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

Comparing trunk/langevin/langevin.tex (file contents):
Revision 3298 by xsun, Wed Jan 2 21:06:31 2008 UTC vs.
Revision 3310 by gezelter, Fri Jan 11 22:08:57 2008 UTC

# Line 17 | Line 17
17   \evensidemargin 0.0cm \topmargin -21pt \headsep 10pt \textheight
18   9.0in \textwidth 6.5in \brokenpenalty=10000
19   \renewcommand{\baselinestretch}{1.2}
20 < \renewcommand\citemid{\ } % no comma in optional reference note
20 > \renewcommand\citemid{\ } % no comma in optional referenc note
21  
22   \begin{document}
23  
24   \title{An algorithm for performing Langevin dynamics on rigid bodies of arbitrary shape }
25  
26 < \author{Teng Lin, Xiuquan Sun and J. Daniel Gezelter\footnote{Corresponding author. \ Electronic mail:
26 > \author{Xiuquan Sun, Teng Lin and J. Daniel Gezelter\footnote{Corresponding author. \ Electronic mail:
27   gezelter@nd.edu} \\
28   Department of Chemistry and Biochemistry\\
29   University of Notre Dame\\
# Line 171 | Line 171 | into the sophisticated rigid body dynamics algorithms.
171   accurate estimation of friction tensor from hydrodynamics theory
172   into the sophisticated rigid body dynamics algorithms.
173  
174 \section{Computational Methods{\label{methodSec}}}
175
174   \subsection{\label{introSection:frictionTensor}Friction Tensor}
175   Theoretically, the friction kernel can be determined using the
176   velocity autocorrelation function. However, this approach becomes
# Line 367 | Line 365 | arbitrary origin $O$ can be written as
365   \begin{eqnarray}
366   \Xi _{}^{tt}  & = & \sum\limits_i {\sum\limits_j {C_{ij} } } \notag , \\
367   \Xi _{}^{tr}  & = & \Xi _{}^{rt}  = \sum\limits_i {\sum\limits_j {U_i C_{ij} } } , \\
368 < \Xi _{}^{rr}  & = &  - \sum\limits_i {\sum\limits_j {U_i C_{ij} } } U_j. \notag \\
368 > \Xi _{}^{rr}  & = &  - \sum\limits_i {\sum\limits_j {U_i C_{ij} } }
369 > U_j  + 6 \eta V {\bf I}. \notag
370   \label{introEquation:ResistanceTensorArbitraryOrigin}
371   \end{eqnarray}
372 + The final term in the expression for $\Xi^{rr}$ is correction that
373 + accounts for errors in the rotational motion of certain kinds of bead
374 + models. The additive correction uses the solvent viscosity ($\eta$)
375 + as well as the total volume of the beads that contribute to the
376 + hydrodynamic model,
377 + \begin{equation}
378 + V = \frac{4 \pi}{3} \sum_{i=1}^{N} \sigma_i^3,
379 + \end{equation}
380 + where $\sigma_i$ is the radius of bead $i$.  This correction term was
381 + rigorously tested and compared with the analytical results for
382 + two-sphere and ellipsoidal systems by Garcia de la Torre and
383 + Rodes.\cite{Torre:1983lr}
384 +
385 +
386   The resistance tensor depends on the origin to which they refer. The
387   proper location for applying the friction force is the center of
388   resistance (or center of reaction), at which the trace of rotational
# Line 426 | Line 439 | joining center of resistance $R$ and origin $O$.
439   where $x_OR$, $y_OR$, $z_OR$ are the components of the vector
440   joining center of resistance $R$ and origin $O$.
441  
429 \subsection{Langevin Dynamics for Rigid Particles of Arbitrary Shape\label{LDRB}}
442  
443 + \section{Langevin Dynamics for Rigid Particles of Arbitrary Shape\label{LDRB}}
444   Consider the Langevin equations of motion in generalized coordinates
445   \begin{equation}
446   M_i \dot V_i (t) = F_{s,i} (t) + F_{f,i(t)}  + F_{r,i} (t)
# Line 574 | Line 587 | the velocities can be advanced to the same time value.
587      + \frac{h}{2} {\bf \tau}^b(t + h) .
588   \end{align*}
589  
590 < \section{Results and Discussion}
591 <
592 < The Langevin algorithm described in previous section has been
593 < implemented in {\sc oopse}\cite{Meineke2005} and applied to studies
594 < of the static and dynamic properties in several systems.
595 <
596 < \subsection{Temperature Control}
597 <
598 < As shown in Eq.~\ref{randomForce}, random collisions associated with
599 < the solvent's thermal motions is controlled by the external
600 < temperature. The capability to maintain the temperature of the whole
601 < system was usually used to measure the stability and efficiency of
602 < the algorithm. In order to verify the stability of this new
603 < algorithm, a series of simulations are performed on system
604 < consisiting of 256 SSD water molecules with different viscosities.
605 < The initial configuration for the simulations is taken from a 1ns
606 < NVT simulation with a cubic box of 19.7166~\AA. All simulation are
607 < carried out with cutoff radius of 9~\AA and 2 fs time step for 1 ns
608 < with reference temperature at 300~K. The average temperature as a
609 < function of $\eta$ is shown in Table \ref{langevin:viscosity} where
610 < the temperatures range from 303.04~K to 300.47~K for $\eta = 0.01 -
611 < 1$ poise. The better temperature control at higher viscosity can be
612 < explained by the finite size effect and relative slow relaxation
613 < rate at lower viscosity regime.
614 < \begin{table}
615 < \caption{AVERAGE TEMPERATURES FROM LANGEVIN DYNAMICS SIMULATIONS OF
603 < SSD WATER MOLECULES WITH REFERENCE TEMPERATURE AT 300~K.}
604 < \label{langevin:viscosity}
590 > \section{Validating the Method\label{sec:validating}}
591 > In order to validate our Langevin integrator for arbitrarily-shaped
592 > rigid bodies, we implemented the algorithm in {\sc
593 > oopse}\cite{Meineke2005} and  compared the results of this algorithm
594 > with the known
595 > hydrodynamic limiting behavior for a few model systems, and to
596 > microcanonical molecular dynamics simulations for some more
597 > complicated bodies. The model systems and their analytical behavior
598 > (if known) are summarized below. Parameters for the primary particles
599 > comprising our model systems are given in table \ref{tab:parameters},
600 > and a sketch of the arrangement of these primary particles into the
601 > model rigid bodies is shown in figure \ref{fig:models}. In table
602 > \ref{tab:parameters}, $d$ and $l$ are the physical dimensions of
603 > ellipsoidal (Gay-Berne) particles.  For spherical particles, the value
604 > of the Lennard-Jones $\sigma$ parameter is the particle diameter
605 > ($d$).  Gay-Berne ellipsoids have an energy scaling parameter,
606 > $\epsilon^s$, which describes the well depth for two identical
607 > ellipsoids in a {\it side-by-side} configuration.  Additionally, a
608 > well depth aspect ratio, $\epsilon^r = \epsilon^e / \epsilon^s$,
609 > describes the ratio between the well depths in the {\it end-to-end}
610 > and side-by-side configurations.  For spheres, $\epsilon^r \equiv 1$.
