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 2729 by tim, Mon Apr 24 18:49:04 2006 UTC vs.
Revision 2881 by tim, Fri Jun 23 20:21:54 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 ensemble, 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.
7 > methods have been developed to generate statistical ensembles, such
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 and Nakahara developed a time-reversible
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 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.
22 > long-time energy conservation than Euler angle methods because of
23 > the time-reversible nature. Extending the Trotter-Suzuki
24 > factorization to general system with a flat phase space, Miller and
25 > his colleagues devised a novel symplectic, time-reversible and
26 > volume-preserving integrator in the quaternion representation, which
27 > was shown to be superior to the Matubayasi's time-reversible
28 > integrator. However, all of the integrators in the quaternion
29 > representation suffer from the computational penalty of constructing
30 > a rotation matrix from quaternions to evolve coordinates and
31 > velocities at every time step. An alternative integration scheme
32 > utilizing the rotation matrix directly proposed by Dullweber,
33 > Leimkuhler and McLachlan (DLM) also preserved the same structural
34 > properties of the Hamiltonian flow. In this section, the integration
35 > scheme of DLM method will be reviewed and extended to other
36 > ensembles.
37  
38   \subsection{\label{methodSection:DLM}Integrating the Equations of Motion: the
39   DLM method}
# Line 111 | Line 112 | torques are calculated at the new positions and orient
112      - \left(\frac{\partial V}{\partial {\bf r}}\right)_{{\bf r}(t + h)}, \\
113   %
114   {\bf \tau}^{s}(t + h) &\leftarrow {\bf u}(t + h)
115 <    \times \frac{\partial V}{\partial {\bf u}}, \\
115 >    \times (\frac{\partial V}{\partial {\bf u}})_{u(t+h)}, \\
116   %
117   {\bf \tau}^{b}(t + h) &\leftarrow \mathsf{A}(t + h)
118      \cdot {\bf \tau}^s(t + h).
119   \end{align*}
120  
121 < {\sc oopse} automatically updates ${\bf u}$ when the rotation matrix
121 > ${\bf u}$ is automatically updated when the rotation matrix
122   $\mathsf{A}$ is calculated in {\tt moveA}.  Once the forces and
123   torques have been obtained at the new time step, the velocities can
124   be advanced to the same time value.
# Line 139 | Line 140 | in Fig.~\ref{timestep}.
140   average 7\% increase in computation time using the DLM method in
141   place of quaternions. This cost is more than justified when
142   comparing the energy conservation of the two methods as illustrated
143 < in Fig.~\ref{timestep}.
143 > in Fig.~\ref{methodFig:timestep}.
144  
145   \begin{figure}
146   \centering
# Line 150 | Line 151 | from the true energy baseline for clarity.} \label{tim
151   increasing time step. For each time step, the dotted line is total
152   energy using the DLM integrator, and the solid line comes from the
153   quaternion integrator. The larger time step plots are shifted up
154 < from the true energy baseline for clarity.} \label{timestep}
154 > from the true energy baseline for clarity.}
155 > \label{methodFig:timestep}
156   \end{figure}
157  
158 < In Fig.~\ref{timestep}, the resulting energy drift at various time
159 < steps for both the DLM and quaternion integration schemes is
160 < compared. All of the 1000 molecule water simulations started with
161 < the same configuration, and the only difference was the method for
162 < handling rotational motion. At time steps of 0.1 and 0.5 fs, both
163 < methods for propagating molecule rotation conserve energy fairly
164 < well, with the quaternion method showing a slight energy drift over
165 < time in the 0.5 fs time step simulation. At time steps of 1 and 2
166 < fs, the energy conservation benefits of the DLM method are clearly
167 < demonstrated. Thus, while maintaining the same degree of energy
168 < conservation, one can take considerably longer time steps, leading
169 < to an overall reduction in computation time.
158 > In Fig.~\ref{methodFig:timestep}, the resulting energy drift at
159 > various time steps for both the DLM and quaternion integration
160 > schemes is compared. All of the 1000 molecule water simulations
161 > started with the same configuration, and the only difference was the
162 > method for handling rotational motion. At time steps of 0.1 and 0.5
163 > fs, both methods for propagating molecule rotation conserve energy
164 > fairly well, with the quaternion method showing a slight energy
165 > drift over time in the 0.5 fs time step simulation. At time steps of
166 > 1 and 2 fs, the energy conservation benefits of the DLM method are
167 > clearly demonstrated. Thus, while maintaining the same degree of
168 > energy conservation, one can take considerably longer time steps,
169 > leading to an overall reduction in computation time.
