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 3309 by xsun, Fri Jan 11 17:09:08 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 574 | Line 574 | the velocities can be advanced to the same time value.
574      + \frac{h}{2} {\bf \tau}^b(t + h) .
575   \end{align*}
576  
577 < \section{Results and Discussion}
578 <
579 < The Langevin algorithm described in previous section has been
580 < implemented in {\sc oopse}\cite{Meineke2005} and applied to studies
581 < of the static and dynamic properties in several systems.
582 <
583 < \subsection{Temperature Control}
584 <
585 < As shown in Eq.~\ref{randomForce}, random collisions associated with
586 < the solvent's thermal motions is controlled by the external
587 < temperature. The capability to maintain the temperature of the whole
588 < system was usually used to measure the stability and efficiency of
589 < the algorithm. In order to verify the stability of this new
590 < algorithm, a series of simulations are performed on system
591 < consisiting of 256 SSD water molecules with different viscosities.
592 < The initial configuration for the simulations is taken from a 1ns
593 < NVT simulation with a cubic box of 19.7166~\AA. All simulation are
594 < carried out with cutoff radius of 9~\AA and 2 fs time step for 1 ns
595 < with reference temperature at 300~K. The average temperature as a
596 < function of $\eta$ is shown in Table \ref{langevin:viscosity} where
597 < the temperatures range from 303.04~K to 300.47~K for $\eta = 0.01 -
598 < 1$ poise. The better temperature control at higher viscosity can be
599 < explained by the finite size effect and relative slow relaxation
600 < rate at lower viscosity regime.
601 < \begin{table}
602 < \caption{AVERAGE TEMPERATURES FROM LANGEVIN DYNAMICS SIMULATIONS OF
603 < SSD WATER MOLECULES WITH REFERENCE TEMPERATURE AT 300~K.}
604 < \label{langevin:viscosity}
605 < \begin{center}
606 < \begin{tabular}{lll}
607 <  \hline
608 <  $\eta$ & $\text{T}_{\text{avg}}$ & $\text{T}_{\text{rms}}$ \\
609 <  \hline
610 <  1    & 300.47 & 10.99 \\
611 <  0.1  & 301.19 & 11.136 \\
612 <  0.01 & 303.04 & 11.796 \\
613 <  \hline
577 > \section{Results}
578 > In order to validate our Langevin integrator for arbitrarily-shaped
579 > rigid bodies, we implemented the algorithm in {\sc
580 > oopse}\cite{Meineke2005} and  compared the results of this algorithm
581 > with the known
582 > hydrodynamic limiting behavior for a few model systems, and to
583 > microcanonical molecular dynamics simulations for some more
584 > complicated bodies. The model systems and their analytical behavior
585 > (if known) are summarized below. Parameters for the primary particles
586 > comprising our model systems are given in table \ref{tab:parameters},
587 > and a sketch of the arrangement of these primary particles into the
588 > model rigid bodies is shown in figure \ref{fig:models}. In table
589 > \ref{tab:parameters}, $d$ and $l$ are the physical dimensions of
590 > ellipsoidal (Gay-Berne) particles.  For spherical particles, the value
591 > of the Lennard-Jones $\sigma$ parameter is the particle diameter
592 > ($d$).  Gay-Berne ellipsoids have an energy scaling parameter,
593 > $\epsilon^s$, which describes the well depth for two identical
594 > ellipsoids in a {\it side-by-side} configuration.  Additionally, a
595 > well depth aspect ratio, $\epsilon^r = \epsilon^e / \epsilon^s$,
596 > describes the ratio between the well depths in the {\it end-to-end}
597 > and side-by-side configurations.  For spheres, $\epsilon^r \equiv 1$.
598 > Moments of inertia are also required to describe the motion of primary
599 > particles with orientational degrees of freedom.
600 >
601 > \begin{table*}
602 > \begin{minipage}{\linewidth}
603 > \begin{center}
604 > \caption{Parameters for the primary particles in use by the rigid body
605 > models in figure \ref{fig:models}.}
606 > \begin{tabular}{lrcccccccc}
607 > \hline
608 > & & & & & & & \multicolumn{3}c{$\overleftrightarrow{\mathsf I}$ (amu \AA$^2$)} \\
609 > & & $d$ (\AA) & $l$ (\AA) & $\epsilon^s$ (kcal/mol) & $\epsilon^r$ &
610 > $m$ (amu) & $I_{xx}$ & $I_{yy}$ & $I_{zz}$ \\ \hline
611 > Sphere   & & 6.5 & $= d$ & 0.8 & 1 & 190 & 802.75 & 802.75  & 802.75  \\
612 > Ellipsoid & & 4.6 & 13.8  & 0.8 & 0.2 & 200 & 2105 & 2105 & 421 \\
613 > Dumbbell &(2 identical spheres) & 6.5 & $= d$ & 0.8 & 1   & 190 & 802.75 & 802.75 & 802.75 \\
614 > Banana  &(3 identical ellipsoids)& 4.2 & 11.2  & 0.8 & 0.2 & 240 & 10000 & 10000 & 0 \\
615 > Lipid: & Spherical Head & 6.5 & $= d$ & 0.185 & 1 & 196 & & & \\
616 >       & Ellipsoidal Tail & 4.6 & 13.8  & 0.8   & 0.2 & 760 & 45000 & 45000 & 9000 \\
617 > Solvent &  & 4.7 & $= d$ & 0.8 & 1   & 72.06 & & & \\
618 > \hline
619   \end{tabular}
620 + \label{tab:parameters}
621   \end{center}
622 < \end{table}
623 <
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.