611 > Moments of inertia are also required to describe the motion of primary
612 > particles with orientational degrees of freedom.
613 >
614 > \begin{table*}
615 > \begin{minipage}{\linewidth}
616   \begin{center}
617 < \begin{tabular}{lll}
618 <  \hline
619 <  $\eta$ & $\text{T}_{\text{avg}}$ & $\text{T}_{\text{rms}}$ \\
620 <  \hline
621 <  1    & 300.47 & 10.99 \\
622 <  0.1  & 301.19 & 11.136 \\
623 <  0.01 & 303.04 & 11.796 \\
624 <  \hline
617 > \caption{Parameters for the primary particles in use by the rigid body
618 > models in figure \ref{fig:models}.}
619 > \begin{tabular}{lrcccccccc}
620 > \hline
621 > & & & & & & & \multicolumn{3}c{$\overleftrightarrow{\mathsf I}$ (amu \AA$^2$)} \\
622 > & & $d$ (\AA) & $l$ (\AA) & $\epsilon^s$ (kcal/mol) & $\epsilon^r$ &
623 > $m$ (amu) & $I_{xx}$ & $I_{yy}$ & $I_{zz}$ \\ \hline
624 > Sphere   & & 6.5 & $= d$ & 0.8 & 1 & 190 & 802.75 & 802.75  & 802.75  \\
625 > Ellipsoid & & 4.6 & 13.8  & 0.8 & 0.2 & 200 & 2105 & 2105 & 421 \\
626 > Dumbbell &(2 identical spheres) & 6.5 & $= d$ & 0.8 & 1   & 190 & 802.75 & 802.75 & 802.75 \\
627 > Banana  &(3 identical ellipsoids)& 4.2 & 11.2  & 0.8 & 0.2 & 240 & 10000 & 10000 & 0 \\
628 > Lipid: & Spherical Head & 6.5 & $= d$ & 0.185 & 1 & 196 & & & \\
629 >       & Ellipsoidal Tail & 4.6 & 13.8  & 0.8   & 0.2 & 760 & 45000 & 45000 & 9000 \\
630 > Solvent &  & 4.7 & $= d$ & 0.8 & 1   & 72.06 & & & \\
631 > \hline
632   \end{tabular}
633 + \label{tab:parameters}
634   \end{center}
635 < \end{table}
636 <
618 < Another set of calculations were performed to study the efficiency of
619 < temperature control using different temperature coupling schemes.
620 < The starting configuration is cooled to 173~K and evolved using NVE,
621 < NVT, and Langevin dynamic with time step of 2 fs.
622 < Fig.~\ref{langevin:temperature} shows the heating curve obtained as
623 < the systems reach equilibrium. The orange curve in
624 < Fig.~\ref{langevin:temperature} represents the simulation using
625 < Nos\'e-Hoover temperature scaling scheme with thermostat of 5 ps
626 < which gives reasonable tight coupling, while the blue one from
627 < Langevin dynamics with viscosity of 0.1 poise demonstrates a faster
628 < scaling to the desire temperature. When $ \eta = 0$, Langevin dynamics becomes normal
629 < NVE (see orange curve in Fig.~\ref{langevin:temperature}) which
630 < loses the temperature control ability.
635 > \end{minipage}
636 > \end{table*}
637  
638   \begin{figure}
639   \centering
640 < \includegraphics[width=\linewidth]{temperature.pdf}
641 < \caption[Plot of Temperature Fluctuation Versus Time]{Plot of
642 < temperature fluctuation versus time.} \label{langevin:temperature}
640 > \includegraphics[width=3in]{sketch}
641 > \caption[Sketch of the model systems]{A sketch of the model systems
642 > used in evaluating the behavior of the rigid body Langevin
643 > integrator.} \label{fig:models}
644   \end{figure}
645  
646 < \subsection{Langevin Dynamics simulation {\it vs} NVE simulations}
646 > \subsection{Simulation Methodology}
647 > We performed reference microcanonical simulations with explicit
648 > solvents for each of the different model system.  In each case there
649 > was one solute model and 1929 solvent molecules present in the
650 > simulation box.  All simulations were equilibrated using a
651 > constant-pressure and temperature integrator with target values of 300
652 > K for the temperature and 1 atm for pressure.  Following this stage,
653 > further equilibration and sampling was done in a microcanonical
654 > ensemble.  Since the model bodies are typically quite massive, we were
655 > able to use a time step of 25 fs.
656  
657 < To validate our langevin dynamics simulation. We performed several NVE
658 < simulations with explicit solvents for different shaped
659 < molecules. There are one solute molecule and 1929 solvent molecules in
660 < NVE simulation. The parameters are shown in table
661 < \ref{tab:parameters}. The force field between spheres is standard
662 < Lennard-Jones, and ellipsoids interact with other ellipsoids and
663 < spheres with generalized Gay-Berne potential. All simulations are
664 < carried out at 300 K and 1 Atm. The time step is 25 ns, and a
665 < switching function was applied to all potentials to smoothly turn off
666 < the interactions between a range of $22$ and $25$ \AA.  The switching
667 < function was the standard (cubic) function,
657 > The model systems studied used both Lennard-Jones spheres as well as
658 > uniaxial Gay-Berne ellipoids. In its original form, the Gay-Berne
659 > potential was a single site model for the interactions of rigid
660 > ellipsoidal molecules.\cite{Gay81} It can be thought of as a
661 > modification of the Gaussian overlap model originally described by
662 > Berne and Pechukas.\cite{Berne72} The potential is constructed in the
663 > familiar form of the Lennard-Jones function using
664 > orientation-dependent $\sigma$ and $\epsilon$ parameters,
665 > \begin{equation*}
666 > V_{ij}({\mathbf{\hat u}_i}, {\mathbf{\hat u}_j}, {\mathbf{\hat
667 > r}_{ij}}) = 4\epsilon ({\mathbf{\hat u}_i}, {\mathbf{\hat u}_j},
668 > {\mathbf{\hat r}_{ij}})\left[\left(\frac{\sigma_0}{r_{ij}-\sigma({\mathbf{\hat u
669 > }_i},
670 > {\mathbf{\hat u}_j}, {\mathbf{\hat r}_{ij}})+\sigma_0}\right)^{12}
671 > -\left(\frac{\sigma_0}{r_{ij}-\sigma({\mathbf{\hat u}_i}, {\mathbf{\hat u}_j},
672 > {\mathbf{\hat r}_{ij}})+\sigma_0}\right)^6\right]
673 > \label{eq:gb}
674 > \end{equation*}
675 >
676 > The range $(\sigma({\bf \hat{u}}_{i},{\bf \hat{u}}_{j},{\bf
677 > \hat{r}}_{ij}))$, and strength $(\epsilon({\bf \hat{u}}_{i},{\bf
678 > \hat{u}}_{j},{\bf \hat{r}}_{ij}))$ parameters
679 > are dependent on the relative orientations of the two ellipsoids (${\bf
680 > \hat{u}}_{i},{\bf \hat{u}}_{j}$) as well as the direction of the
681 > inter-ellipsoid separation (${\bf \hat{r}}_{ij}$).  The shape and
682 > attractiveness of each ellipsoid is governed by a relatively small set
683 > of parameters: $l$ and $d$ describe the length and width of each
684 > uniaxial ellipsoid, while $\epsilon^s$, which describes the well depth
685 > for two identical ellipsoids in a {\it side-by-side} configuration.
