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 2851 by tim, Sun Jun 11 02:06:01 2006 UTC

# Line 16 | Line 16 | two decades. Matubayasi and Nakahara developed a time-
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
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
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
26 > quaternion representation, which was shown to be superior to the
27 > Matubayasi's time-reversible integrator. However, all of the
28 > 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
# Line 139 | Line 139 | in Fig.~\ref{timestep}.
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}.
142 > in Fig.~\ref{methodFig:timestep}.
143  
144   \begin{figure}
145   \centering
# Line 474 | Line 474 | Bond constraints are applied at the end of both the {\
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}.
477 > {\tt moveB} portions of the algorithm.
478  
479   \subsection{\label{methodSection:NPTf}Constant-pressure integration with a
480   flexible box (NPTf)}
# Line 608 | Line 607 | assume non-orthorhombic geometries.
607  
608   \subsection{\label{methodSection:otherSpecialEnsembles}Other Special Ensembles}
609  
610 < \subsubsection{\label{methodSection:NPAT}NPAT Ensemble}
610 > \subsubsection{\label{methodSection:NPAT}\textbf{NPAT Ensemble}}
611  
612   A comprehensive understanding of structure¨Cfunction relations of
613   biological membrane system ultimately relies on structure and
# Line 631 | Line 630 | described for the NPTi integrator.
630   Note that the iterative schemes for NPAT are identical to those
631   described for the NPTi integrator.
632  
633 < \subsubsection{\label{methodSection:NPrT}NP$\gamma$T Ensemble}
633 > \subsubsection{\label{methodSection:NPrT}\textbf{NP$\gamma$T Ensemble}}
634  
635   Theoretically, the surface tension $\gamma$ of a stress free
636   membrane system should be zero since its surface free energy $G$ is
# Line 669 | Line 668 | $\gamma$ is set to zero.
668   integrator is a special case of $NP\gamma T$ if the surface tension
669   $\gamma$ is set to zero.
670  
671 < %\section{\label{methodSection:constraintMethod}Constraint Method}
671 > \section{\label{methodSection:zcons}Z-Constraint Method}
672  
673 < %\subsection{\label{methodSection:bondConstraint}Bond Constraint for Rigid Body}
673 > Based on the fluctuation-dissipation theorem, a force
674 > auto-correlation method was developed by Roux and Karplus to
675 > investigate the dynamics of ions inside ion channels\cite{Roux1991}.
676 > The time-dependent friction coefficient can be calculated from the
677 > deviation of the instantaneous force from its mean force.
678 > \begin{equation}
679 > \xi(z,t)=\langle\delta F(z,t)\delta F(z,0)\rangle/k_{B}T,
680 > \end{equation}
681 > where%
682 > \begin{equation}
683 > \delta F(z,t)=F(z,t)-\langle F(z,t)\rangle.
684 > \end{equation}
685  
686 < %\subsection{\label{methodSection:zcons}Z-constraint Method}
686 > If the time-dependent friction decays rapidly, the static friction
687 > coefficient can be approximated by
688 > \begin{equation}
689 > \xi_{\text{static}}(z)=\int_{0}^{\infty}\langle\delta F(z,t)\delta
690 > F(z,0)\rangle dt.
691 > \end{equation}
692 > Allowing diffusion constant to then be calculated through the
693 > Einstein relation:\cite{Marrink1994}
694 > \begin{equation}
695 > D(z)=\frac{k_{B}T}{\xi_{\text{static}}(z)}=\frac{(k_{B}T)^{2}}{\int_{0}^{\infty
696 > }\langle\delta F(z,t)\delta F(z,0)\rangle dt}.%
697 > \end{equation}
698  