622 > \end{minipage}
623 > \end{table*}
624  
625   \begin{figure}
626   \centering
627 < \includegraphics[width=\linewidth]{temperature.pdf}
628 < \caption[Plot of Temperature Fluctuation Versus Time]{Plot of
629 < temperature fluctuation versus time.} \label{langevin:temperature}
627 > \includegraphics[width=3in]{sketch}
628 > \caption[Sketch of the model systems]{A sketch of the model systems
629 > used in evaluating the behavior of the rigid body Langevin
630 > integrator.} \label{fig:models}
631   \end{figure}
632  
633 < \subsection{Langevin Dynamics simulation {\it vs} NVE simulations}
633 > \subsection{Simulation Methodology}
634  
635 < To validate our langevin dynamics simulation. We performed several NVE
636 < simulations with explicit solvents for different shaped
637 < molecules. There are one solute molecule and 1929 solvent molecules in
638 < NVE simulation. The parameters are shown in table
639 < \ref{tab:parameters}. The force field between spheres is standard
640 < Lennard-Jones, and ellipsoids interact with other ellipsoids and
641 < spheres with generalized Gay-Berne potential. All simulations are
642 < carried out at 300 K and 1 Atm. The time step is 25 ns, and a
643 < switching function was applied to all potentials to smoothly turn off
644 < the interactions between a range of $22$ and $25$ \AA.  The switching
645 < function was the standard (cubic) function,
635 > We performed reference microcanonical simulations with explicit
636 > solvents for each of the different model system.  In each case there
637 > was one solute model and 1929 solvent molecules present in the
638 > simulation box.  All simulations were equilibrated using a
639 > constant-pressure and temperature integrator with target values of 300
640 > K for the temperature and 1 atm for pressure.  Following this stage,
641 > further equilibration and sampling was done in a microcanonical
642 > ensemble.  Since the model bodies are typically quite massive, we were
643 > able to use a time step of 25 fs.  A switching function was applied to
644 > all potentials to smoothly turn off the interactions between a range
645 > of $22$ and $25$ \AA.  The switching function was the standard (cubic)
646 > function,
647   \begin{equation}
648   s(r) =
649          \begin{cases}
# Line 660 | Line 655 | We have computed translational diffusion constants for
655          \end{cases}
656   \label{eq:switchingFunc}
657   \end{equation}
658 < We have computed translational diffusion constants for lipid molecules
659 < from the mean-square displacement,
658 > To measure shear viscosities from our microcanonical simulations, we
659 > used the Einstein form of the pressure correlation function,\cite{hess:209}
660   \begin{equation}
661 < D = \lim_{t\rightarrow \infty} \frac{1}{6 t} \langle {|\left({\bf r}_{i}(t) - {\bf r}_{i}(0) \right)|}^2 \rangle,
661 > \eta = \lim_{t->\infty} \frac{V}{2 k_B T} \frac{d}{dt} \langle \left(
662 > \int_{t_0}^{t_0 + t} P_{xz}(t') dt' \right)^2 \rangle_{t_0}.
663 > \label{eq:shear}
664   \end{equation}
665 < of the solute molecules. Translational diffusion constants for the
669 < different shaped molecules are shown in table
670 < \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,
665 > A similar form exists for the bulk viscosity
666   \begin{equation}
667 < C_{\ell}(t)  =  \langle P_{\ell}\left({\bf \mu}_{i}(t) \cdot {\bf
668 < \mu}_{i}(0) \right)
667 > \kappa = \lim_{t->\infty} \frac{V}{2 k_B T} \frac{d}{dt} \langle \left(
668 > \int_{t_0}^{t_0 + t}
669 > \left(P\left(t'\right)-\langle P \rangle \right)dt'
670 > \right)^2 \rangle_{t_0}.