686 > Additionally, a well depth aspect ratio, $\epsilon^r = \epsilon^e /
687 > \epsilon^s$, describes the ratio between the well depths in the {\it
688 > end-to-end} and side-by-side configurations.  Details of the potential
689 > are given elsewhere,\cite{Luckhurst90,Golubkov06,SunGezelter08} and an
690 > excellent overview of the computational methods that can be used to
691 > efficiently compute forces and torques for this potential can be found
692 > in Ref. \citen{Golubkov06}
693 >
694 > For the interaction between nonequivalent uniaxial ellipsoids (or
695 > between spheres and ellipsoids), the spheres are treated as ellipsoids
696 > with an aspect ratio of 1 ($d = l$) and with an well depth ratio
697 > ($\epsilon^r$) of 1 ($\epsilon^e = \epsilon^s$).  The form of the
698 > Gay-Berne potential we are using was generalized by Cleaver {\it et
699 > al.} and is appropriate for dissimilar uniaxial
700 > ellipsoids.\cite{Cleaver96}
701 >
702 > A switching function was applied to all potentials to smoothly turn
703 > off the interactions between a range of $22$ and $25$ \AA.  The
704 > switching function was the standard (cubic) function,
705   \begin{equation}
706   s(r) =
707          \begin{cases}
# Line 660 | Line 713 | We have computed translational diffusion constants for
713          \end{cases}
714   \label{eq:switchingFunc}
715   \end{equation}
716 < We have computed translational diffusion constants for lipid molecules
717 < from the mean-square displacement,
716 >
717 > To measure shear viscosities from our microcanonical simulations, we
718 > used the Einstein form of the pressure correlation function,\cite{hess:209}
719   \begin{equation}
720 < D = \lim_{t\rightarrow \infty} \frac{1}{6 t} \langle {|\left({\bf r}_{i}(t) - {\bf r}_{i}(0) \right)|}^2 \rangle,
721 < \end{equation}
722 < of the solute molecules. Translational diffusion constants for the
723 < different shaped molecules are shown in table
724 < \ref{tab:translation}.  We have also computed orientational correlation
671 < times for different shaped molecules from fits of the second-order
672 < Legendre polynomial correlation function,
720 > \eta = \lim_{t->\infty} \frac{V}{2 k_B T} \frac{d}{dt} \left\langle \left(
721 > \int_{t_0}^{t_0 + t} P_{xz}(t') dt' \right)^2 \right\rangle_{t_0}.
722 > \label{eq:shear}
723 > \end{equation}
724 > A similar form exists for the bulk viscosity
725   \begin{equation}
726 < C_{\ell}(t)  =  \langle P_{\ell}\left({\bf \mu}_{i}(t) \cdot {\bf
727 < \mu}_{i}(0) \right)
726 > \kappa = \lim_{t->\infty} \frac{V}{2 k_B T} \frac{d}{dt} \left\langle \left(
727 > \int_{t_0}^{t_0 + t}
728 > \left(P\left(t'\right)-\left\langle P \right\rangle \right)dt'
729 > \right)^2 \right\rangle_{t_0}.
730   \end{equation}
731 < the results are shown in table \ref{tab:rotation}. We used einstein
732 < format of the pressure correlation function,
731 > Alternatively, the shear viscosity can also be calculated using a
732 > Green-Kubo formula with the off-diagonal pressure tensor correlation function,
733   \begin{equation}
734 < C_{\ell}(t)  =  \langle P_{\ell}\left({\bf \mu}_{i}(t) \cdot {\bf
735 < \mu}_{i}(0) \right)
734 > \eta = \frac{V}{k_B T} \int_0^{\infty} \left\langle P_{xz}(t_0) P_{xz}(t_0
735 > + t) \right\rangle_{t_0} dt,
736   \end{equation}
737 < to estimate the viscosity of the systems from NVE simulations. The
738 < viscosity can also be calculated by Green-Kubo pressure correlaton
739 < function,
737 > although this method converges extremely slowly and is not practical
738 > for obtaining viscosities from molecular dynamics simulations.
739 >
740 > The Langevin dynamics for the different model systems were performed
741 > at the same temperature as the average temperature of the
742 > microcanonical simulations and with a solvent viscosity taken from
743 > Eq. (\ref{eq:shear}) applied to these simulations.  We used 1024
744 > independent solute simulations to obtain statistics on our Langevin
745 > integrator.
746 >
747 > \subsection{Analysis}
748 >
749 > The quantities of interest when comparing the Langevin integrator to
750 > analytic hydrodynamic equations and to molecular dynamics simulations
751 > are typically translational diffusion constants and orientational
752 > relaxation times.  Translational diffusion constants for point
753 > particles are computed easily from the long-time slope of the
754 > mean-square displacement,
755   \begin{equation}
756 < C_{\ell}(t)  =  \langle P_{\ell}\left({\bf \mu}_{i}(t) \cdot {\bf
688 < \mu}_{i}(0) \right)
756 > D = \lim_{t\rightarrow \infty} \frac{1}{6 t} \left\langle {|\left({\bf r}_{i}(t) - {\bf r}_{i}(0) \right)|}^2 \right\rangle,
757   \end{equation}
758 < However, this method converges slowly, and the statistics are not good
759 < enough to give us a very accurate value. The langevin dynamics
760 < simulations for different shaped molecules are performed at the same
761 < conditions as the NVE simulations with viscosity estimated from NVE
762 < simulations. To get better statistics, 1024 non-interacting solute
763 < molecules are put into one simulation box for each langevin
764 < simulation, this is equal to 1024 simulations for single solute
765 < systems. The diffusion constants and rotation relaxation times for
766 < different shaped molecules are shown in table \ref{tab:translation}
767 < and \ref{tab:rotation} to compare to the results calculated from NVE
768 < simulations. The theoretical values for sphere is calculated from the
769 < Stokes-Einstein law, the theoretical values for ellipsoid is
770 < calculated from Perrin's fomula, the theoretical values for dumbbell
771 < is calculated from StinXX and Davis theory. The exact method is
772 < applied to the langevin dynamics simulations for sphere and ellipsoid,
773 < the bead model is applied to the simulation for dumbbell molecule, and
774 < the rough shell model is applied to ellipsoid, dumbbell, banana and
775 < lipid molecules. The results from all the langevin dynamics
776 < simulations, including exact, bead model and rough shell, match the
777 < theoretical values perfectly for all different shaped molecules. This
778 < indicates that our simulation package for langevin dynamics is working
779 < well. The approxiate methods ( bead model and rough shell model) are
780 < accurate enough for the current simulations. The goal of the langevin
781 < dynamics theory is to replace the explicit solvents by the friction
782 < forces. We compared the dynamic properties of different shaped
783 < molecules in langevin dynamics simulations with that in NVE
784 < simulations. The results are reasonable close. Overall, the
785 < translational diffusion constants calculated from langevin dynamics
786 < simulations are very close to the values from the NVE simulation. For
787 < sphere and lipid molecules, the diffusion constants are a little bit
788 < off from the NVE simulation results. One possible reason is that the
789 < calculation of the viscosity is very difficult to be accurate. Another
790 < possible reason is that although we save very frequently during the
791 < NVE simulations and run pretty long time simulations, there is only
792 < one solute molecule in the system which makes the calculation for the
793 < diffusion constant difficult. The sphere molecule behaves as a free
794 < rotor in the solvent, so there is no rotation relaxation time
795 < calculated from NVE simulations. The rotation relaxation time is not
796 < very close to the NVE simulations results. The banana and lipid
797 < molecules match the NVE simulations results pretty well. The mismatch
798 < between langevin dynamics and NVE simulation for ellipsoid is possibly
799 < caused by the slip boundary condition. For dumbbell, the mismatch is
800 < caused by the size of the solvent molecule is pretty large compared to
801 < dumbbell molecule in NVE simulations.