699 < \section{\label{methodSection:langevin}Integrators for Langevin Dynamics of Rigid Bodies}
700 <
701 < \subsection{\label{methodSection:temperature}Temperature Control}
702 <
703 < \subsection{\label{methodSection:pressureControl}Pressure Control}
704 <
705 < \section{\label{methodSection:hydrodynamics}Hydrodynamics}
699 > The Z-Constraint method, which fixes the z coordinates of the
700 > molecules with respect to the center of the mass of the system, has
701 > been a method suggested to obtain the forces required for the force
702 > auto-correlation calculation.\cite{Marrink1994} However, simply
703 > resetting the coordinate will move the center of the mass of the
704 > whole system. To avoid this problem, we reset the forces of
705 > z-constrained molecules as well as subtract the total constraint
706 > forces from the rest of the system after the force calculation at
707 > each time step instead of resetting the coordinate.
708  
709 < %\section{\label{methodSection:coarseGrained}Coarse-Grained Modeling}
709 > After the force calculation, define $G_\alpha$ as
710 > \begin{equation}
711 > G_{\alpha} = \sum_i F_{\alpha i}, \label{oopseEq:zc1}
712 > \end{equation}
713 > where $F_{\alpha i}$ is the force in the z direction of atom $i$ in
714 > z-constrained molecule $\alpha$. The forces of the z constrained
715 > molecule are then set to:
716 > \begin{equation}
717 > F_{\alpha i} = F_{\alpha i} -
718 >    \frac{m_{\alpha i} G_{\alpha}}{\sum_i m_{\alpha i}}.
719 > \end{equation}
720 > Here, $m_{\alpha i}$ is the mass of atom $i$ in the z-constrained
721 > molecule. Having rescaled the forces, the velocities must also be
722 > rescaled to subtract out any center of mass velocity in the z
723 > direction.
724 > \begin{equation}
725 > v_{\alpha i} = v_{\alpha i} -
726 >    \frac{\sum_i m_{\alpha i} v_{\alpha i}}{\sum_i m_{\alpha i}},
727 > \end{equation}
728 > where $v_{\alpha i}$ is the velocity of atom $i$ in the z direction.
729 > Lastly, all of the accumulated z constrained forces must be
730 > subtracted from the system to keep the system center of mass from
731 > drifting.
732 > \begin{equation}
733 > F_{\beta i} = F_{\beta i} - \frac{m_{\beta i} \sum_{\alpha}
734 > G_{\alpha}}
735 >    {\sum_{\beta}\sum_i m_{\beta i}},
736 > \end{equation}
737 > where $\beta$ are all of the unconstrained molecules in the system.
738 > Similarly, the velocities of the unconstrained molecules must also
739 > be scaled.
740 > \begin{equation}
741 > v_{\beta i} = v_{\beta i} + \sum_{\alpha}
742 >    \frac{\sum_i m_{\alpha i} v_{\alpha i}}{\sum_i m_{\alpha i}}.
743 > \end{equation}
744  
745 < %\section{\label{methodSection:moleculeScale}Molecular-Scale Modeling}
745 > At the very beginning of the simulation, the molecules may not be at
746 > their constrained positions. To move a z-constrained molecule to its
747 > specified position, a simple harmonic potential is used
748 > \begin{equation}
749 > U(t)=\frac{1}{2}k_{\text{Harmonic}}(z(t)-z_{\text{cons}})^{2},%
750 > \end{equation}
751 > where $k_{\text{Harmonic}}$ is the harmonic force constant, $z(t)$
752 > is the current $z$ coordinate of the center of mass of the
753 > constrained molecule, and $z_{\text{cons}}$ is the constrained
754 > position. The harmonic force operating on the z-constrained molecule
755 > at time $t$ can be calculated by
756 > \begin{equation}
757 > F_{z_{\text{Harmonic}}}(t)=-\frac{\partial U(t)}{\partial z(t)}=
758 >    -k_{\text{Harmonic}}(z(t)-z_{\text{cons}}).
759 > \end{equation}

Diff Legend

Removed lines
+ Added lines
< Changed lines
> Changed lines