170  
171   \subsection{\label{methodSection:NVT}Nos\'{e}-Hoover Thermostatting}
172  
173 < The Nos\'e-Hoover equations of motion are given by\cite{Hoover85}
173 > The Nos\'e-Hoover equations of motion are given by\cite{Hoover1985}
174   \begin{eqnarray}
175   \dot{{\bf r}} & = & {\bf v}, \\
176   \dot{{\bf v}} & = & \frac{{\bf f}}{m} - \chi {\bf v} ,\\
# Line 197 | Line 199 | and $K$ is the total kinetic energy,
199   \begin{equation}
200   f = 3 N + 3 N_{\mathrm{orient}} - N_{\mathrm{constraints}},
201   \end{equation}
202 < and $K$ is the total kinetic energy,
202 > where $N_{\mathrm{orient}}$ is the number of molecules with
203 > orientational degrees of freedom, and $K$ is the total kinetic
204 > energy,
205   \begin{equation}
206   K = \sum_{i=1}^{N} \frac{1}{2} m_i {\bf v}_i^T \cdot {\bf v}_i +
207   \sum_{i=1}^{N_{\mathrm{orient}}}  \frac{1}{2} {\bf j}_i^T \cdot
# Line 205 | Line 209 | relaxation of the temperature to the target value.  To
209   \end{equation}
210  
211   In eq.(\ref{eq:nosehooverext}), $\tau_T$ is the time constant for
212 < relaxation of the temperature to the target value.  To set values
213 < for $\tau_T$ or $T_{\mathrm{target}}$ in a simulation, one would use
214 < the {\tt tauThermostat} and {\tt targetTemperature} keywords in the
211 < {\tt .bass} file.  The units for {\tt tauThermostat} are fs, and the
212 < units for the {\tt targetTemperature} are degrees K.   The
213 < integration of the equations of motion is carried out in a
214 < velocity-Verlet style 2 part algorithm:
212 > relaxation of the temperature to the target value. The integration
213 > of the equations of motion is carried out in a velocity-Verlet style
214 > 2 part algorithm:
215  
216   {\tt moveA:}
217   \begin{align*}
# Line 277 | Line 277 | self-consistent.  The relative tolerance for the self-
277   caclculate $T(t + h)$ as well as $\chi(t + h)$, they indirectly
278   depend on their own values at time $t + h$.  {\tt moveB} is
279   therefore done in an iterative fashion until $\chi(t + h)$ becomes
280 < self-consistent.  The relative tolerance for the self-consistency
281 < check defaults to a value of $\mbox{10}^{-6}$, but {\sc oopse} will
282 < terminate the iteration after 4 loops even if the consistency check
283 < has not been satisfied.
280 > self-consistent.
281  
282   The Nos\'e-Hoover algorithm is known to conserve a Hamiltonian for
283   the extended system that is, to within a constant, identical to the
284 < Helmholtz free energy,\cite{melchionna93}
284 > Helmholtz free energy,\cite{Melchionna1993}
285   \begin{equation}
286   H_{\mathrm{NVT}} = V + K + f k_B T_{\mathrm{target}} \left(
287   \frac{\tau_{T}^2 \chi^2(t)}{2} + \int_{0}^{t} \chi(t^\prime)
# Line 295 | Line 292 | Bond constraints are applied at the end of both the {\
292   last column of the {\tt .stat} file to allow checks on the quality
293   of the integration.
294  
298 Bond constraints are applied at the end of both the {\tt moveA} and
299 {\tt moveB} portions of the algorithm.  Details on the constraint
300 algorithms are given in section \ref{oopseSec:rattle}.
