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 3306 by xsun, Fri Jan 11 15:32:15 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 608 | Line 621 | Sphere   & & 6.5 & $= d$ & 0.8 & 1 & 190 & & &  \\
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 & & &  \\
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 & & & \\
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 \\
# 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 854 | Line 914 | The agreement with the translational diffusion constan
914   exact treatment of the diffusion tensor as well as the rough-shell
915   model for the ellipsoid.
916  
917 < The agreement with the translational diffusion constants from the
918 < microcanonical simulations is quite good, although the rotational
919 < correlation times are a factor of 2 shorter than the predictions of
920 < the Perrin hydrodynamic model.
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 < \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 891 | Line 958 | inversion of a 6 x 6 matrix).
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).
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
# Line 904 | Line 985 | For rotational motion, Langevin integration yields
985   diffusion constants are within 6\% of the diffusion constant predicted
986   from the fully solvated system.
987  
988 < For rotational motion, Langevin integration yields
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  
909 \subsubsection{Ellipsoidal-composite banana-shaped molecules}
999  
1000 < Banana-shaped rigid bodies composed of composites of Gay-Berne
1001 < ellipsoids 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 overlapping
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 leave out the dipolar interactions of the
1030 < original Orlandi model.
1031 <
1032 < \subsubsection{Composite sphero-ellipsoids}
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{Temperature Control}
929 <
930 < As shown in Eq.~\ref{randomForce}, random collisions associated with
931 < the solvent's thermal motions is controlled by the external
932 < temperature. The capability to maintain the temperature of the whole
933 < system was usually used to measure the stability and efficiency of
934 < the algorithm. In order to verify the stability of this new
935 < algorithm, a series of simulations are performed on system
936 < consisiting of 256 SSD water molecules with different viscosities.
937 < The initial configuration for the simulations is taken from a 1ns
938 < NVT simulation with a cubic box of 19.7166~\AA. All simulation are
939 < carried out with cutoff radius of 9~\AA and 2 fs time step for 1 ns
940 < with reference temperature at 300~K. The average temperature as a
941 < function of $\eta$ is shown in Table \ref{langevin:viscosity} where
942 < the temperatures range from 303.04~K to 300.47~K for $\eta = 0.01 -
943 < 1$ poise. The better temperature control at higher viscosity can be
944 < explained by the finite size effect and relative slow relaxation
945 < rate at lower viscosity regime.
946 < \begin{table}
947 < \caption{AVERAGE TEMPERATURES FROM LANGEVIN DYNAMICS SIMULATIONS OF
948 < SSD WATER MOLECULES WITH REFERENCE TEMPERATURE AT 300~K.}
949 < \label{langevin:viscosity}
950 < \begin{center}
951 < \begin{tabular}{lll}
952 <  \hline
953 <  $\eta$ & $\text{T}_{\text{avg}}$ & $\text{T}_{\text{rms}}$ \\
954 <  \hline
955 <  1    & 300.47 & 10.99 \\
956 <  0.1  & 301.19 & 11.136 \\
957 <  0.01 & 303.04 & 11.796 \\
958 <  \hline
959 < \end{tabular}
960 < \end{center}
961 < \end{table}
962 <
963 < Another set of calculations were performed to study the efficiency of
964 < temperature control using different temperature coupling schemes.
965 < The starting configuration is cooled to 173~K and evolved using NVE,
966 < NVT, and Langevin dynamic with time step of 2 fs.
967 < Fig.~\ref{langevin:temperature} shows the heating curve obtained as
968 < the systems reach equilibrium. The orange curve in
969 < Fig.~\ref{langevin:temperature} represents the simulation using
970 < Nos\'e-Hoover temperature scaling scheme with thermostat of 5 ps
971 < which gives reasonable tight coupling, while the blue one from
972 < Langevin dynamics with viscosity of 0.1 poise demonstrates a faster
973 < scaling to the desire temperature. When $ \eta = 0$, Langevin dynamics becomes normal
974 < NVE (see orange curve in Fig.~\ref{langevin:temperature}) which
975 < loses the temperature control ability.