758 > of the solute molecules.  For models in which the translational
759 > diffusion tensor (${\bf D}_{tt}$) has non-degenerate eigenvalues
760 > (i.e. any non-spherically-symmetric rigid body), it is possible to
761 > compute the diffusive behavior for motion parallel to each body-fixed
762 > axis by projecting the displacement of the particle onto the
763 > body-fixed reference frame at $t=0$.  With an isotropic solvent, as we
764 > have used in this study, there are differences between the three
765 > diffusion constants, but these must converge to the same value at
766 > longer times.  Translational diffusion constants for the different
767 > shaped models are shown in table \ref{tab:translation}.
768 >
769 > In general, the three eigenvalues ($D_1, D_2, D_3$) of the rotational
770 > diffusion tensor (${\bf D}_{rr}$) measure the diffusion of an object
771 > {\it around} a particular body-fixed axis and {\it not} the diffusion
772 > of a vector pointing along the axis.  However, these eigenvalues can
773 > be combined to find 5 characteristic rotational relaxation
774 > times,\cite{PhysRev.119.53,Berne90}
775 > \begin{eqnarray}
776 > 1 / \tau_1 & = & 6 D_r + 2 \Delta \\
777 > 1 / \tau_2 & = & 6 D_r - 2 \Delta \\
778 > 1 / \tau_3 & = & 3 (D_r + D_1)  \\
779 > 1 / \tau_4 & = & 3 (D_r + D_2) \\
780 > 1 / \tau_5 & = & 3 (D_r + D_3)
781 > \end{eqnarray}
782 > where
783 > \begin{equation}
784 > D_r = \frac{1}{3} \left(D_1 + D_2 + D_3 \right)
785 > \end{equation}
786 > and
787 > \begin{equation}
788 > \Delta = \left( (D_1 - D_2)^2 + (D_3 - D_1 )(D_3 - D_2)\right)^{1/2}
789 > \end{equation}
790 > Each of these characteristic times can be used to predict the decay of
791 > part of the rotational correlation function when $\ell = 2$,
792 > \begin{equation}
793 > C_2(t)  =  \frac{a^2}{N^2} e^{-t/\tau_1} + \frac{b^2}{N^2} e^{-t/\tau_2}.
794 > \end{equation}
795 > This is the same as the $F^2_{0,0}(t)$ correlation function that
796 > appears in Ref. \citen{Berne90}.  The amplitudes of the two decay
797 > terms are expressed in terms of three dimensionless functions of the
798 > eigenvalues: $a = \sqrt{3} (D_1 - D_2)$, $b = (2D_3 - D_1 - D_2 +
799 > 2\Delta)$, and $N = 2 \sqrt{\Delta b}$.  Similar expressions can be
800 > obtained for other angular momentum correlation
801 > functions.\cite{PhysRev.119.53,Berne90} In all of the model systems we
802 > studied, only one of the amplitudes of the two decay terms was
803 > non-zero, so it was possible to derive a single relaxation time for
804 > each of the hydrodynamic tensors. In many cases, these characteristic
805 > times are averaged and reported in the literature as a single relaxation
806 > time,\cite{Garcia-de-la-Torre:1997qy}
807 > \begin{equation}
808 > 1 / \tau_0 = \frac{1}{5} \sum_{i=1}^5 \tau_{i}^{-1},
809 > \end{equation}
810 > although for the cases reported here, this averaging is not necessary
811 > and only one of the five relaxation times is relevant.
812 >
813 > To test the Langevin integrator's behavior for rotational relaxation,
814 > we have compared the analytical orientational relaxation times (if
815 > they are known) with the general result from the diffusion tensor and
816 > with the results from both the explicitly solvated molecular dynamics
817 > and Langevin simulations.  Relaxation times from simulations (both
818 > microcanonical and Langevin), were computed using Legendre polynomial
819 > correlation functions for a unit vector (${\bf u}$) fixed along one or
820 > more of the body-fixed axes of the model.
821 > \begin{equation}
822 > C_{\ell}(t)  =  \left\langle P_{\ell}\left({\bf u}_{i}(t) \cdot {\bf
823 > u}_{i}(0) \right) \right\rangle
824 > \end{equation}
825 > For simulations in the high-friction limit, orientational correlation
826 > times can then be obtained from exponential fits of this function, or by
827 > integrating,
828 > \begin{equation}
829 > \tau = \ell (\ell + 1) \int_0^{\infty} C_{\ell}(t) dt.
830 > \end{equation}
831 > In lower-friction solvents, the Legendre correlation functions often
832 > exhibit non-exponential decay, and may not be characterized by a
833 > single decay constant.
834 >
835 > In table \ref{tab:rotation} we show the characteristic rotational
836 > relaxation times (based on the diffusion tensor) for each of the model
837 > systems compared with the values obtained via microcanonical and Langevin
838 > simulations.
839 >
840 > \subsection{Spherical particles}
841 > Our model system for spherical particles was a Lennard-Jones sphere of
842 > diameter ($\sigma$) 6.5 \AA\ in a sea of smaller spheres ($\sigma$ =
843 > 4.7 \AA).  The well depth ($\epsilon$) for both particles was set to
844 > an arbitrary value of 0.8 kcal/mol.  