671   \end{equation}
672 < the results are shown in table \ref{tab:rotation}. We used einstein
673 < format of the pressure correlation function,
672 > Alternatively, the shear viscosity can also be calculated using a
673 > Green-Kubo formula with the off-diagonal pressure tensor correlation function,
674   \begin{equation}
675 < C_{\ell}(t)  =  \langle P_{\ell}\left({\bf \mu}_{i}(t) \cdot {\bf
676 < \mu}_{i}(0) \right)
675 > \eta = \frac{V}{k_B T} \int_0^{\infty} \langle P_{xz}(t_0) P_{xz}(t_0
676 > + t) \rangle_{t_0} dt,
677   \end{equation}
678 < to estimate the viscosity of the systems from NVE simulations. The
679 < viscosity can also be calculated by Green-Kubo pressure correlaton
680 < function,
681 < \begin{equation}
682 < C_{\ell}(t)  =  \langle P_{\ell}\left({\bf \mu}_{i}(t) \cdot {\bf
683 < \mu}_{i}(0) \right)
678 > although this method converges extremely slowly and is not practical
679 > for obtaining viscosities from molecular dynamics simulations.
680 >
681 > The Langevin dynamics for the different model systems were performed
682 > at the same temperature as the average temperature of the
683 > microcanonical simulations and with a solvent viscosity taken from
684 > Eq. (\ref{eq:shear}) applied to these simulations.  We used 1024
685 > independent solute simulations to obtain statistics on our Langevin
686 > integrator.
687 >
688 > \subsection{Analysis}
689 >
690 > The quantities of interest when comparing the Langevin integrator to
691 > analytic hydrodynamic equations and to molecular dynamics simulations
692 > are typically translational diffusion constants and orientational
693 > relaxation times.  Translational diffusion constants for point
694 > particles are computed easily from the long-time slope of the
695 > mean-square displacement,
696 > \begin{equation}
697 > D = \lim_{t\rightarrow \infty} \frac{1}{6 t} \langle {|\left({\bf r}_{i}(t) - {\bf r}_{i}(0) \right)|}^2 \rangle,
698   \end{equation}
699 < However, this method converges slowly, and the statistics are not good
700 < enough to give us a very accurate value. The langevin dynamics
701 < simulations for different shaped molecules are performed at the same
702 < conditions as the NVE simulations with viscosity estimated from NVE
703 < simulations. To get better statistics, 1024 non-interacting solute
704 < molecules are put into one simulation box for each langevin
705 < simulation, this is equal to 1024 simulations for single solute
706 < systems. The diffusion constants and rotation relaxation times for
699 > of the solute molecules.  For models in which the translational
700 > diffusion tensor (${\bf D}_{tt}$) has non-degenerate eigenvalues
701 > (i.e. any non-spherically-symmetric rigid body), it is possible to
702 > compute the diffusive behavior for motion parallel to each body-fixed
703 > axis by projecting the displacement of the particle onto the
704 > body-fixed reference frame at $t=0$.  With an isotropic solvent, as we
705 > have used in this study, there are differences between the three
706 > diffusion constants, but these must converge to the same value at
707 > longer times.  Translational diffusion constants for the different
708 > shaped models are shown in table \ref{tab:translation}.
709 >
710 > In general, the three eigenvalues ($D_1, D_2, D_3$) of the rotational
711 > diffusion tensor (${\bf D}_{rr}$) measure the diffusion of an object
712 > {\it around} a particular body-fixed axis and {\it not} the diffusion
713 > of a vector pointing along the axis.  However, these eigenvalues can
714 > be combined to find 5 characteristic rotational relaxation
715 > times,\cite{PhysRev.119.53,Berne90}
716 > \begin{eqnarray}
717 > 1 / \tau_1 & = & 6 D_r + 2 \Delta \\
718 > 1 / \tau_2 & = & 6 D_r - 2 \Delta \\
719 > 1 / \tau_3 & = & 3 (D_r + D_1)  \\
720 > 1 / \tau_4 & = & 3 (D_r + D_2) \\
721 > 1 / \tau_5 & = & 3 (D_r + D_3)
722 > \end{eqnarray}
723 > where
724 > \begin{equation}
725 > D_r = \frac{1}{3} \left(D_1 + D_2 + D_3 \right)
726 > \end{equation}
727 > and
728 > \begin{equation}
729 > \Delta = \left( (D_1 - D_2)^2 + (D_3 - D_1 )(D_3 - D_2)\right)^{1/2}
730 > \end{equation}
731 > Each of these characteristic times can be used to predict the decay of
732 > part of the rotational correlation function when $\ell = 2$,
733 > \begin{equation}
734 > C_2(t)  =  \frac{a^2}{N^2} e^{-t/\tau_1} + \frac{b^2}{N^2} e^{-t/\tau_2}.