976 <
977 < \begin{figure}
978 < \centering
979 < \includegraphics[width=\linewidth]{temperature}
980 < \caption[Plot of Temperature Fluctuation Versus Time]{Plot of
981 < temperature fluctuation versus time.} \label{langevin:temperature}
982 < \end{figure}
983 <
984 <
985 < The diffusion constants and rotation relaxation times for
986 < different shaped molecules are shown in table \ref{tab:translation}
987 < and \ref{tab:rotation} to compare to the results calculated from NVE
988 < simulations. The theoretical values for sphere is calculated from the
989 < Stokes-Einstein law, the theoretical values for ellipsoid is
990 < calculated from Perrin's fomula,  The exact method is
991 < applied to the langevin dynamics simulations for sphere and ellipsoid,
992 < the bead model is applied to the simulation for dumbbell molecule, and
993 < the rough shell model is applied to ellipsoid, dumbbell, banana and
994 < lipid molecules. The results from all the langevin dynamics
995 < simulations, including exact, bead model and rough shell, match the
996 < theoretical values perfectly for all different shaped molecules. This
997 < indicates that our simulation package for langevin dynamics is working
998 < well. The approxiate methods ( bead model and rough shell model) are
999 < accurate enough for the current simulations. The goal of the langevin
1000 < dynamics theory is to replace the explicit solvents by the friction
1001 < forces. We compared the dynamic properties of different shaped
1002 < molecules in langevin dynamics simulations with that in NVE
1003 < simulations. The results are reasonable close. Overall, the
1004 < translational diffusion constants calculated from langevin dynamics
1005 < simulations are very close to the values from the NVE simulation. For
1006 < sphere and lipid molecules, the diffusion constants are a little bit
1007 < off from the NVE simulation results. One possible reason is that the
1008 < calculation of the viscosity is very difficult to be accurate. Another
1009 < possible reason is that although we save very frequently during the
1010 < NVE simulations and run pretty long time simulations, there is only
1011 < one solute molecule in the system which makes the calculation for the
1012 < diffusion constant difficult. The sphere molecule behaves as a free
1013 < rotor in the solvent, so there is no rotation relaxation time
1014 < calculated from NVE simulations. The rotation relaxation time is not
1015 < very close to the NVE simulations results. The banana and lipid
1016 < molecules match the NVE simulations results pretty well. The mismatch
1017 < between langevin dynamics and NVE simulation for ellipsoid is possibly
1018 < caused by the slip boundary condition. For dumbbell, the mismatch is
1019 < caused by the size of the solvent molecule is pretty large compared to
1020 < dumbbell molecule in NVE simulations.
1021 <
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 1085 | 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  
1098 \subsection{Langevin Dynamics of Banana Shaped Molecules}
1099
1100 In order to verify that Langevin dynamics can mimic the dynamics of
1101 the systems absent of explicit solvents, we carried out two sets of
1102 simulations and compare their dynamic properties.
1103 Fig.~\ref{langevin:twoBanana} shows a snapshot of the simulation
1104 made of 256 pentane molecules and two banana shaped molecules at
1105 273~K. It has an equivalent implicit solvent system containing only
1106 two banana shaped molecules with viscosity of 0.289 center poise. To
1107 calculate the hydrodynamic properties of the banana shaped molecule,
1108 we created a rough shell model (see Fig.~\ref{langevin:roughShell}),
1109 in which the banana shaped molecule is represented as a ``shell''
1110 made of 2266 small identical beads with size of 0.3 \AA on the
1111 surface. Applying the procedure described in
1112 Sec.~\ref{introEquation:ResistanceTensorArbitraryOrigin}, we
1113 identified the center of resistance at (0 $\rm{\AA}$, 0.7482 $\rm{\AA}$,
1114 -0.1988 $\rm{\AA}$), as well as the resistance tensor,
1115 \[
1116 \left( {\begin{array}{*{20}c}
1117 0.9261 & 0 & 0&0&0.08585&0.2057\\
1118 0& 0.9270&-0.007063& 0.08585&0&0\\
1119 0&-0.007063&0.7494&0.2057&0&0\\
1120 0&0.0858&0.2057& 58.64& 0&0\\
1121 0.08585&0&0&0&48.30&3.219&\\
1122 0.2057&0&0&0&3.219&10.7373\\
1123 \end{array}} \right).
1124 \]
1125 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.
1126 Curves of the velocity auto-correlation functions in
1127 Fig.~\ref{langevin:vacf} were shown to match each other very well.
1128 However, because of the stochastic nature, simulation using Langevin
1129 dynamics was shown to decay slightly faster than MD. In order to
1130 study the rotational motion of the molecules, we also calculated the
1131 auto-correlation function of the principle axis of the second GB
1132 particle, $u$. The discrepancy shown in Fig.~\ref{langevin:uacf} was
1133 probably due to the reason that we used the experimental viscosity directly instead of calculating bulk viscosity from simulation.
1134
1154   \begin{figure}
1155   \centering
1156 < \includegraphics[width=\linewidth]{roughShell}
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  
1142 \begin{figure}
1143 \centering
1144 \includegraphics[width=\linewidth]{twoBanana}
1145 \caption[Snapshot from Simulation of Two Banana Shaped Molecules and
1146 256 Pentane Molecules]{Snapshot from simulation of two Banana shaped
1147 molecules and 256 pentane molecules.} \label{langevin:twoBanana}
1148 \end{figure}
1149
1150 \begin{figure}
1151 \centering
1152 \includegraphics[width=\linewidth]{vacf}
1153 \caption[Plots of Velocity Auto-correlation Functions]{Velocity
1154 auto-correlation functions of NVE (explicit solvent) in blue and
1155 Langevin dynamics (implicit solvent) in red.} \label{langevin:vacf}
1156 \end{figure}
1157
1158 \begin{figure}
1159 \centering
1160 \includegraphics[width=\linewidth]{uacf}
1161 \caption[Auto-correlation functions of the principle axis of the
1162 middle GB particle]{Auto-correlation functions of the principle axis
1163 of the middle GB particle of NVE (blue) and Langevin dynamics
1164 (red).} \label{langevin:uacf}
1165 \end{figure}
1166
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
1175 < Nos\'{e}-Hoover method. Further studies in systems involving banana
1176 < shaped molecules illustrated that the dynamic properties could be
1177 < 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}

Diff Legend

Removed lines
+ Added lines
< Changed lines
> Changed lines