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

Comparing trunk/tengDissertation/Introduction.tex (file contents):
Revision 2702 by tim, Wed Apr 12 03:37:46 2006 UTC vs.
Revision 2801 by tim, Tue Jun 6 14:56:36 2006 UTC

# Line 93 | Line 93 | the kinetic, $K$, and potential energies, $U$ \cite{to
93   The actual trajectory, along which a dynamical system may move from
94   one point to another within a specified time, is derived by finding
95   the path which minimizes the time integral of the difference between
96 < the kinetic, $K$, and potential energies, $U$ \cite{tolman79}.
96 > the kinetic, $K$, and potential energies, $U$ \cite{Tolman1979}.
97   \begin{equation}
98   \delta \int_{t_1 }^{t_2 } {(K - U)dt = 0} ,
99   \label{introEquation:halmitonianPrinciple1}
# Line 189 | Line 189 | known as the canonical equations of motions \cite{Gold
189   Eq.~\ref{introEquation:motionHamiltonianCoordinate} and
190   Eq.~\ref{introEquation:motionHamiltonianMomentum} are Hamilton's
191   equation of motion. Due to their symmetrical formula, they are also
192 < known as the canonical equations of motions \cite{Goldstein01}.
192 > known as the canonical equations of motions \cite{Goldstein2001}.
193  
194   An important difference between Lagrangian approach and the
195   Hamiltonian approach is that the Lagrangian is considered to be a
# Line 200 | Line 200 | equations\cite{Marion90}.
200   appropriate for application to statistical mechanics and quantum
201   mechanics, since it treats the coordinate and its time derivative as
202   independent variables and it only works with 1st-order differential
203 < equations\cite{Marion90}.
203 > equations\cite{Marion1990}.
204  
205   In Newtonian Mechanics, a system described by conservative forces
206   conserves the total energy \ref{introEquation:energyConservation}.
# Line 315 | Line 315 | partition function like,
315   isolated and conserve energy, Microcanonical ensemble(NVE) has a
316   partition function like,
317   \begin{equation}
318 < \Omega (N,V,E) = e^{\beta TS}
319 < \label{introEqaution:NVEPartition}.
318 > \Omega (N,V,E) = e^{\beta TS} \label{introEquation:NVEPartition}.
319   \end{equation}
320   A canonical ensemble(NVT)is an ensemble of systems, each of which
321   can share its energy with a large heat reservoir. The distribution
# Line 471 | Line 470 | statistical ensemble are identical \cite{Frenkel1996,
470   many-body system in Statistical Mechanics. Fortunately, Ergodic
471   Hypothesis is proposed to make a connection between time average and
472   ensemble average. It states that time average and average over the
473 < statistical ensemble are identical \cite{Frenkel1996, leach01:mm}.
473 > statistical ensemble are identical \cite{Frenkel1996, Leach2001}.
474   \begin{equation}
475   \langle A(q , p) \rangle_t = \mathop {\lim }\limits_{t \to \infty }
476   \frac{1}{t}\int\limits_0^t {A(q(t),p(t))dt = \int\limits_\Gamma
# Line 485 | Line 484 | reasonable, the Monte Carlo techniques\cite{metropolis
484   a properly weighted statistical average. This allows the researcher
485   freedom of choice when deciding how best to measure a given
486   observable. In case an ensemble averaged approach sounds most
487 < reasonable, the Monte Carlo techniques\cite{metropolis:1949} can be
487 > reasonable, the Monte Carlo techniques\cite{Metropolis1949} can be
488   utilized. Or if the system lends itself to a time averaging
489   approach, the Molecular Dynamics techniques in
490   Sec.~\ref{introSection:molecularDynamics} will be the best
# Line 499 | Line 498 | issue. The velocity verlet method, which happens to be
498   within the equations. Since 1990, geometric integrators, which
499   preserve various phase-flow invariants such as symplectic structure,
500   volume and time reversal symmetry, are developed to address this
501 < issue. The velocity verlet method, which happens to be a simple
502 < example of symplectic integrator, continues to gain its popularity
503 < in molecular dynamics community. This fact can be partly explained
504 < by its geometric nature.
501 > issue\cite{Dullweber1997, McLachlan1998, Leimkuhler1999}. The
502 > velocity verlet method, which happens to be a simple example of
503 > symplectic integrator, continues to gain its popularity in molecular
504 > dynamics community. This fact can be partly explained by its
505 > geometric nature.