735 > \end{equation}
736 > This is the same as the $F^2_{0,0}(t)$ correlation function that
737 > appears in Ref. \citen{Berne90}.  The amplitudes of the two decay
738 > terms are expressed in terms of three dimensionless functions of the
739 > eigenvalues: $a = \sqrt{3} (D_1 - D_2)$, $b = (2D_3 - D_1 - D_2 +
740 > 2\Delta)$, and $N = 2 \sqrt{\Delta b}$.  Similar expressions can be
741 > obtained for other angular momentum correlation
742 > functions.\cite{PhysRev.119.53,Berne90} In all of the model systems we
743 > studied, only one of the amplitudes of the two decay terms was
744 > non-zero, so it was possible to derive a single relaxation time for
745 > each of the hydrodynamic tensors. In many cases, these characteristic
746 > times are averaged and reported in the literature as a single relaxation
747 > time,\cite{Garcia-de-la-Torre:1997qy}
748 > \begin{equation}
749 > 1 / \tau_0 = \frac{1}{5} \sum_{i=1}^5 \tau_{i}^{-1},
750 > \end{equation}
751 > although for the cases reported here, this averaging is not necessary
752 > and only one of the five relaxation times is relevant.
753 >
754 > To test the Langevin integrator's behavior for rotational relaxation,
755 > we have compared the analytical orientational relaxation times (if
756 > they are known) with the general result from the diffusion tensor and
757 > with the results from both the explicitly solvated molecular dynamics
758 > and Langevin simulations.  Relaxation times from simulations (both
759 > microcanonical and Langevin), were computed using Legendre polynomial
760 > correlation functions for a unit vector (${\bf u}$) fixed along one or
761 > more of the body-fixed axes of the model.
762 > \begin{equation}
763 > C_{\ell}(t)  =  \langle P_{\ell}\left({\bf u}_{i}(t) \cdot {\bf
764 > u}_{i}(0) \right)
765 > \end{equation}
766 > For simulations in the high-friction limit, orientational correlation
767 > times can then be obtained from exponential fits of this function, or by
768 > integrating,
769 > \begin{equation}
770 > \tau = \ell (\ell + 1) \int_0^{\infty} C_{\ell}(t) dt.
771 > \end{equation}
772 > In lower-friction solvents, the Legendre correlation functions often
773 > exhibit non-exponential decay, and may not be characterized by a
774 > single decay constant.
775 >
776 > In table \ref{tab:rotation} we show the characteristic rotational
777 > relaxation times (based on the diffusion tensor) for each of the model
778 > systems compared with the values obtained via microcanonical and Langevin
779 > simulations.
780 >
781 > \subsection{Spherical particles}
782 > Our model system for spherical particles was a Lennard-Jones sphere of
783 > diameter ($\sigma$) 6.5 \AA\ in a sea of smaller spheres ($\sigma$ =
784 > 4.7 \AA).  The well depth ($\epsilon$) for both particles was set to
785 > an arbitrary value of 0.8 kcal/mol.  
786 >
787 > The Stokes-Einstein behavior of large spherical particles in
788 > hydrodynamic flows is well known, giving translational friction
789 > coefficients of $6 \pi \eta R$ (stick boundary conditions) and
790 > rotational friction coefficients of $8 \pi \eta R^3$.  Recently,
791 > Schmidt and Skinner have computed the behavior of spherical tag
792 > particles in molecular dynamics simulations, and have shown that {\it
793 > slip} boundary conditions ($\Xi_{tt} = 4 \pi \eta R$) may be more
794 > appropriate for molecule-sized spheres embedded in a sea of spherical
795 > qsolvent particles.\cite{Schmidt:2004fj,Schmidt:2003kx}
796 >
797 > Our simulation results show similar behavior to the behavior observed
798 > by Schmidt and Skinner.  The diffusion constant obtained from our
799 > microcanonical molecular dynamics simulations lies between the slip
800 > and stick boundary condition results obtained via Stokes-Einstein
801 > behavior.  Since the Langevin integrator assumes Stokes-Einstein stick
802 > boundary conditions in calculating the drag and random forces for
803 > spherical particles, our Langevin routine obtains nearly quantitative
804 > agreement with the hydrodynamic results for spherical particles.  One
805 > avenue for improvement of the method would be to compute elements of
806 > $\Xi_{tt}$ assuming behavior intermediate between the two boundary
807 > conditions.
808 >
809 > In these simulations, our spherical particles were structureless, so
810 > there is no way to obtain rotational correlation times for this model
811 > system.
812 >
813 > \subsubsection{Ellipsoids}
814 > For uniaxial ellipsoids ($a > b = c$) , Perrin's formulae for both
815 > translational and rotational diffusion of each of the body-fixed axes
816 > can be combined to give a single translational diffusion
817 > constant,\cite{Berne90}
818 > \begin{equation}
819 > D = \frac{k_B T}{6 \pi \eta a} G(\rho),
820 > \label{Dperrin}
821 > \end{equation}
822 > as well as a single rotational diffusion coefficient,
823 > \begin{equation}
824 > \Theta = \frac{3 k_B T}{16 \pi \eta a^3} \left\{ \frac{(2 - \rho^2)
825 > G(\rho) - 1}{1 - \rho^4} \right\}.
