)$, $\chi(t + h)$, and $\eta(t +
409   h)$, they indirectly depend on their own values at time $t + h$.
# Line 450 | Line 418 | Poor choices of $\delta t$, $\tau_T$, or $\tau_B$ can
418   \frac{\tau_{T}^2 \chi^2(t)}{2} + \int_{0}^{t} \chi(t^\prime)
419   dt^\prime \right) + P_{\mathrm{target}} \mathcal{V}(t).
420   \end{equation}
421 < Poor choices of $\delta t$, $\tau_T$, or $\tau_B$ can result in
422 < non-conservation of $H_{\mathrm{NPTi}}$, so the conserved quantity
423 < is maintained in the last column of the {\tt .stat} file to allow
456 < checks on the quality of the integration.  It is also known that
457 < this algorithm samples the equilibrium distribution for the enthalpy
458 < (including contributions for the thermostat and barostat),
421 > It is also known that this algorithm samples the equilibrium
422 > distribution for the enthalpy (including contributions for the
423 > thermostat and barostat),
424   \begin{equation}
425   H_{\mathrm{NPTi}} = V + K + \frac{f k_B T_{\mathrm{target}}}{2}
426   \left( \chi^2 \tau_T^2 + \eta^2 \tau_B^2 \right) +
# Line 476 | Line 441 | method are
441   \dot{{\bf r}} & = & {\bf v} + \overleftrightarrow{\eta} \cdot \left( {\bf r} - {\bf R}_0 \right), \\
442   \dot{{\bf v}} & = & \frac{{\bf f}}{m} - (\overleftrightarrow{\eta} +
443   \chi \cdot \mathsf{1}) {\bf v}, \\
444 < \dot{\mathsf{A}} & = & \mathsf{A} \cdot
444 > \dot{\mathsf{Q}} & = & \mathsf{Q} \cdot
445   \mbox{ skew}\left(\overleftrightarrow{I}^{-1} \cdot {\bf j}\right) ,\\
446   \dot{{\bf j}} & = & {\bf j} \times \left(
447   \overleftrightarrow{I}^{-1} \cdot {\bf j} \right) - \mbox{
448 < rot}\left(\mathsf{A}^{T} \cdot \frac{\partial
449 < V}{\partial \mathsf{A}} \right) - \chi {\bf j} ,\\
448 > rot}\left(\mathsf{Q}^{T} \cdot \frac{\partial
449 > V}{\partial \mathsf{Q}} \right) - \chi {\bf j} ,\\
450   \dot{\chi} & = & \frac{1}{\tau_{T}^2} \left(
451   \frac{T}{T_{\mathrm{target}}} - 1 \right) ,\\
452   \dot{\overleftrightarrow{\eta}} & = & \frac{1}{\tau_{B}^2 f k_B
# Line 514 | Line 479 | r}(t)\right\},
479      + \frac{h}{2} \left( {\bf \tau}^b(t) - {\bf j}(t)
480      \chi(t) \right), \\
481   %
482 < \mathsf{A}(t + h) &\leftarrow \mathrm{rotate}\left(h *
482 > \mathsf{Q}(t + h) &\leftarrow \mathrm{rotate}\left(h *
483      {\bf j}(t + h / 2) \overleftrightarrow{\mathsf{I}}^{-1}
484      \right), \\
485   %
# Line 536 | Line 501 | exponential operation is used to scale the simulation
501      \overleftrightarrow{\eta}(t + h / 2)} .
502   \end{align*}
503   Here, a power series expansion truncated at second order for the
504 < exponential operation is used to scale the simulation box.
505 <
506 < The {\tt moveB} portion of the algorithm is largely unchanged from
542 < the NPTi integrator:
504 > exponential operation is used to scale the simulation box. The {\tt
505 > moveB} portion of the algorithm is largely unchanged from the NPTi
506 > integrator:
507  
508   {\tt moveB:}
509   \begin{align*}
# Line 570 | Line 534 | T(t + h) &\leftarrow \left\{{\bf v}(t + h)\right\},
534      + h / 2 \right) + \frac{h}{2} \left( {\bf \tau}^b(t
535      + h) - {\bf j}(t + h) \chi(t + h) \right) .
536   \end{align*}
573
537   The iterative schemes for both {\tt moveA} and {\tt moveB} are
538 < identical to those described for the NPTi integrator.
539 <
577 < The NPTf integrator is known to conserve the following Hamiltonian:
538 > identical to those described for the NPTi integrator. The NPTf
539 > integrator is known to conserve the following Hamiltonian:
540   \begin{eqnarray*}
541   H_{\mathrm{NPTf}} & = & V + K + f k_B T_{\mathrm{target}} \left(
542   \frac{\tau_{T}^2 \chi^2(t)}{2} + \int_{0}^{t} \chi(t^\prime)
543   dt^\prime \right) \\
544 < + P_{\mathrm{target}} \mathcal{V}(t) + \frac{f
545 < k_B T_{\mathrm{target}}}{2}
544 > & & + P_{\mathrm{target}} \mathcal{V}(t) + \frac{f k_B
545 > T_{\mathrm{target}}}{2}
546   \mathrm{Tr}\left[\overleftrightarrow{\eta}(t)\right]^2 \tau_B^2.
547   \end{eqnarray*}
586
548   This integrator must be used with care, particularly in liquid
549   simulations.  Liquids have very small restoring forces in the
550   off-diagonal directions, and the simulation box can very quickly
# Line 592 | Line 553 | assume non-orthorhombic geometries.
553   finds most use in simulating crystals or liquid crystals which
554   assume non-orthorhombic geometries.