506  
507   \subsection{\label{introSection:symplecticManifold}Symplectic Manifold}
508   A \emph{manifold} is an abstract mathematical space. It locally
# Line 566 | Line 566 | Another generalization of Hamiltonian dynamics is Pois
566   \end{equation}In this case, $f$ is
567   called a \emph{Hamiltonian vector field}.
568  
569 < Another generalization of Hamiltonian dynamics is Poisson Dynamics,
569 > Another generalization of Hamiltonian dynamics is Poisson
570 > Dynamics\cite{Olver1986},
571   \begin{equation}
572   \dot x = J(x)\nabla _x H \label{introEquation:poissonHamiltonian}
573   \end{equation}
574   The most obvious change being that matrix $J$ now depends on $x$.
574 The free rigid body is an example of Poisson system (actually a
575 Lie-Poisson system) with Hamiltonian function of angular kinetic
576 energy.
577 \begin{equation}
578 J(\pi ) = \left( {\begin{array}{*{20}c}
579   0 & {\pi _3 } & { - \pi _2 }  \\
580   { - \pi _3 } & 0 & {\pi _1 }  \\
581   {\pi _2 } & { - \pi _1 } & 0  \\
582 \end{array}} \right)
583 \end{equation}
575  
585 \begin{equation}
586 H = \frac{1}{2}\left( {\frac{{\pi _1^2 }}{{I_1 }} + \frac{{\pi _2^2
587 }}{{I_2 }} + \frac{{\pi _3^2 }}{{I_3 }}} \right)
588 \end{equation}
589
576   \subsection{\label{introSection:exactFlow}Exact Flow}
577  
578   Let $x(t)$ be the exact solution of the ODE system,
# Line 628 | Line 614 | The hidden geometric properties of ODE and its flow pl
614  
615   \subsection{\label{introSection:geometricProperties}Geometric Properties}
616  
617 < The hidden geometric properties of ODE and its flow play important
618 < roles in numerical studies. Many of them can be found in systems
619 < which occur naturally in applications.
617 > The hidden geometric properties\cite{Budd1999, Marsden1998} of ODE
618 > and its flow play important roles in numerical studies. Many of them
619 > can be found in systems which occur naturally in applications.
620  
621   Let $\varphi$ be the flow of Hamiltonian vector field, $\varphi$ is
622   a \emph{symplectic} flow if it satisfies,
623   \begin{equation}
624 < '\varphi^T J '\varphi = J.
624 > {\varphi '}^T J \varphi ' = J.
625   \end{equation}
626   According to Liouville's theorem, the symplectic volume is invariant
627   under a Hamiltonian flow, which is the basis for classical
# Line 643 | Line 629 | symplectomorphism. As to the Poisson system,
629   field on a symplectic manifold can be shown to be a
630   symplectomorphism. As to the Poisson system,
631   \begin{equation}
632 < '\varphi ^T J '\varphi  = J \circ \varphi
632 > {\varphi '}^T J \varphi ' = J \circ \varphi
633   \end{equation}
634   is the property must be preserved by the integrator.
635  
# Line 660 | Line 646 | only if $ h \circ \varphi ^{ - 1}  = \varphi  \circ h
646   The vector filed $f$ has reversing symmetry $h$ if $f = - \tilde f$.
647   In other words, the flow of this vector field is reversible if and
648   only if $ h \circ \varphi ^{ - 1}  = \varphi  \circ h $.
649 +
650 + A \emph{first integral}, or conserved quantity of a general
651 + differential function is a function $ G:R^{2d}  \to R^d $ which is
652 + constant for all solutions of the ODE $\frac{{dx}}{{dt}} = f(x)$ ,
653 + \[
654 + \frac{{dG(x(t))}}{{dt}} = 0.
