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 3308 by gezelter, Fri Jan 11 17:04:12 2008 UTC vs.
Revision 3310 by gezelter, Fri Jan 11 22:08:57 2008 UTC

# Line 170 | Line 170 | into the sophisticated rigid body dynamics algorithms.
170   algorithm for arbitrary-shaped rigid particles by integrating the
171   accurate estimation of friction tensor from hydrodynamics theory
172   into the sophisticated rigid body dynamics algorithms.
173
174 \section{Computational Methods{\label{methodSec}}}
173  
174   \subsection{\label{introSection:frictionTensor}Friction Tensor}
175   Theoretically, the friction kernel can be determined using the
# 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}
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
# Line 631 | Line 644 | integrator.} \label{fig:models}
644   \end{figure}
645  
646   \subsection{Simulation Methodology}
634
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
# Line 640 | Line 652 | able to use a time step of 25 fs.  A switching functio
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.  A switching function was applied to
656 < all potentials to smoothly turn off the interactions between a range
657 < of $22$ and $25$ \AA.  The switching function was the standard (cubic)
658 < function,
655 > able to use a time step of 25 fs.
656 >
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 655 | Line 713 | To measure shear viscosities from our microcanonical s
713          \end{cases}
714   \label{eq:switchingFunc}
715   \end{equation}
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 < \eta = \lim_{t->\infty} \frac{V}{2 k_B T} \frac{d}{dt} \langle \left(
721 < \int_{t_0}^{t_0 + t} P_{xz}(t') dt' \right)^2 \rangle_{t_0}.
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 < \kappa = \lim_{t->\infty} \frac{V}{2 k_B T} \frac{d}{dt} \langle \left(
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)-\langle P \rangle \right)dt'
729 < \right)^2 \rangle_{t_0}.
728 > \left(P\left(t'\right)-\left\langle P \right\rangle \right)dt'
729 > \right)^2 \right\rangle_{t_0}.
730   \end{equation}
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 < \eta = \frac{V}{k_B T} \int_0^{\infty} \langle P_{xz}(t_0) P_{xz}(t_0
735 < + t) \rangle_{t_0} dt,
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   although this method converges extremely slowly and is not practical
738   for obtaining viscosities from molecular dynamics simulations.
# Line 694 | Line 753 | D = \lim_{t\rightarrow \infty} \frac{1}{6 t} \langle {
753   particles are computed easily from the long-time slope of the
754   mean-square displacement,
755   \begin{equation}
756 < D = \lim_{t\rightarrow \infty} \frac{1}{6 t} \langle {|\left({\bf r}_{i}(t) - {\bf r}_{i}(0) \right)|}^2 \rangle,
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   of the solute molecules.  For models in which the translational
759   diffusion tensor (${\bf D}_{tt}$) has non-degenerate eigenvalues
# Line 760 | Line 819 | C_{\ell}(t)  =  \langle P_{\ell}\left({\bf u}_{i}(t) \
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)  =  \langle P_{\ell}\left({\bf u}_{i}(t) \cdot {\bf
823 < u}_{i}(0) \right)
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
# Line 792 | Line 851 | qsolvent particles.\cite{Schmidt:2004fj,Schmidt:2003kx
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 < qsolvent particles.\cite{Schmidt:2004fj,Schmidt:2003kx}
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
# Line 806 | Line 865 | In these simulations, our spherical particles were str
865   $\Xi_{tt}$ assuming behavior intermediate between the two boundary
866   conditions.
867  
868 < In these simulations, our spherical particles were structureless, so
869 < there is no way to obtain rotational correlation times for this model
870 < system.
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 < \subsubsection{Ellipsoids}
874 < For uniaxial ellipsoids ($a > b = c$) , Perrin's formulae for both
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}
# Line 841 | Line 901 | of 0.25 \AA) to approximate the shape of the modle ell
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 modle ellipsoid.  We
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
# Line 866 | Line 926 | exceed explicitly solvated correlation times by a fact
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 < \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}
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
# Line 909 | Line 962 | inversion of a 6 x 6 matrix).  
962  
963   \begin{figure}
964   \centering
965 < \includegraphics[width=3in]{RoughShell}
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
# Line 943 | Line 996 | solvent simulations is shown in figure \ref{fig:explan
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 < \subsubsection{Ellipsoidal-composite banana-shaped molecules}
1000 < Banana-shaped rigid bodies composed of three Gay-Berne ellipsoids
1001 < have been used by Orlandi {\it et al.} to observe
1002 < mesophases in coarse-grained models bent-core liquid crystalline