845 >
846 > The Stokes-Einstein behavior of large spherical particles in
847 > hydrodynamic flows is well known, giving translational friction
848 > coefficients of $6 \pi \eta R$ (stick boundary conditions) and
849 > rotational friction coefficients of $8 \pi \eta R^3$.  Recently,
850 > Schmidt and Skinner have computed the behavior of spherical tag
851 > particles in molecular dynamics simulations, and have shown that {\it
852 > slip} boundary conditions ($\Xi_{tt} = 4 \pi \eta R$) may be more
853 > appropriate for molecule-sized spheres embedded in a sea of spherical
854 > solvent particles.\cite{Schmidt:2004fj,Schmidt:2003kx}
855 >
856 > Our simulation results show similar behavior to the behavior observed
857 > by Schmidt and Skinner.  The diffusion constant obtained from our
858 > microcanonical molecular dynamics simulations lies between the slip
859 > and stick boundary condition results obtained via Stokes-Einstein
860 > behavior.  Since the Langevin integrator assumes Stokes-Einstein stick
861 > boundary conditions in calculating the drag and random forces for
862 > spherical particles, our Langevin routine obtains nearly quantitative
863 > agreement with the hydrodynamic results for spherical particles.  One
864 > avenue for improvement of the method would be to compute elements of
865 > $\Xi_{tt}$ assuming behavior intermediate between the two boundary
866 > conditions.
867 >
868 > In the explicit solvent simulations, both our solute and solvent
869 > particles were structureless, exerting no torques upon each other.
870 > Therefore, there are not rotational correlation times available for
871 > this model system.
872  
873 + \subsection{Ellipsoids}
874 + For uniaxial ellipsoids ($a > b = c$), Perrin's formulae for both
875 + translational and rotational diffusion of each of the body-fixed axes
876 + can be combined to give a single translational diffusion
877 + constant,\cite{Berne90}
878 + \begin{equation}
879 + D = \frac{k_B T}{6 \pi \eta a} G(\rho),
880 + \label{Dperrin}
881 + \end{equation}
882 + as well as a single rotational diffusion coefficient,
883 + \begin{equation}
884 + \Theta = \frac{3 k_B T}{16 \pi \eta a^3} \left\{ \frac{(2 - \rho^2)
885 + G(\rho) - 1}{1 - \rho^4} \right\}.
886 + \label{ThetaPerrin}
887 + \end{equation}
888 + In these expressions, $G(\rho)$ is a function of the axial ratio
889 + ($\rho = b / a$), which for prolate ellipsoids, is
890 + \begin{equation}
891 + G(\rho) = (1- \rho^2)^{-1/2} \ln \left\{ \frac{1 + (1 -
892 + \rho^2)^{1/2}}{\rho} \right\}
893 + \label{GPerrin}
894 + \end{equation}
895 + Again, there is some uncertainty about the correct boundary conditions
896 + to use for molecular-scale ellipsoids in a sea of similarly-sized
897 + solvent particles.  Ravichandran and Bagchi found that {\it slip}
898 + boundary conditions most closely resembled the simulation
899 + results,\cite{Ravichandran:1999fk} in agreement with earlier work of
900 + Tang and Evans.\cite{TANG:1993lr}
901 +
902 + Even though there are analytic resistance tensors for ellipsoids, we
903 + constructed a rough-shell model using 2135 beads (each with a diameter
904 + of 0.25 \AA) to approximate the shape of the model ellipsoid.  We
905 + compared the Langevin dynamics from both the simple ellipsoidal
906 + resistance tensor and the rough shell approximation with
907 + microcanonical simulations and the predictions of Perrin.  As in the
908 + case of our spherical model system, the Langevin integrator reproduces
909 + almost exactly the behavior of the Perrin formulae (which is
910 + unsurprising given that the Perrin formulae were used to derive the
911 + drag and random forces applied to the ellipsoid).  We obtain
912 + translational diffusion constants and rotational correlation times
913 + that are within a few percent of the analytic values for both the
914 + exact treatment of the diffusion tensor as well as the rough-shell
915 + model for the ellipsoid.
916 +
917 + The translational diffusion constants from the microcanonical simulations
918 + agree well with the predictions of the Perrin model, although the rotational
919 + correlation times are a factor of 2 shorter than expected from hydrodynamic
920 + theory.  One explanation for the slower rotation
921 + of explicitly-solvated ellipsoids is the possibility that solute-solvent
922 + collisions happen at both ends of the solute whenever the principal
923 + axis of the ellipsoid is turning. In the upper portion of figure
924 + \ref{fig:explanation} we sketch a physical picture of this explanation.
925 + Since our Langevin integrator is providing nearly quantitative agreement with
926 + the Perrin model, it also predicts orientational diffusion for ellipsoids that
927 + exceed explicitly solvated correlation times by a factor of two.
928 +
929 + \subsection{Rigid dumbbells}
930 + Perhaps the only {\it composite} rigid body for which analytic
931 + expressions for the hydrodynamic tensor are available is the
932 + two-sphere dumbbell model.  This model consists of two non-overlapping
933 + spheres held by a rigid bond connecting their centers. There are
934 + competing expressions for the 6x6 resistance tensor for this
935 + model. Equation (\ref{introEquation:oseenTensor}) above gives the
936 + original Oseen tensor, while the second order expression introduced by
937 + Rotne and Prager,\cite{Rotne1969} and improved by Garc\'{i}a de la
938 + Torre and Bloomfield,\cite{Torre1977} is given above as
939 + Eq. (\ref{introEquation:RPTensorNonOverlapped}).  In our case, we use
940 + a model dumbbell in which the two spheres are identical Lennard-Jones
941 + particles ($\sigma$ = 6.5 \AA\ , $\epsilon$ = 0.8 kcal / mol) held at
942 + a distance of 6.532 \AA.
943 +
944 + The theoretical values for the translational diffusion constant of the
945 + dumbbell are calculated from the work of Stimson and Jeffery, who
946 + studied the motion of this system in a flow parallel to the
947 + inter-sphere axis,\cite{Stimson:1926qy} and Davis, who studied the
948 + motion in a flow {\it perpendicular} to the inter-sphere
949 + axis.\cite{Davis:1969uq} We know of no analytic solutions for the {\it
950 + orientational} correlation times for this model system (other than
951 + those derived from the 6 x 6 tensors mentioned above).
952 +
953 + The bead model for this model system comprises the two large spheres
954 + by themselves, while the rough shell approximation used 3368 separate
955 + beads (each with a diameter of 0.25 \AA) to approximate the shape of
956 + the rigid body.  The hydrodynamics tensors computed from both the bead
957 + and rough shell models are remarkably similar.  Computing the initial
958 + hydrodynamic tensor for a rough shell model can be quite expensive (in
959 + this case it requires inverting a 10104 x 10104 matrix), while the
960 + bead model is typically easy to compute (in this case requiring
961 + inversion of a 6 x 6 matrix).  