826 > \label{ThetaPerrin}
827 > \end{equation}
828 > In these expressions, $G(\rho)$ is a function of the axial ratio
829 > ($\rho = b / a$), which for prolate ellipsoids, is
830 > \begin{equation}
831 > G(\rho) = (1- \rho^2)^{-1/2} \ln \left\{ \frac{1 + (1 -
832 > \rho^2)^{1/2}}{\rho} \right\}
833 > \label{GPerrin}
834 > \end{equation}
835 > Again, there is some uncertainty about the correct boundary conditions
836 > to use for molecular-scale ellipsoids in a sea of similarly-sized
837 > solvent particles.  Ravichandran and Bagchi found that {\it slip}
838 > boundary conditions most closely resembled the simulation
839 > results,\cite{Ravichandran:1999fk} in agreement with earlier work of
840 > Tang and Evans.\cite{TANG:1993lr}
841 >
842 > Even though there are analytic resistance tensors for ellipsoids, we
843 > constructed a rough-shell model using 2135 beads (each with a diameter
844 > of 0.25 \AA) to approximate the shape of the modle ellipsoid.  We
845 > compared the Langevin dynamics from both the simple ellipsoidal
846 > resistance tensor and the rough shell approximation with
847 > microcanonical simulations and the predictions of Perrin.  As in the
848 > case of our spherical model system, the Langevin integrator reproduces
849 > almost exactly the behavior of the Perrin formulae (which is
850 > unsurprising given that the Perrin formulae were used to derive the
851 > drag and random forces applied to the ellipsoid).  We obtain
852 > translational diffusion constants and rotational correlation times
853 > that are within a few percent of the analytic values for both the
854 > exact treatment of the diffusion tensor as well as the rough-shell
855 > model for the ellipsoid.
856 >
857 > The translational diffusion constants from the microcanonical simulations
858 > agree well with the predictions of the Perrin model, although the rotational
859 > correlation times are a factor of 2 shorter than expected from hydrodynamic
860 > theory.  One explanation for the slower rotation
861 > of explicitly-solvated ellipsoids is the possibility that solute-solvent
862 > collisions happen at both ends of the solute whenever the principal
863 > axis of the ellipsoid is turning. In the upper portion of figure
864 > \ref{fig:explanation} we sketch a physical picture of this explanation.
865 > Since our Langevin integrator is providing nearly quantitative agreement with
866 > the Perrin model, it also predicts orientational diffusion for ellipsoids that
867 > exceed explicitly solvated correlation times by a factor of two.
868 >
869 > \begin{figure}
870 > \centering
871 > \includegraphics[width=6in]{explanation}
872 > \caption[Explanations of the differences between orientational correlation times for explicitly-solvated models and hydrodynamics predictions]{Explanations of the differences between orientational correlation times for explicitly-solvated models and hydrodynamic predictions.   For the ellipsoids (upper figures), rotation of the principal axis can involve correlated collisions at both sides of the solute.  In the rigid dumbbell model (lower figures), the large size of the explicit solvent spheres prevents them from coming in contact with a substantial fraction of the surface area of the dumbbell.  Therefore, the explicit solvent only provides drag over a substantially reduced surface area of this model, where the hydrodynamic theories utilize the entire surface area for estimating rotational diffusion.
873 > } \label{fig:explanation}
874 > \end{figure}
875 >
876 > \subsubsection{Rigid dumbbells}
877 > Perhaps the only {\it composite} rigid body for which analytic
878 > expressions for the hydrodynamic tensor are available is the
879 > two-sphere dumbbell model.  This model consists of two non-overlapping
880 > spheres held by a rigid bond connecting their centers. There are
881 > competing expressions for the 6x6 resistance tensor for this
882 > model. Equation (\ref{introEquation:oseenTensor}) above gives the
883 > original Oseen tensor, while the second order expression introduced by
884 > Rotne and Prager,\cite{Rotne1969} and improved by Garc\'{i}a de la
885 > Torre and Bloomfield,\cite{Torre1977} is given above as
886 > Eq. (\ref{introEquation:RPTensorNonOverlapped}).  In our case, we use
887 > a model dumbbell in which the two spheres are identical Lennard-Jones
888 > particles ($\sigma$ = 6.5 \AA\ , $\epsilon$ = 0.8 kcal / mol) held at
889 > a distance of 6.532 \AA.
890 >
891 > The theoretical values for the translational diffusion constant of the
892 > dumbbell are calculated from the work of Stimson and Jeffery, who
893 > studied the motion of this system in a flow parallel to the
894 > inter-sphere axis,\cite{Stimson:1926qy} and Davis, who studied the
895 > motion in a flow {\it perpendicular} to the inter-sphere
896 > axis.\cite{Davis:1969uq} We know of no analytic solutions for the {\it
897 > orientational} correlation times for this model system (other than
898 > those derived from the 6 x 6 tensors mentioned above).