301
295   \subsection{\label{methodSection:NPTi}Constant-pressure integration with
296   isotropic box deformations (NPTi)}
297  
298 < To carry out isobaric-isothermal ensemble calculations {\sc oopse}
299 < implements the Melchionna modifications to the
300 < Nos\'e-Hoover-Andersen equations of motion,\cite{melchionna93}
298 > We can used an isobaric-isothermal ensemble integrator which is
299 > implemented using the Melchionna modifications to the
300 > Nos\'e-Hoover-Andersen equations of motion,\cite{Melchionna1993}
301  
302   \begin{eqnarray}
303   \dot{{\bf r}} & = & {\bf v} + \eta \left( {\bf r} - {\bf R}_0 \right), \\
# Line 359 | Line 352 | relaxation of the pressure to the target value.  To se
352   \end{equation}
353  
354   In eq.(\ref{eq:melchionna1}), $\tau_B$ is the time constant for
355 < relaxation of the pressure to the target value.  To set values for
363 < $\tau_B$ or $P_{\mathrm{target}}$ in a simulation, one would use the
364 < {\tt tauBarostat} and {\tt targetPressure} keywords in the {\tt
365 < .bass} file.  The units for {\tt tauBarostat} are fs, and the units
366 < for the {\tt targetPressure} are atmospheres.  Like in the NVT
355 > relaxation of the pressure to the target value. Like in the NVT
356   integrator, the integration of the equations of motion is carried
357   out in a velocity-Verlet style 2 part algorithm:
358  
# Line 404 | Line 393 | depends on the positions at the same time.  {\sc oopse
393  
394   Most of these equations are identical to their counterparts in the
395   NVT integrator, but the propagation of positions to time $t + h$
396 < depends on the positions at the same time.  {\sc oopse} carries out
397 < this step iteratively (with a limit of 5 passes through the
398 < iterative loop).  Also, the simulation box $\mathsf{H}$ is scaled
399 < uniformly for one full time step by an exponential factor that
400 < depends on the value of $\eta$ at time $t + h / 2$.  Reshaping the
412 < box uniformly also scales the volume of the box by
396 > depends on the positions at the same time. The simulation box
397 > $\mathsf{H}$ is scaled uniformly for one full time step by an
398 > exponential factor that depends on the value of $\eta$ at time $t +
399 > h / 2$.  Reshaping the box uniformly also scales the volume of the
400 > box by
401   \begin{equation}
402   \mathcal{V}(t + h) \leftarrow e^{ - 3 h \eta(t + h /2)}.
403   \mathcal{V}(t)
# Line 451 | Line 439 | + h)$ and $\eta(t + h)$ become self-consistent.  The r
439   to caclculate $T(t + h)$, $P(t + h)$, $\chi(t + h)$, and $\eta(t +
440   h)$, they indirectly depend on their own values at time $t + h$.
441   {\tt moveB} is therefore done in an iterative fashion until $\chi(t
442 < + h)$ and $\eta(t + h)$ become self-consistent.  The relative
455 < tolerance for the self-consistency check defaults to a value of
456 < $\mbox{10}^{-6}$, but {\sc oopse} will terminate the iteration after
457 < 4 loops even if the consistency check has not been satisfied.
442 > + h)$ and $\eta(t + h)$ become self-consistent.
443  
444   The Melchionna modification of the Nos\'e-Hoover-Andersen algorithm
445   is known to conserve a Hamiltonian for the extended system that is,
# Line 476 | Line 461 | Bond constraints are applied at the end of both the {\
461   P_{\mathrm{target}} \mathcal{V}(t).
462   \end{equation}
463  
479 Bond constraints are applied at the end of both the {\tt moveA} and
480 {\tt moveB} portions of the algorithm.  Details on the constraint
481 algorithms are given in section \ref{oopseSec:rattle}.
482
464   \subsection{\label{methodSection:NPTf}Constant-pressure integration with a
465   flexible box (NPTf)}
466  
# Line 553 | Line 534 | r}(t)\right\},
534   \mathsf{H}(t + h) &\leftarrow \mathsf{H}(t) \cdot e^{-h
535      \overleftrightarrow{\eta}(t + h / 2)} .
536   \end{align*}
537 < {\sc oopse} uses a power series expansion truncated at second order
538 < for the exponential operation which scales the simulation box.
537 > Here, a power series expansion truncated at second order for the
538 > exponential operation is used to scale the simulation box.
539  
540   The {\tt moveB} portion of the algorithm is largely unchanged from
541   the NPTi integrator:
# Line 593 | Line 574 | The NPTf integrator is known to conserve the following
574   identical to those described for the NPTi integrator.
575  
576   The NPTf integrator is known to conserve the following Hamiltonian:
577 < \begin{equation}
578 < H_{\mathrm{NPTf}} = V + K + f k_B T_{\mathrm{target}} \left(
577 > \begin{eqnarray*}
578 > H_{\mathrm{NPTf}} & = & V + K + f k_B T_{\mathrm{target}} \left(
579   \frac{\tau_{T}^2 \chi^2(t)}{2} + \int_{0}^{t} \chi(t^\prime)
580 < dt^\prime \right) + P_{\mathrm{target}} \mathcal{V}(t) + \frac{f k_B
580 > dt^\prime \right) \\
581 > & & + P_{\mathrm{target}} \mathcal{V}(t) + \frac{f k_B
582   T_{\mathrm{target}}}{2}
583   \mathrm{Tr}\left[\overleftrightarrow{\eta}(t)\right]^2 \tau_B^2.