962 +
963 + \begin{figure}
964 + \centering
965 + \includegraphics[width=2in]{RoughShell}
966 + \caption[Model rigid bodies and their rough shell approximations]{The
967 + model rigid bodies (left column) used to test this algorithm and their
968 + rough-shell approximations (right-column) that were used to compute
969 + the hydrodynamic tensors.  The top two models (ellipsoid and dumbbell)
970 + have analytic solutions and were used to test the rough shell
971 + approximation.  The lower two models (banana and lipid) were compared
972 + with explicitly-solvated molecular dynamics simulations. }
973 + \label{fig:roughShell}
974 + \end{figure}
975 +
976 +
977 + Once the hydrodynamic tensor has been computed, there is no additional
978 + penalty for carrying out a Langevin simulation with either of the two
979 + different hydrodynamics models.  Our naive expectation is that since
980 + the rigid body's surface is roughened under the various shell models,
981 + the diffusion constants will be even farther from the ``slip''
982 + boundary conditions than observed for the bead model (which uses a
983 + Stokes-Einstein model to arrive at the hydrodynamic tensor).  For the
984 + dumbbell, this prediction is correct although all of the Langevin
985 + diffusion constants are within 6\% of the diffusion constant predicted
986 + from the fully solvated system.
987 +
988 + For rotational motion, Langevin integration (and the hydrodynamic tensor)
989 + yields rotational correlation times that are substantially shorter than those
990 + obtained from explicitly-solvated simulations.  It is likely that this is due
991 + to the large size of the explicit solvent spheres, a feature that prevents
992 + the solvent from coming in contact with a substantial fraction of the surface
993 + area of the dumbbell.  Therefore, the explicit solvent only provides drag
994 + over a substantially reduced surface area of this model, while the
995 + hydrodynamic theories utilize the entire surface area for estimating
996 + rotational diffusion.  A sketch of the free volume available in the explicit
997 + solvent simulations is shown in figure \ref{fig:explanation}.
998 +
999 +
1000 + \begin{figure}
1001 + \centering
1002 + \includegraphics[width=6in]{explanation}
1003 + \caption[Explanations of the differences between orientational
1004 + correlation times for explicitly-solvated models and hydrodynamics
1005 + predictions]{Explanations of the differences between orientational
1006 + correlation times for explicitly-solvated models and hydrodynamic
1007 + predictions.   For the ellipsoids (upper figures), rotation of the
1008 + principal axis can involve correlated collisions at both sides of the
1009 + solute.  In the rigid dumbbell model (lower figures), the large size
1010 + of the explicit solvent spheres prevents them from coming in contact
1011 + with a substantial fraction of the surface area of the dumbbell.
1012 + Therefore, the explicit solvent only provides drag over a
1013 + substantially reduced surface area of this model, where the
1014 + hydrodynamic theories utilize the entire surface area for estimating
1015 + rotational diffusion.
1016 + } \label{fig:explanation}
1017 + \end{figure}
1018 +
1019 +
1020 +
1021 + \subsection{Composite banana-shaped molecules}
1022 + Banana-shaped rigid bodies composed of three Gay-Berne ellipsoids have
1023 + been used by Orlandi {\it et al.} to observe mesophases in
1024 + coarse-grained models for bent-core liquid crystalline
1025 + molecules.\cite{Orlandi:2006fk} We have used the same overlapping
1026 + ellipsoids as a way to test the behavior of our algorithm for a
1027 + structure of some interest to the materials science community,
1028 + although since we are interested in capturing only the hydrodynamic
1029 + behavior of this model, we have left out the dipolar interactions of
1030 + the original Orlandi model.
1031 +
1032 + A reference system composed of a single banana rigid body embedded in a
1033 + sea of 1929 solvent particles was created and run under standard
1034 + (microcanonical) molecular dynamics.  The resulting viscosity of this
1035 + mixture was 0.298 centipoise (as estimated using Eq. (\ref{eq:shear})).
1036 + To calculate the hydrodynamic properties of the banana rigid body model,
1037 + we created a rough shell (see Fig.~\ref{fig:roughShell}), in which
1038 + the banana is represented as a ``shell'' made of 3321 identical beads
1039 + (0.25 \AA\  in diameter) distributed on the surface.  Applying the
1040 + procedure described in Sec.~\ref{introEquation:ResistanceTensorArbitraryOrigin}, we
1041 + identified the center of resistance, ${\bf r} = $(0 \AA, 0.81 \AA, 0 \AA), as
1042 + well as the resistance tensor,
1043 + \begin{equation*}
1044 + \Xi =
1045 + \left( {\begin{array}{*{20}c}
1046 + 0.9261 & 0 & 0&0&0.08585&0.2057\\
1047 + 0& 0.9270&-0.007063& 0.08585&0&0\\
1048 + 0&-0.007063&0.7494&0.2057&0&0\\
1049 + 0&0.0858&0.2057& 58.64& 0&0\\0.08585&0&0&0&48.30&3.219&\\0.2057&0&0&0&3.219&10.7373\\\end{array}} \right),
1050 + \end{equation*}
1051 + where the units for translational, translation-rotation coupling and
1052 + rotational tensors are (kcal fs / mol \AA$^2$), (kcal fs / mol \AA\ rad),
1053 + and (kcal fs / mol rad$^2$), respectively.
1054 +
1055 + The Langevin rigid-body integrator (and the hydrodynamic diffusion tensor)
1056 + are essentially quantitative for translational diffusion of this model.  
1057 + Orientational correlation times under the Langevin rigid-body integrator
1058 + are within 11\% of the values obtained from explicit solvent, but these
1059 + models also exhibit some solvent inaccessible surface area in the
1060 + explicitly-solvated case.  
1061 +
1062 + \subsection{Composite sphero-ellipsoids}
1063 + Spherical heads perched on the ends of Gay-Berne ellipsoids have been
1064 + used recently as models for lipid molecules.\cite{SunGezelter08,Ayton01}
1065 +
1066 + MORE DETAILS
1067 +
1068 +
1069 + \subsection{Summary}
1070   According to our simulations, the langevin dynamics is a reliable
1071   theory to apply to replace the explicit solvents, especially for the
1072   translation properties. For large molecules, the rotation properties
# Line 740 | Line 1075 | are also mimiced reasonablly well.
1075   \begin{table*}
1076   \begin{minipage}{\linewidth}
1077   \begin{center}
1078 < \caption{}
1079 < \begin{tabular}{llccccccc}
1080 < \hline
1081 <  & & Sphere & Ellipsoid & Dumbbell(2 spheres) & Banana(3 ellpsoids) &
1082 < Lipid(head) & lipid(tail) & Solvent \\
1083 < \hline
1084 < $d$ (\AA) & & 6.5 & 4.6  & 6.5 &  4.2 & 6.5 & 4.6 & 4.7 \\
1085 < $l$ (\AA) & & $= d$ & 13.8 & $=d$ & 11.2 & $=d$ & 13.8 & 4.7 \\
1086 < $\epsilon^s$ (kcal/mol) & & 0.8 & 0.8 & 0.8 & 0.8 & 0.185 & 0.8 & 0.8 \\
1087 < $\epsilon_r$ (well-depth aspect ratio)& & 1 & 0.2 & 1 & 0.2 & 1 & 0.2 & 1 \\
753 < $m$ (amu) & & 190 & 200 & 190 & 240 & 196 & 760 & 72.06 \\
754 < %$\overleftrightarrow{\mathsf I}$ (amu \AA$^2$) & & & & \\
755 < %\multicolumn{2}c{$I_{xx}$} & 1125 & 45000 & N/A \\
756 < %\multicolumn{2}c{$I_{yy}$} & 1125 & 45000 & N/A \\
757 < %\multicolumn{2}c{$I_{zz}$} &  0 &    9000 & N/A \\
758 < %$\mu$ (Debye) & & varied & 0 & 0 \\
759 < \end{tabular}
760 < \label{tab:parameters}
761 < \end{center}
762 < \end{minipage}
763 < \end{table*}
764 <
765 < \begin{table*}
766 < \begin{minipage}{\linewidth}
767 < \begin{center}
768 < \caption{}
769 < \begin{tabular}{lccccc}
1078 > \caption{Translational diffusion constants (D) for the model systems
1079 > calculated using microcanonical simulations (with explicit solvent),
1080 > theoretical predictions, and Langevin simulations (with implicit solvent).