899 >
900 > The bead model for this model system comprises the two large spheres
901 > by themselves, while the rough shell approximation used 3368 separate
902 > beads (each with a diameter of 0.25 \AA) to approximate the shape of
903 > the rigid body.  The hydrodynamics tensors computed from both the bead
904 > and rough shell models are remarkably similar.  Computing the initial
905 > hydrodynamic tensor for a rough shell model can be quite expensive (in
906 > this case it requires inverting a 10104 x 10104 matrix), while the
907 > bead model is typically easy to compute (in this case requiring
908 > inversion of a 6 x 6 matrix).  
909 >
910 > \begin{figure}
911 > \centering
912 > \includegraphics[width=3in]{RoughShell}
913 > \caption[Model rigid bodies and their rough shell approximations]{The
914 > model rigid bodies (left column) used to test this algorithm and their
915 > rough-shell approximations (right-column) that were used to compute
916 > the hydrodynamic tensors.  The top two models (ellipsoid and dumbbell)
917 > have analytic solutions and were used to test the rough shell
918 > approximation.  The lower two models (banana and lipid) were compared
919 > with explicitly-solvated molecular dynamics simulations. }
920 > \label{fig:roughShell}
921 > \end{figure}
922 >
923 >
924 > Once the hydrodynamic tensor has been computed, there is no additional
925 > penalty for carrying out a Langevin simulation with either of the two
926 > different hydrodynamics models.  Our naive expectation is that since
927 > the rigid body's surface is roughened under the various shell models,
928 > the diffusion constants will be even farther from the ``slip''
929 > boundary conditions than observed for the bead model (which uses a
930 > Stokes-Einstein model to arrive at the hydrodynamic tensor).  For the
931 > dumbbell, this prediction is correct although all of the Langevin
932 > diffusion constants are within 6\% of the diffusion constant predicted
933 > from the fully solvated system.
934 >
935 > For rotational motion, Langevin integration (and the hydrodynamic tensor)
936 > yields rotational correlation times that are substantially shorter than those
937 > obtained from explicitly-solvated simulations.  It is likely that this is due
938 > to the large size of the explicit solvent spheres, a feature that prevents
939 > the solvent from coming in contact with a substantial fraction of the surface
940 > area of the dumbbell.  Therefore, the explicit solvent only provides drag
941 > over a substantially reduced surface area of this model, while the
942 > hydrodynamic theories utilize the entire surface area for estimating
943 > rotational diffusion.  A sketch of the free volume available in the explicit
944 > solvent simulations is shown in figure \ref{fig:explanation}.
945 >
946 > \subsubsection{Ellipsoidal-composite banana-shaped molecules}
947 > Banana-shaped rigid bodies composed of three Gay-Berne ellipsoids
948 > have been used by Orlandi {\it et al.} to observe
949 > mesophases in coarse-grained models bent-core liquid crystalline
950 > molecules.\cite{Orlandi:2006fk}  We have used the same overlapping
951 > ellipsoids as a way to test the behavior of our algorithm for a
952 > structure of some interest to the materials science community,
953 > although since we are interested in capturing only the hydrodynamic
954 > behavior of this model, we have left out the dipolar interactions of the
955 > original Orlandi model.
956 >
957 > A reference system composed of a single banana rigid body embedded in a
958 > sea of 1929 solvent particles was created and run under standard
959 > (microcanonical) molecular dynamics.  The resulting viscosity of this
960 > mixture was 0.298 centipoise (as estimated using Eq. (\ref{eq:shear})).
961 > To calculate the hydrodynamic properties of the banana rigid body model,
962 > we created a rough shell (see Fig.~\ref{langevin:roughShell}), in which
963 > the banana is represented as a ``shell'' made of 3321 identical beads
964 > (0.25 \AA in diameter) distributed on the surface.  Applying the
965 > procedure described in Sec.~\ref{introEquation:ResistanceTensorArbitraryOrigin}, we
966 > identified the center of resistance at (0 $\rm{\AA}$, 0.81 $\rm{\AA}$, 0 $\rm{\AA}$), as well as the resistance tensor,
967 > \[
968 > \left( {\begin{array}{*{20}c}
969 > 0.9261 & 0 & 0&0&0.08585&0.2057\\
970 > 0& 0.9270&-0.007063& 0.08585&0&0\\
971 > 0&-0.007063&0.7494&0.2057&0&0\\
972 > 0&0.0858&0.2057& 58.64& 0&0\\0.08585&0&0&0&48.30&3.219&\\0.2057&0&0&0&3.219&10.7373\\\end{array}} \right).
973 > \]
974 > where the units for translational, translation-rotation coupling and rotational
975 > tensors are $\frac{kcal \cdot fs}{mol \cdot \rm{\AA}^2}$, $\frac{kcal \cdot fs}{
976 > mol \cdot \rm{\AA} \cdot rad}$ and $\frac{kcal \cdot fs}{mol \cdot rad^2}$ respe
977 > ctively.
978 >
979 > The Langevin rigid-body integrator (and the hydrodynamic diffusion tensor)
980 > are essentially quantitative for translational diffusion of this model.  
