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 2801 by tim, Tue Jun 6 14:56:36 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 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 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 198 | 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 206 | 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
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:
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 278 | 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
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.
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
# Line 299 | Line 295 | To carry out isobaric-isothermal ensemble calculations
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
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}
# Line 356 | 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
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
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 401 | 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
409 < 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 448 | 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
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.
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 473 | Line 461 | Bond constraints are applied at the end of both the {\
461   P_{\mathrm{target}} \mathcal{V}(t).
462   \end{equation}
463  
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
464   \subsection{\label{methodSection:NPTf}Constant-pressure integration with a
465   flexible box (NPTf)}
466  
# Line 550 | 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 590 | 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 606 | 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:otherSpecialEnsembles}Other Special Ensembles}
594 > \subsection{\label{methodSection:NPAT}NPAT Ensemble}
595  
596 < \subsubsection{\label{methodSection:NPAT}NPAT Ensemble}
597 <
598 < A comprehensive understanding of structure¨Cfunction relations of
599 < biological membrane system ultimately relies on structure and
600 < dynamics of lipid bilayer, which are strongly affected by the
601 < interfacial interaction between lipid molecules and surrounding
602 < media. One quantity to describe the interfacial interaction is so
603 < called the average surface area per lipid. Constat area and constant
604 < lateral pressure simulation can be achieved by extending the
620 < standard NPT ensemble with a different pressure control strategy
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   \begin{equation}
607   \dot {\overleftrightarrow{\eta}} _{\alpha \beta}=\left\{\begin{array}{ll}
# Line 631 | Line 615 | described for the NPTi integrator.
615   Note that the iterative schemes for NPAT are identical to those
616   described for the NPTi integrator.
617  
618 < \subsubsection{\label{methodSection:NPrT}NP$\gamma$T Ensemble}
618 > \subsection{\label{methodSection:NPrT}NP$\gamma$T
619 > Ensemble}
620  
621   Theoretically, the surface tension $\gamma$ of a stress free
622   membrane system should be zero since its surface free energy $G$ is
# Line 641 | Line 626 | the membrane simulation, a special ensemble, NP$\gamma
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, is
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 cell size control tensor,
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}
# Line 667 | Line 652 | $\gamma$ is set to zero.
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.
655 > $\gamma$ is set to zero, and if $x$ and $y$ can move independently.
656  
657 < %\section{\label{methodSection:constraintMethod}Constraint Method}
657 > \section{\label{methodSection:zcons}The Z-Constraint Method}
658  
659 < %\subsection{\label{methodSection:bondConstraint}Bond Constraint for Rigid Body}
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{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}
665 > \xi(z,t)=\langle\delta F(z,t)\delta F(z,0)\rangle/k_{B}T,
666 > \end{equation}
667 > where%
668 > \begin{equation}
669 > \delta F(z,t)=F(z,t)-\langle F(z,t)\rangle.
670 > \end{equation}
671  
672 < %\subsection{\label{methodSection:zcons}Z-constraint Method}
672 > If the time-dependent friction decays rapidly, the static friction
673 > coefficient can be approximated by
674 > \begin{equation}
675 > \xi_{\text{static}}(z)=\int_{0}^{\infty}\langle\delta F(z,t)\delta
676 > F(z,0)\rangle dt.
677 > \end{equation}
678 > Allowing diffusion constant to then be calculated through the
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}.%
683 > \end{equation}
684  
685 < \section{\label{methodSection:langevin}Integrators for Langevin Dynamics of Rigid Bodies}
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{Marrink1994} However, simply
689 > resetting the coordinate will move the center of the mass of the
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 instead of resetting the coordinate.
694  
695 < \subsection{\label{methodSection:temperature}Temperature Control}
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}
699 > where $F_{\alpha i}$ is the force in the z direction of atom $i$ in
700 > z-constrained molecule $\alpha$. The forces of the z constrained
701 > molecule are then set to:
702 > \begin{equation}
703 > F_{\alpha i} = F_{\alpha i} -
704 >    \frac{m_{\alpha i} G_{\alpha}}{\sum_i m_{\alpha i}}.
705 > \end{equation}
706 > Here, $m_{\alpha i}$ is the mass of atom $i$ in the z-constrained
707 > molecule. Having rescaled the forces, the velocities must also be
708 > rescaled to subtract out any center of mass velocity in the z
709 > direction.
710 > \begin{equation}
711 > v_{\alpha i} = v_{\alpha i} -
712 >    \frac{\sum_i m_{\alpha i} v_{\alpha i}}{\sum_i m_{\alpha i}},
713 > \end{equation}
714 > where $v_{\alpha i}$ is the velocity of atom $i$ in the z direction.
715 > Lastly, all of the accumulated z constrained forces must be
716 > subtracted from the system to keep the system center of mass from
717 > drifting.
718 > \begin{equation}
719 > F_{\beta i} = F_{\beta i} - \frac{m_{\beta i} \sum_{\alpha}
720 > G_{\alpha}}
721 >    {\sum_{\beta}\sum_i m_{\beta i}},
722 > \end{equation}
723 > where $\beta$ are all of the unconstrained molecules in the system.
724 > Similarly, the velocities of the unconstrained molecules must also
725 > be scaled.
726 > \begin{equation}
727 > v_{\beta i} = v_{\beta i} + \sum_{\alpha}
728 >    \frac{\sum_i m_{\alpha i} v_{\alpha i}}{\sum_i m_{\alpha i}}.
729 > \end{equation}
730  
731 < \subsection{\label{methodSection:pressureControl}Pressure Control}
732 <
733 < \section{\label{methodSection:hydrodynamics}Hydrodynamics}
734 <
735 < %\section{\label{methodSection:coarseGrained}Coarse-Grained Modeling}
736 <
737 < %\section{\label{methodSection:moleculeScale}Molecular-Scale Modeling}
731 > At the very beginning of the simulation, the molecules may not be at
732 > their constrained positions. To move a z-constrained molecule to its
733 > specified position, a simple harmonic potential is used
734 > \begin{equation}
735 > U(t)=\frac{1}{2}k_{\text{Harmonic}}(z(t)-z_{\text{cons}})^{2},%
736 > \end{equation}
737 > where $k_{\text{Harmonic}}$ is the harmonic force constant, $z(t)$
738 > is the current $z$ coordinate of the center of mass of the
739 > constrained molecule, and $z_{\text{cons}}$ is the constrained
740 > position. The harmonic force operating on the z-constrained molecule
741 > at time $t$ can be calculated by
742 > \begin{equation}
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}

Diff Legend

Removed lines
+ Added lines
< Changed lines
> Changed lines