1081 > Analytical solutions for the exactly-solved hydrodynamics models are
1082 > from Refs. \citen{Einstein05} (sphere), \citen{Perrin1934} and \citen{Perrin1936}
1083 > (ellipsoid), \citen{Stimson:1926qy} and \citen{Davis:1969uq}
1084 > (dumbbell). The other model systems have no known analytic solution.
1085 > All  diffusion constants are reported in units of $10^{-3}$ cm$^2$ / ps (=
1086 > $10^{-4}$ \AA$^2$  / fs). }
1087 > \begin{tabular}{lccccccc}
1088   \hline
1089 < & & & & &Translation \\
1090 < \hline
1091 < & NVE &  & Theoretical & Langevin & \\
774 < \hline
775 < & $\eta$ & D & D & method & D \\
1089 > & \multicolumn{2}c{microcanonical simulation} & & \multicolumn{3}c{Theoretical} & Langevin \\
1090 > \cline{2-3} \cline{5-7}
1091 > model & $\eta$ (centipoise)  & D & & Analytical & method & Hydrodynamics & simulation \\
1092   \hline
1093 < sphere & 3.480159e-03 & 1.643135e-04 & 1.942779e-04 & exact & 1.982283e-04 \\
1094 < ellipsoid & 2.551262e-03 & 2.437492e-04 & 2.335756e-04 & exact & 2.374905e-04 \\
1095 < & 2.551262e-03  & 2.437492e-04 & 2.335756e-04 & rough shell & 2.284088e-04 \\
1096 < dumbell & 2.41276e-03  & 2.129432e-04 & 2.090239e-04 & bead model & 2.148098e-04 \\
1097 < & 2.41276e-03 & 2.129432e-04 & 2.090239e-04 & rough shell & 2.013219e-04 \\
1098 < banana & 2.9846e-03 & 1.527819e-04 &  & rough shell & 1.54807e-04 \\
1099 < lipid & 3.488661e-03 & 0.9562979e-04 &  & rough shell & 1.320987e-04 \\
1093 > sphere    & 0.261  & ?    & & 2.59 & exact       & 2.59 & 2.56 \\
1094 > ellipsoid & 0.255  & 2.44 & & 2.34 & exact       & 2.34 & 2.37 \\
1095 >          & 0.255  & 2.44 & & 2.34 & rough shell & 2.36 & 2.28 \\
1096 > dumbbell  & 0.322  & ?    & & 1.57 & bead model  & 1.57 & 1.57 \\
1097 >          & 0.322  & ?    & & 1.57 & rough shell & ?    & ?    \\
1098 > banana    & 0.298  & 1.53 & &      & rough shell & 1.56 & 1.55 \\
1099 > lipid     & 0.349  & 0.96 & &      & rough shell & 1.33 & 1.32 \\
1100   \end{tabular}
1101   \label{tab:translation}
1102   \end{center}
# Line 790 | Line 1106 | lipid & 3.488661e-03 & 0.9562979e-04 &  & rough shell
1106   \begin{table*}
1107   \begin{minipage}{\linewidth}
1108   \begin{center}
1109 < \caption{}
1110 < \begin{tabular}{lccccc}
1109 > \caption{Orientational relaxation times ($\tau$) for the model systems using
1110 > microcanonical simulation (with explicit solvent), theoretical
1111 > predictions, and Langevin simulations (with implicit solvent). All
1112 > relaxation times are for the rotational correlation function with
1113 > $\ell = 2$ and are reported in units of ps.  The ellipsoidal model has
1114 > an exact solution for the orientational correlation time due to
1115 > Perrin, but the other model systems have no known analytic solution.}
1116 > \begin{tabular}{lccccccc}
1117   \hline
1118 < & & & & &Rotation \\
1119 < \hline
1120 < & NVE &  & Theoretical & Langevin & \\
799 < \hline
800 < & $\eta$ & $\tau_0$ & $\tau_0$ & method & $\tau_0$ \\
1118 > & \multicolumn{2}c{microcanonical simulation} & & \multicolumn{3}c{Theoretical} & Langevin \\
1119 > \cline{2-3} \cline{5-7}
1120 > model & $\eta$ (centipoise) & $\tau$ & & Perrin & method & Hydrodynamic  & simulation \\
1121   \hline
1122 < sphere & 3.480159e-03 &  & 1.208178e+04 & exact & 1.20628e+04 \\
1123 < ellipsoid & 2.551262e-03 & 4.66806e+04 & 2.198986e+04 & exact & 2.21507e+04 \\
1124 < & 2.551262e-03 & 4.66806e+04 & 2.198986e+04 & rough shell & 2.21714e+04 \\
1125 < dumbell & 2.41276e-03 & 1.42974e+04 &  & bead model & 7.12435e+04 \\
1126 < & 2.41276e-03 & 1.42974e+04 &  & rough shell & 7.04765e+04 \\
1127 < banana & 2.9846e-03 & 6.38323e+04 &  & rough shell & 7.0945e+04 \\
1128 < lipid & 3.488661e-03 & 7.79595e+04 &  & rough shell & 7.78886e+04 \\
1122 > sphere    & 0.261  &      & & 9.06 & exact       & 9.06 & 9.11 \\
1123 > ellipsoid & 0.255  & 46.7 & & 22.0 & exact       & 22.0 & 22.2 \\
1124 >          & 0.255  & 46.7 & & 22.0 & rough shell & 22.6 & 22.2 \\
1125 > dumbbell  & 0.322  & 14.0 & &      & bead model  & 52.3 & 52.8 \\
1126 >          & 0.322  & 14.0 & &      & rough shell & ?    & ?    \\
1127 > banana    & 0.298  & 63.8 & &      & rough shell & 70.9 & 70.9 \\
1128 > lipid     & 0.349  & 78.0 & &      & rough shell & 76.9 & 77.9 \\
1129 > \hline
1130   \end{tabular}
1131   \label{tab:rotation}
1132   \end{center}
1133   \end{minipage}
1134   \end{table*}
1135  
1136 < Langevin dynamics simulations are applied to study the formation of
1137 < the ripple phase of lipid membranes. The initial configuration is
1136 > \section{Application: A rigid-body lipid bilayer}
1137 >
1138 > The Langevin dynamics integrator was applied to study the formation of
1139 > corrugated structures emerging from simulations of the coarse grained
1140 > lipid molecular models presented above.  The initial configuration is
1141   taken from our molecular dynamics studies on lipid bilayers with
1142 < lennard-Jones sphere solvents. The solvent molecules are excluded from
1143 < the system, the experimental value of water viscosity is applied to
1144 < mimic the heat bath. Fig. XXX is the snapshot of the stable
1145 < configuration of the system, the ripple structure stayed stable after
1146 < 100 ns run. The efficiency of the simulation is increased by one order
1142 > lennard-Jones sphere solvents. The solvent molecules were excluded
1143 > from the system, and the experimental value for the viscosity of water
1144 > at 20C ($\eta = 1.00$ cp) was used to mimic the hydrodynamic effects
1145 > of the solvent.  The absence of explicit solvent molecules and the
1146 > stability of the integrator allowed us to take timesteps of 50 fs.  A
1147 > total simulation run time of 100 ns was sampled.