555  
556 < \subsubsection{\label{methodSection:NPAT}NPAT Ensemble}
556 > \subsection{\label{methodSection:NPAT}NPAT Ensemble}
557  
558 < A comprehensive understanding of structure¨Cfunction relations of
559 < biological membrane system ultimately relies on structure and
560 < dynamics of lipid bilayer, which are strongly affected by the
561 < interfacial interaction between lipid molecules and surrounding
562 < media. One quantity to describe the interfacial interaction is so
563 < called the average surface area per lipid. Constat area and constant
564 < lateral pressure simulation can be achieved by extending the
565 < standard NPT ensemble with a different pressure control strategy
558 > A comprehensive understanding of relations between structures and
559 > functions in biological membrane system ultimately relies on
560 > structure and dynamics of lipid bilayers, which are strongly
561 > affected by the interfacial interaction between lipid molecules and
562 > surrounding media. One quantity used to describe the interfacial
563 > interaction is the average surface area per lipid.
564 > Constant area and constant lateral pressure simulations can be
565 > achieved by extending the standard NPT ensemble with a different
566 > pressure control strategy
567  
568   \begin{equation}
569   \dot {\overleftrightarrow{\eta}} _{\alpha \beta}=\left\{\begin{array}{ll}
# Line 620 | Line 582 | minimum with respect to surface area $A$
582  
583   Theoretically, the surface tension $\gamma$ of a stress free
584   membrane system should be zero since its surface free energy $G$ is
585 < minimum with respect to surface area $A$
586 < \[
587 < \gamma  = \frac{{\partial G}}{{\partial A}}.
588 < \]
589 < However, a surface tension of zero is not appropriate for relatively
590 < small patches of membrane. In order to eliminate the edge effect of
591 < the membrane simulation, a special ensemble, NP$\gamma$T, is
630 < proposed to maintain the lateral surface tension and normal
631 < pressure. The equation of motion for cell size control tensor,
632 < $\eta$, in $NP\gamma T$ is
585 > minimum with respect to surface area $A$, $\gamma  = \frac{{\partial
586 > G}}{{\partial A}}.$ However, a surface tension of zero is not
587 > appropriate for relatively small patches of membrane. In order to
588 > eliminate the edge effect of membrane simulations, a special
589 > ensemble, NP$\gamma$T, has been proposed to maintain the lateral
590 > surface tension and normal pressure. The equation of motion for the
591 > cell size control tensor, $\eta$, in $NP\gamma T$ is
592   \begin{equation}
593   \dot {\overleftrightarrow{\eta}} _{\alpha \beta}=\left\{\begin{array}{ll}
594      - A_{xy} (\gamma _\alpha   - \gamma _{{\rm{target}}} ) & \mbox{$\alpha  = \beta  = x$ or $\alpha  = \beta  = y$}\\
# Line 645 | Line 604 | - P_{{\rm{target}}} )
604   - P_{{\rm{target}}} )
605   \label{methodEquation:instantaneousSurfaceTensor}
606   \end{equation}
648
607   There is one additional extended system integrator (NPTxyz), in
608   which each attempt to preserve the target pressure along the box
609   walls perpendicular to that particular axis.  The lengths of the box
610   axes are allowed to fluctuate independently, but the angle between
611   the box axes does not change. It should be noted that the NPTxyz
612   integrator is a special case of $NP\gamma T$ if the surface tension
613 < $\gamma$ is set to zero.
613 > $\gamma$ is set to zero, and if $x$ and $y$ can move independently.
614  
615 < \section{\label{methodSection:zcons}Z-Constraint Method}
615 > \section{\label{methodSection:zcons}The Z-Constraint Method}
616  
617   Based on the fluctuation-dissipation theorem, a force
618   auto-correlation method was developed by Roux and Karplus to
# Line 668 | Line 626 | where%
626   \begin{equation}
627   \delta F(z,t)=F(z,t)-\langle F(z,t)\rangle.
628   \end{equation}
671
629   If the time-dependent friction decays rapidly, the static friction
630   coefficient can be approximated by
631   \begin{equation}
# Line 681 | Line 638 | D(z)=\frac{k_{B}T}{\xi_{\text{static}}(z)}=\frac{(k_{B
638   D(z)=\frac{k_{B}T}{\xi_{\text{static}}(z)}=\frac{(k_{B}T)^{2}}{\int_{0}^{\infty
639   }\langle\delta F(z,t)\delta F(z,0)\rangle dt}.%
640   \end{equation}
684
641   The Z-Constraint method, which fixes the z coordinates of the
642   molecules with respect to the center of the mass of the system, has
643   been a method suggested to obtain the forces required for the force
# Line 690 | Line 646 | each time step instead of resetting the coordinate.
646   whole system. To avoid this problem, we reset the forces of
647   z-constrained molecules as well as subtract the total constraint
648   forces from the rest of the system after the force calculation at
649 < each time step instead of resetting the coordinate.
650 <
695 < After the force calculation, define $G_\alpha$ as
649 > each time step instead of resetting the coordinate. After the force
650 > calculation, we define $G_\alpha$ as
651   \begin{equation}
652   G_{\alpha} = \sum_i F_{\alpha i}, \label{oopseEq:zc1}
653   \end{equation}
# Line 727 | Line 682 | v_{\beta i} = v_{\beta i} + \sum_{\alpha}
682   v_{\beta i} = v_{\beta i} + \sum_{\alpha}
683      \frac{\sum_i m_{\alpha i} v_{\alpha i}}{\sum_i m_{\alpha i}}.
684   \end{equation}
730
685   At the very beginning of the simulation, the molecules may not be at
686   their constrained positions. To move a z-constrained molecule to its
687   specified position, a simple harmonic potential is used

Diff Legend

Removed lines
+ Added lines
< Changed lines
> Changed lines