981 > Orientational correlation times under the Langevin rigid-body integrator
982 > are within 11\% of the values obtained from explicit solvent, but these
983 > models also exhibit some solvent inaccessible surface area in the
984 > explicitly-solvated case.  
985 >
986 > \subsubsection{Composite sphero-ellipsoids}
987 > Spherical heads perched on the ends of Gay-Berne ellipsoids have been
988 > used recently as models for lipid molecules.\cite{SunGezelter08,Ayton01}
989 >
990 > The diffusion constants and rotation relaxation times for
991   different shaped molecules are shown in table \ref{tab:translation}
992   and \ref{tab:rotation} to compare to the results calculated from NVE
993   simulations. The theoretical values for sphere is calculated from the
994   Stokes-Einstein law, the theoretical values for ellipsoid is
995 < calculated from Perrin's fomula, the theoretical values for dumbbell
703 < is calculated from StinXX and Davis theory. The exact method is
995 > calculated from Perrin's fomula,  The exact method is
996   applied to the langevin dynamics simulations for sphere and ellipsoid,
997   the bead model is applied to the simulation for dumbbell molecule, and
998   the rough shell model is applied to ellipsoid, dumbbell, banana and
# Line 740 | Line 1032 | are also mimiced reasonablly well.
1032   \begin{table*}
1033   \begin{minipage}{\linewidth}
1034   \begin{center}
1035 < \caption{}
1036 < \begin{tabular}{llccccccc}
1037 < \hline
1038 <  & & Sphere & Ellipsoid & Dumbbell(2 spheres) & Banana(3 ellpsoids) &
1039 < Lipid(head) & lipid(tail) & Solvent \\
1040 < \hline
1041 < $d$ (\AA) & & 6.5 & 4.6  & 6.5 &  4.2 & 6.5 & 4.6 & 4.7 \\
1042 < $l$ (\AA) & & $= d$ & 13.8 & $=d$ & 11.2 & $=d$ & 13.8 & 4.7 \\
1043 < $\epsilon^s$ (kcal/mol) & & 0.8 & 0.8 & 0.8 & 0.8 & 0.185 & 0.8 & 0.8 \\
1044 < $\epsilon_r$ (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}
1035 > \caption{Translational diffusion constants (D) for the model systems
1036 > calculated using microcanonical simulations (with explicit solvent),
1037 > theoretical predictions, and Langevin simulations (with implicit solvent).
1038 > Analytical solutions for the exactly-solved hydrodynamics models are
1039 > from Refs. \citen{Einstein05} (sphere), \citen{Perrin1934} and \citen{Perrin1936}
1040 > (ellipsoid), \citen{Stimson:1926qy} and \citen{Davis:1969uq}
1041 > (dumbbell). The other model systems have no known analytic solution.
1042 > All  diffusion constants are reported in units of $10^{-3}$ cm$^2$ / ps (=
1043 > $10^{-4}$ \AA$^2$  / fs). }
1044 > \begin{tabular}{lccccccc}
1045   \hline
1046 < & & & & &Translation \\
1047 < \hline
1048 < & NVE &  & Theoretical & Langevin & \\
774 < \hline
775 < & $\eta$ & D & D & method & D \\
1046 > & \multicolumn{2}c{microcanonical simulation} & & \multicolumn{3}c{Theoretical} & Langevin \\
1047 > \cline{2-3} \cline{5-7}
1048 > model & $\eta$ (centipoise)  & D & & Analytical & method & Hydrodynamics & simulation \\
1049   \hline
1050 < sphere & 3.480159e-03 & 1.643135e-04 & 1.942779e-04 & exact & 1.982283e-04 \\
1051 < ellipsoid & 2.551262e-03 & 2.437492e-04 & 2.335756e-04 & exact & 2.374905e-04 \\
1052 < & 2.551262e-03  & 2.437492e-04 & 2.335756e-04 & rough shell & 2.284088e-04 \\
1053 < dumbell & 2.41276e-03  & 2.129432e-04 & 2.090239e-04 & bead model & 2.148098e-04 \\
1054 < & 2.41276e-03 & 2.129432e-04 & 2.090239e-04 & rough shell & 2.013219e-04 \\
1055 < banana & 2.9846e-03 & 1.527819e-04 &  & rough shell & 1.54807e-04 \\
1056 < lipid & 3.488661e-03 & 0.9562979e-04 &  & rough shell & 1.320987e-04 \\
1050 > sphere    & 0.261  & ?    & & 2.59 & exact       & 2.59 & 2.56 \\
1051 > ellipsoid & 0.255  & 2.44 & & 2.34 & exact       & 2.34 & 2.37 \\
1052 >          & 0.255  & 2.44 & & 2.34 & rough shell & 2.36 & 2.28 \\
1053 > dumbbell  & 0.322  & ?    & & 1.57 & bead model  & 1.57 & 1.57 \\
1054 >          & 0.322  & ?    & & 1.57 & rough shell & ?    & ?    \\
1055 > banana    & 0.298  & 1.53 & &      & rough shell & 1.56 & 1.55 \\