584 < \end{equation}
584 > \end{eqnarray*}
585  
586   This integrator must be used with care, particularly in liquid
587   simulations.  Liquids have very small restoring forces in the
# Line 609 | Line 591 | assume non-orthorhombic geometries.
591   finds most use in simulating crystals or liquid crystals which
592   assume non-orthorhombic geometries.
593  
594 < \subsection{\label{methodSection:NPAT}Constant Lateral Pressure and Constant Surface Area (NPAT)}
594 > \subsection{\label{methodSection:NPAT}NPAT Ensemble}
595  
596 < \subsection{\label{methodSection:NPrT}Constant lateral Pressure and Constant Surface Tension (NP\gamma T) }
596 > A comprehensive understanding of relations between structures and
597 > functions in biological membrane system ultimately relies on
598 > structure and dynamics of lipid bilayers, which are strongly
599 > affected by the interfacial interaction between lipid molecules and
600 > surrounding media. One quantity to describe the interfacial
601 > interaction is so called the average surface area per lipid.
602 > Constant area and constant lateral pressure simulations can be
603 > achieved by extending the standard NPT ensemble with a different
604 > pressure control strategy
605  
606 < \subsection{\label{methodSection:NPTxyz}Constant pressure in 3 axes (NPTxyz)}
606 > \begin{equation}
607 > \dot {\overleftrightarrow{\eta}} _{\alpha \beta}=\left\{\begin{array}{ll}
608 >                  \frac{{V(P_{\alpha \beta }  - P_{{\rm{target}}} )}}{{\tau_{\rm{B}}^{\rm{2}} fk_B T_{{\rm{target}}} }}
609 >                  & \mbox{if $ \alpha = \beta  = z$}\\
610 >                  0 & \mbox{otherwise}\\
611 >           \end{array}
612 >    \right.
613 > \end{equation}
614  
615 < There is one additional extended system integrator which is somewhat
616 < simpler than the NPTf method described above.  In this case, the
620 < three axes have independent barostats which each attempt to preserve
621 < the target pressure along the box walls perpendicular to that
622 < particular axis.  The lengths of the box axes are allowed to
623 < fluctuate independently, but the angle between the box axes does not
624 < change. The equations of motion are identical to those described
625 < above, but only the {\it diagonal} elements of
626 < $\overleftrightarrow{\eta}$ are computed.  The off-diagonal elements
627 < are set to zero (even when the pressure tensor has non-zero
628 < off-diagonal elements). It should be noted that the NPTxyz
629 < integrator is a special case of $NP\gamma T$ if the surface tension
630 < $\gamma$ is set to zero.
615 > Note that the iterative schemes for NPAT are identical to those
616 > described for the NPTi integrator.
617  
618 + \subsection{\label{methodSection:NPrT}NP$\gamma$T
619 + Ensemble}
620  
621 < \section{\label{methodSection:constraintMethods}Constraint Methods}
622 <
623 < \subsection{\label{methodSection:rattle}The {\sc rattle} Method for Bond
624 <    Constraints}
621 > Theoretically, the surface tension $\gamma$ of a stress free
622 > membrane system should be zero since its surface free energy $G$ is
623 > minimum with respect to surface area $A$
624 > \[
625 > \gamma  = \frac{{\partial G}}{{\partial A}}.
626 > \]
627 > However, a surface tension of zero is not appropriate for relatively
628 > small patches of membrane. In order to eliminate the edge effect of
629 > the membrane simulation, a special ensemble, NP$\gamma$T, has been
630 > proposed to maintain the lateral surface tension and normal
631 > pressure. The equation of motion for the cell size control tensor,
632 > $\eta$, in $NP\gamma T$ is
633 > \begin{equation}
634 > \dot {\overleftrightarrow{\eta}} _{\alpha \beta}=\left\{\begin{array}{ll}
635 >    - A_{xy} (\gamma _\alpha   - \gamma _{{\rm{target}}} ) & \mbox{$\alpha  = \beta  = x$ or $\alpha  = \beta  = y$}\\
636 >    \frac{{V(P_{\alpha \beta }  - P_{{\rm{target}}} )}}{{\tau _{\rm{B}}^{\rm{2}} fk_B T_{{\rm{target}}}}} & \mbox{$\alpha  = \beta  = z$} \\
637 >    0 & \mbox{$\alpha  \ne \beta$} \\
638 >       \end{array}
639 >    \right.