655 + \]
656 + Using chain rule, one may obtain,
657 + \[
658 + \sum\limits_i {\frac{{dG}}{{dx_i }}} f_i (x) = f \bullet \nabla G,
659 + \]
660 + which is the condition for conserving \emph{first integral}. For a
661 + canonical Hamiltonian system, the time evolution of an arbitrary
662 + smooth function $G$ is given by,
663  
664 + \begin{eqnarray}
665 + \frac{{dG(x(t))}}{{dt}} & = & [\nabla _x G(x(t))]^T \dot x(t) \\
666 +                        & = & [\nabla _x G(x(t))]^T J\nabla _x H(x(t)). \\
667 + \label{introEquation:firstIntegral1}
668 + \end{eqnarray}
669 +
670 +
671 + Using poisson bracket notion, Equation
672 + \ref{introEquation:firstIntegral1} can be rewritten as
673 + \[
674 + \frac{d}{{dt}}G(x(t)) = \left\{ {G,H} \right\}(x(t)).
675 + \]
676 + Therefore, the sufficient condition for $G$ to be the \emph{first
677 + integral} of a Hamiltonian system is
678 + \[
679 + \left\{ {G,H} \right\} = 0.
680 + \]
681 + As well known, the Hamiltonian (or energy) H of a Hamiltonian system
682 + is a \emph{first integral}, which is due to the fact $\{ H,H\}  =
683 + 0$.
684 +
685   When designing any numerical methods, one should always try to
686   preserve the structural properties of the original ODE and its flow.
687  
# Line 678 | Line 699 | Generating function tends to lead to methods which are
699   \item Splitting methods
700   \end{enumerate}
701  
702 < Generating function tends to lead to methods which are cumbersome
703 < and difficult to use. In dissipative systems, variational methods
704 < can capture the decay of energy accurately. Since their
705 < geometrically unstable nature against non-Hamiltonian perturbations,
706 < ordinary implicit Runge-Kutta methods are not suitable for
707 < Hamiltonian system. Recently, various high-order explicit
708 < Runge--Kutta methods have been developed to overcome this
709 < instability \cite{}. However, due to computational penalty involved
710 < in implementing the Runge-Kutta methods, they do not attract too
711 < much attention from Molecular Dynamics community. Instead, splitting
712 < have been widely accepted since they exploit natural decompositions
713 < of the system\cite{Tuckerman92}.
702 > Generating function\cite{Channell1990} tends to lead to methods
703 > which are cumbersome and difficult to use. In dissipative systems,
704 > variational methods can capture the decay of energy
705 > accurately\cite{Kane2000}. Since their geometrically unstable nature
706 > against non-Hamiltonian perturbations, ordinary implicit Runge-Kutta
707 > methods are not suitable for Hamiltonian system. Recently, various
708 > high-order explicit Runge-Kutta methods
709 > \cite{Owren1992,Chen2003}have been developed to overcome this
710 > instability. However, due to computational penalty involved in
711 > implementing the Runge-Kutta methods, they do not attract too much
712 > attention from Molecular Dynamics community. Instead, splitting have
713 > been widely accepted since they exploit natural decompositions of
714 > the system\cite{Tuckerman1992, McLachlan1998}.
715  
716   \subsubsection{\label{introSection:splittingMethod}Splitting Method}
717  
# Line 736 | Line 758 | _{1,h/2} ,
758   splitting gives a second-order decomposition,
759   \begin{equation}
760   \varphi _h  = \varphi _{1,h/2}  \circ \varphi _{2,h}  \circ \varphi
761 < _{1,h/2} ,
740 < \label{introEqaution:secondOrderSplitting}
761 > _{1,h/2} , \label{introEquation:secondOrderSplitting}
762   \end{equation}
763   which has a local error proportional to $h^3$. Sprang splitting's
764   popularity in molecular simulation community attribute to its
765   symmetric property,
766   \begin{equation}
767   \varphi _h^{ - 1} = \varphi _{ - h}.