1056 > lipid     & 0.349  & 0.96 & &      & rough shell & 1.33 & 1.32 \\
1057   \end{tabular}
1058   \label{tab:translation}
1059   \end{center}
# Line 790 | Line 1063 | lipid & 3.488661e-03 & 0.9562979e-04 &  & rough shell
1063   \begin{table*}
1064   \begin{minipage}{\linewidth}
1065   \begin{center}
1066 < \caption{}
1067 < \begin{tabular}{lccccc}
1066 > \caption{Orientational relaxation times ($\tau$) for the model systems using
1067 > microcanonical simulation (with explicit solvent), theoretical
1068 > predictions, and Langevin simulations (with implicit solvent). All
1069 > relaxation times are for the rotational correlation function with
1070 > $\ell = 2$ and are reported in units of ps.  The ellipsoidal model has
1071 > an exact solution for the orientational correlation time due to
1072 > Perrin, but the other model systems have no known analytic solution.}
1073 > \begin{tabular}{lccccccc}
1074   \hline
1075 < & & & & &Rotation \\
1076 < \hline
1077 < & NVE &  & Theoretical & Langevin & \\
799 < \hline
800 < & $\eta$ & $\tau_0$ & $\tau_0$ & method & $\tau_0$ \\
1075 > & \multicolumn{2}c{microcanonical simulation} & & \multicolumn{3}c{Theoretical} & Langevin \\
1076 > \cline{2-3} \cline{5-7}
1077 > model & $\eta$ (centipoise) & $\tau$ & & Perrin & method & Hydrodynamic  & simulation \\
1078   \hline
1079 < sphere & 3.480159e-03 &  & 1.208178e+04 & exact & 1.20628e+04 \\
1080 < ellipsoid & 2.551262e-03 & 4.66806e+04 & 2.198986e+04 & exact & 2.21507e+04 \\
1081 < & 2.551262e-03 & 4.66806e+04 & 2.198986e+04 & rough shell & 2.21714e+04 \\
1082 < dumbell & 2.41276e-03 & 1.42974e+04 &  & bead model & 7.12435e+04 \\
1083 < & 2.41276e-03 & 1.42974e+04 &  & rough shell & 7.04765e+04 \\
1084 < banana & 2.9846e-03 & 6.38323e+04 &  & rough shell & 7.0945e+04 \\
1085 < lipid & 3.488661e-03 & 7.79595e+04 &  & rough shell & 7.78886e+04 \\
1079 > sphere    & 0.261  &      & & 9.06 & exact       & 9.06 & 9.11 \\
1080 > ellipsoid & 0.255  & 46.7 & & 22.0 & exact       & 22.0 & 22.2 \\
1081 >          & 0.255  & 46.7 & & 22.0 & rough shell & 22.6 & 22.2 \\
1082 > dumbbell  & 0.322  & 14.0 & &      & bead model  & 52.3 & 52.8 \\
1083 >          & 0.322  & 14.0 & &      & rough shell & ?    & ?    \\
1084 > banana    & 0.298  & 63.8 & &      & rough shell & 70.9 & 70.9 \\
1085 > lipid     & 0.349  & 78.0 & &      & rough shell & 76.9 & 77.9 \\
1086 > \hline
1087   \end{tabular}
1088   \label{tab:rotation}
1089   \end{center}
# Line 822 | Line 1100 | of magnitude.
1100   100 ns run. The efficiency of the simulation is increased by one order
1101   of magnitude.
1102  
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
862 \begin{figure}
863 \centering
864 \includegraphics[width=\linewidth]{roughShell.pdf}
865 \caption[Rough shell model for banana shaped molecule]{Rough shell
866 model for banana shaped molecule.} \label{langevin:roughShell}
867 \end{figure}
868
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
1103   \section{Conclusions}
1104  
1105   We have presented a new Langevin algorithm by incorporating the
1106   hydrodynamics properties of arbitrary shaped molecules into an
1107 < advanced symplectic integration scheme. The temperature control
1108 < ability of this algorithm was demonstrated by a set of simulations
1109 < with different viscosities. It was also shown to have significant
1110 < 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.
1107 > advanced symplectic integration scheme. Further studies in systems
1108 > involving banana shaped molecules illustrated that the dynamic
1109 > properties could be preserved by using this new algorithm as an
1110 > implicit solvent model.
1111  
1112  
1113   \section{Acknowledgments}
# Line 911 | Line 1117 | of Notre Dame.
1117   of Notre Dame.
1118   \newpage
1119  
1120 < \bibliographystyle{jcp2}
1120 > \bibliographystyle{jcp}
1121   \bibliography{langevin}
1122  
1123   \end{document}

Diff Legend

Removed lines
+ Added lines
< Changed lines
> Changed lines