640 > \end{equation}
641 > where $ \gamma _{{\rm{target}}}$ is the external surface tension and
642 > the instantaneous surface tensor $\gamma _\alpha$ is given by
643 > \begin{equation}
644 > \gamma _\alpha   =  - h_z( \overleftrightarrow{P} _{\alpha \alpha }
645 > - P_{{\rm{target}}} )
646 > \label{methodEquation:instantaneousSurfaceTensor}
647 > \end{equation}
648  
649 < \subsection{\label{methodSection:zcons}Z-Constraint Method}
649 > There is one additional extended system integrator (NPTxyz), in
650 > which each attempt to preserve the target pressure along the box
651 > walls perpendicular to that particular axis.  The lengths of the box
652 > axes are allowed to fluctuate independently, but the angle between
653 > the box axes does not change. It should be noted that the NPTxyz
654 > integrator is a special case of $NP\gamma T$ if the surface tension
655 > $\gamma$ is set to zero, and if $x$ and $y$ can move independently.
656  
657 + \section{\label{methodSection:zcons}The Z-Constraint Method}
658 +
659   Based on the fluctuation-dissipation theorem, a force
660   auto-correlation method was developed by Roux and Karplus to
661 < investigate the dynamics of ions inside ion channels.\cite{Roux91}
661 > investigate the dynamics of ions inside ion channels\cite{Roux1991}.
662   The time-dependent friction coefficient can be calculated from the
663   deviation of the instantaneous force from its mean force.
664   \begin{equation}
# Line 657 | Line 676 | Einstein relation:\cite{Marrink94}
676   F(z,0)\rangle dt.
677   \end{equation}
678   Allowing diffusion constant to then be calculated through the
679 < Einstein relation:\cite{Marrink94}
679 > Einstein relation:\cite{Marrink1994}
680   \begin{equation}
681   D(z)=\frac{k_{B}T}{\xi_{\text{static}}(z)}=\frac{(k_{B}T)^{2}}{\int_{0}^{\infty
682   }\langle\delta F(z,t)\delta F(z,0)\rangle dt}.%
# Line 666 | Line 685 | auto-correlation calculation.\cite{Marrink94} However,
685   The Z-Constraint method, which fixes the z coordinates of the
686   molecules with respect to the center of the mass of the system, has
687   been a method suggested to obtain the forces required for the force
688 < auto-correlation calculation.\cite{Marrink94} However, simply
688 > auto-correlation calculation.\cite{Marrink1994} However, simply
689   resetting the coordinate will move the center of the mass of the
690 < whole system. To avoid this problem, a new method was used in {\sc
672 < oopse}. Instead of resetting the coordinate, we reset the forces of
690 > whole system. To avoid this problem, we reset the forces of
691   z-constrained molecules as well as subtract the total constraint
692   forces from the rest of the system after the force calculation at
693 < each time step.
693 > each time step instead of resetting the coordinate.
694  
695 < After the force calculation, define $G_\alpha$ as
695 > After the force calculation, we define $G_\alpha$ as
696   \begin{equation}
697   G_{\alpha} = \sum_i F_{\alpha i}, \label{oopseEq:zc1}
698   \end{equation}
# Line 725 | Line 743 | F_{z_{\text{Harmonic}}}(t)=-\frac{\partial U(t)}{\part
743   F_{z_{\text{Harmonic}}}(t)=-\frac{\partial U(t)}{\partial z(t)}=
744      -k_{\text{Harmonic}}(z(t)-z_{\text{cons}}).
745   \end{equation}
728
729 \section{\label{methodSection:langevin}Integrators for Langevin Dynamics of Rigid Bodies}
730
731 \subsection{\label{methodSection:temperature}Temperature Control}
732
733 \subsection{\label{methodSection:pressureControl}Pressure Control}
734
735 \section{\label{methodSection:hydrodynamics}Hydrodynamics}
736
737 %\section{\label{methodSection:coarseGrained}Coarse-Grained Modeling}
738
739 %\section{\label{methodSection:moleculeScale}Molecular-Scale Modeling}

Diff Legend

Removed lines
+ Added lines
< Changed lines
> Changed lines