1148 > Fig. \ref{fig:bilayer} shows the configuration of the system after 100
1149 > ns, and the ripple structure remains stable during the entire
1150 > trajectory.  Compared with using explicit bead-model solvent
1151 > molecules, the efficiency of the simulation has increased by an order
1152   of magnitude.
1153  
825 \subsection{Langevin Dynamics of Banana Shaped Molecules}
826
827 In order to verify that Langevin dynamics can mimic the dynamics of
828 the systems absent of explicit solvents, we carried out two sets of
829 simulations and compare their dynamic properties.
830 Fig.~\ref{langevin:twoBanana} shows a snapshot of the simulation
831 made of 256 pentane molecules and two banana shaped molecules at
832 273~K. It has an equivalent implicit solvent system containing only
833 two banana shaped molecules with viscosity of 0.289 center poise. To
834 calculate the hydrodynamic properties of the banana shaped molecule,
835 we created a rough shell model (see Fig.~\ref{langevin:roughShell}),
836 in which the banana shaped molecule is represented as a ``shell''
837 made of 2266 small identical beads with size of 0.3 \AA on the
838 surface. Applying the procedure described in
839 Sec.~\ref{introEquation:ResistanceTensorArbitraryOrigin}, we
840 identified the center of resistance at (0 $\rm{\AA}$, 0.7482 $\rm{\AA}$,
841 -0.1988 $\rm{\AA}$), as well as the resistance tensor,
842 \[
843 \left( {\begin{array}{*{20}c}
844 0.9261 & 0 & 0&0&0.08585&0.2057\\
845 0& 0.9270&-0.007063& 0.08585&0&0\\
846 0&-0.007063&0.7494&0.2057&0&0\\
847 0&0.0858&0.2057& 58.64& 0&0\\
848 0.08585&0&0&0&48.30&3.219&\\
849 0.2057&0&0&0&3.219&10.7373\\
850 \end{array}} \right).
851 \]
852 where the units for translational, translation-rotation coupling and rotational tensors are $\frac{kcal \cdot fs}{mol \cdot \rm{\AA}^2}$, $\frac{kcal \cdot fs}{mol \cdot \rm{\AA} \cdot rad}$ and $\frac{kcal \cdot fs}{mol \cdot rad^2}$ respectively.
853 Curves of the velocity auto-correlation functions in
854 Fig.~\ref{langevin:vacf} were shown to match each other very well.
855 However, because of the stochastic nature, simulation using Langevin
856 dynamics was shown to decay slightly faster than MD. In order to
857 study the rotational motion of the molecules, we also calculated the
858 auto-correlation function of the principle axis of the second GB
859 particle, $u$. The discrepancy shown in Fig.~\ref{langevin:uacf} was
860 probably due to the reason that we used the experimental viscosity directly instead of calculating bulk viscosity from simulation.
861
1154   \begin{figure}
1155   \centering
1156 < \includegraphics[width=\linewidth]{roughShell.pdf}
1157 < \caption[Rough shell model for banana shaped molecule]{Rough shell
1158 < model for banana shaped molecule.} \label{langevin:roughShell}
1156 > \includegraphics[width=\linewidth]{bilayer}
1157 > \caption[Snapshot of a bilayer of rigid-body models for lipids]{A
1158 > snapshot of a bilayer composed of rigid-body models for lipid
1159 > molecules evolving using the Langevin integrator described in this
1160 > work.} \label{fig:bilayer}
1161   \end{figure}
1162  
869 \begin{figure}
870 \centering
871 \includegraphics[width=\linewidth]{twoBanana.pdf}
872 \caption[Snapshot from Simulation of Two Banana Shaped Molecules and
873 256 Pentane Molecules]{Snapshot from simulation of two Banana shaped
874 molecules and 256 pentane molecules.} \label{langevin:twoBanana}
875 \end{figure}
876
877 \begin{figure}
878 \centering
879 \includegraphics[width=\linewidth]{vacf.pdf}
880 \caption[Plots of Velocity Auto-correlation Functions]{Velocity
881 auto-correlation functions of NVE (explicit solvent) in blue and
882 Langevin dynamics (implicit solvent) in red.} \label{langevin:vacf}
883 \end{figure}
884
885 \begin{figure}
886 \centering
887 \includegraphics[width=\linewidth]{uacf.pdf}
888 \caption[Auto-correlation functions of the principle axis of the
889 middle GB particle]{Auto-correlation functions of the principle axis
890 of the middle GB particle of NVE (blue) and Langevin dynamics
891 (red).} \label{langevin:uacf}
892 \end{figure}
893
1163   \section{Conclusions}
1164  
1165   We have presented a new Langevin algorithm by incorporating the
1166   hydrodynamics properties of arbitrary shaped molecules into an
1167 < advanced symplectic integration scheme. The temperature control
1168 < ability of this algorithm was demonstrated by a set of simulations
1169 < with different viscosities. It was also shown to have significant
1170 < advantage of producing rapid thermal equilibration over
902 < Nos\'{e}-Hoover method. Further studies in systems involving banana
903 < shaped molecules illustrated that the dynamic properties could be
904 < preserved by using this new algorithm as an implicit solvent model.
1167 > advanced symplectic integration scheme. Further studies in systems
1168 > involving banana shaped molecules illustrated that the dynamic
1169 > properties could be preserved by using this new algorithm as an
1170 > implicit solvent model.
1171  
1172  
1173   \section{Acknowledgments}
# Line 911 | Line 1177 | of Notre Dame.
1177   of Notre Dame.
1178   \newpage
1179  
1180 < \bibliographystyle{jcp2}
1180 > \bibliographystyle{jcp}
1181   \bibliography{langevin}
1182  
1183   \end{document}

Diff Legend

Removed lines
+ Added lines
< Changed lines
> Changed lines