1003 < molecules.\cite{Orlandi:2006fk}  We have used the same overlapping
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 the
1030 < original Orlandi model.
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{langevin:roughShell}), in which
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
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 at (0 $\rm{\AA}$, 0.81 $\rm{\AA}$, 0 $\rm{\AA}$), as well as the resistance tensor,
1042 < \[
1041 > identified the center of resistance, ${\bf r} = $(0 \AA, 0.81 \AA, 0 \AA), as
1042 > well as the resistance tensor,
1043 > \begin{equation*}
1044 > \Xi =
1045   \left( {\begin{array}{*{20}c}
1046   0.9261 & 0 & 0&0&0.08585&0.2057\\
1047   0& 0.9270&-0.007063& 0.08585&0&0\\
1048   0&-0.007063&0.7494&0.2057&0&0\\
1049 < 0&0.0858&0.2057& 58.64& 0&0\\0.08585&0&0&0&48.30&3.219&\\0.2057&0&0&0&3.219&10.7373\\\end{array}} \right).
1050 < \]
1051 < where the units for translational, translation-rotation coupling and rotational
1052 < tensors are $\frac{kcal \cdot fs}{mol \cdot \rm{\AA}^2}$, $\frac{kcal \cdot fs}{
1053 < mol \cdot \rm{\AA} \cdot rad}$ and $\frac{kcal \cdot fs}{mol \cdot rad^2}$ respe
977 < ctively.
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.  
# Line 983 | Line 1059 | explicitly-solvated case.  
1059   models also exhibit some solvent inaccessible surface area in the
1060   explicitly-solvated case.  
1061  
1062 < \subsubsection{Composite sphero-ellipsoids}
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 < 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 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
999 < lipid molecules. The results from all the langevin dynamics
1000 < simulations, including exact, bead model and rough shell, match the
1001 < theoretical values perfectly for all different shaped molecules. This
1002 < indicates that our simulation package for langevin dynamics is working
1003 < well. The approxiate methods ( bead model and rough shell model) are
1004 < accurate enough for the current simulations. The goal of the langevin
1005 < dynamics theory is to replace the explicit solvents by the friction
1006 < forces. We compared the dynamic properties of different shaped
1007 < molecules in langevin dynamics simulations with that in NVE
1008 < simulations. The results are reasonable close. Overall, the
1009 < translational diffusion constants calculated from langevin dynamics
1010 < simulations are very close to the values from the NVE simulation. For
1011 < sphere and lipid molecules, the diffusion constants are a little bit
1012 < off from the NVE simulation results. One possible reason is that the
1013 < calculation of the viscosity is very difficult to be accurate. Another
1014 < possible reason is that although we save very frequently during the
1015 < NVE simulations and run pretty long time simulations, there is only
1016 < one solute molecule in the system which makes the calculation for the
1017 < diffusion constant difficult. The sphere molecule behaves as a free
1018 < rotor in the solvent, so there is no rotation relaxation time
1019 < calculated from NVE simulations. The rotation relaxation time is not
1020 < very close to the NVE simulations results. The banana and lipid
1021 < molecules match the NVE simulations results pretty well. The mismatch
1022 < between langevin dynamics and NVE simulation for ellipsoid is possibly
1023 < caused by the slip boundary condition. For dumbbell, the mismatch is
1024 < caused by the size of the solvent molecule is pretty large compared to
1025 < dumbbell molecule in NVE simulations.
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 1047 | Line 1090 | sphere    & 0.348  & 1.64 & & 1.94 & exact       & 1.9
1090   \cline{2-3} \cline{5-7}
1091   model & $\eta$ (centipoise)  & D & & Analytical & method & Hydrodynamics & simulation \\
1092   \hline
1093 < sphere    & 0.348  & 1.64 & & 1.94 & exact       & 1.94 & 1.98 \\
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  & 0.241  & 2.13 & & 2.09 & bead model  & 2.10 & 2.15 \\
1097 <          & 0.241  & 2.13 & & 2.09 & rough shell & 2.03 & 2.01 \\
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}
# Line 1076 | Line 1119 | ellipsoid & 0.255  & 46.7 & & 22.0 & exact       & 22.
1119   \cline{2-3} \cline{5-7}
1120   model & $\eta$ (centipoise) & $\tau$ & & Perrin & method & Hydrodynamic  & simulation \\
1121   \hline
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  & 0.241  & 14.3 & &      & bead model  & 39.2 & 71.2 \\
1126 <          & 0.241  & 14.3 & &      & rough shell & 32.6 & 70.5 \\
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
# Line 1089 | Line 1133 | Langevin dynamics simulations are applied to study the
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  
1154 + \begin{figure}
1155 + \centering
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 +
1163   \section{Conclusions}
1164  
1165   We have presented a new Langevin algorithm by incorporating the

Diff Legend

Removed lines
+ Added lines
< Changed lines
> Changed lines