768 < \lable{introEquation:timeReversible}
768 > \label{introEquation:timeReversible}
769   \end{equation}
770  
771   \subsubsection{\label{introSection:exampleSplittingMethod}Example of Splitting Method}
# Line 802 | Line 823 | q(\Delta t) = q(0) + \frac{{\Delta t}}{2}\left[ {\dot
823   \frac{{\Delta t}}{{2m}}\dot q(0)} \right], %
824   \label{introEquation:positionVerlet1} \\%
825   %
826 < q(\Delta t) = q(0) + \frac{{\Delta t}}{2}\left[ {\dot q(0) + \dot
826 > q(\Delta t) &= q(0) + \frac{{\Delta t}}{2}\left[ {\dot q(0) + \dot
827   q(\Delta t)} \right]. %
828 < \label{introEquation:positionVerlet1}
828 > \label{introEquation:positionVerlet2}
829   \end{align}
830  
831   \subsubsection{\label{introSection:errorAnalysis}Error Analysis and Higher Order Methods}
# Line 813 | Line 834 | $\varphi_1(t)$ and $\varphi_2(t$ respectively , we hav
834   error of splitting method in terms of commutator of the
835   operators(\ref{introEquation:exponentialOperator}) associated with
836   the sub-flow. For operators $hX$ and $hY$ which are associate to
837 < $\varphi_1(t)$ and $\varphi_2(t$ respectively , we have
837 > $\varphi_1(t)$ and $\varphi_2(t)$ respectively , we have
838   \begin{equation}
839   \exp (hX + hY) = \exp (hZ)
840   \end{equation}
# Line 826 | Line 847 | Applying Baker-Campbell-Hausdorff formula to Sprang sp
847   \[
848   [X,Y] = XY - YX .
849   \]
850 < Applying Baker-Campbell-Hausdorff formula to Sprang splitting, we
851 < can obtain
852 < \begin{eqnarray}
853 < \exp (h X/2)\exp (h Y)\exp (h X/2) & = & \exp (h X + h Y + h^2
854 < [X,Y]/4 + h^2 [Y,X]/4 \\ & & \mbox{} + h^2 [X,X]/8 + h^2 [Y,Y]/8 +
855 < h^3 [Y,[Y,X]]/12 - h^3 [X,[X,Y]]/24 +  \ldots )
856 < \end{eqnarray}
850 > Applying Baker-Campbell-Hausdorff formula\cite{Varadarajan1974} to
851 > Sprang splitting, we can obtain
852 > \begin{eqnarray*}
853 > \exp (h X/2)\exp (h Y)\exp (h X/2) & = & \exp (h X + h Y + h^2 [X,Y]/4 + h^2 [Y,X]/4 \\
854 >                                   &   & \mbox{} + h^2 [X,X]/8 + h^2 [Y,Y]/8 \\
855 >                                   &   & \mbox{} + h^3 [Y,[Y,X]]/12 - h^3[X,[X,Y]]/24 + \ldots )
856 > \end{eqnarray*}
857   Since \[ [X,Y] + [Y,X] = 0\] and \[ [X,X] = 0\], the dominant local
858   error of Spring splitting is proportional to $h^3$. The same
859   procedure can be applied to general splitting,  of the form
# Line 840 | Line 861 | Careful choice of coefficient $a_1 ,\ldot , b_m$ will
861   \varphi _{b_m h}^2  \circ \varphi _{a_m h}^1  \circ \varphi _{b_{m -
862   1} h}^2  \circ  \ldots  \circ \varphi _{a_1 h}^1 .
863   \end{equation}
864 < Careful choice of coefficient $a_1 ,\ldot , b_m$ will lead to higher
864 > Careful choice of coefficient $a_1 \ldots b_m$ will lead to higher
865   order method. Yoshida proposed an elegant way to compose higher
866 < order methods based on symmetric splitting. Given a symmetric second
867 < order base method $ \varphi _h^{(2)} $, a fourth-order symmetric
868 < method can be constructed by composing,
866 > order methods based on symmetric splitting\cite{Yoshida1990}. Given
867 > a symmetric second order base method $ \varphi _h^{(2)} $, a
868 > fourth-order symmetric method can be constructed by composing,
869   \[
870   \varphi _h^{(4)}  = \varphi _{\alpha h}^{(2)}  \circ \varphi _{\beta
871   h}^{(2)}  \circ \varphi _{\alpha h}^{(2)}
# Line 864 | Line 885 | As a special discipline of molecular modeling, Molecul
885  
886   \section{\label{introSection:molecularDynamics}Molecular Dynamics}
887  
888 < As a special discipline of molecular modeling, Molecular dynamics
889 < has proven to be a powerful tool for studying the functions of
890 < biological systems, providing structural, thermodynamic and
891 < dynamical information.
888 > As one of the principal tools of molecular modeling, Molecular
889 > dynamics has proven to be a powerful tool for studying the functions
890 > of biological systems, providing structural, thermodynamic and
891 > dynamical information. The basic idea of molecular dynamics is that
892 > macroscopic properties are related to microscopic behavior and
893 > microscopic behavior can be calculated from the trajectories in
894 > simulations. For instance, instantaneous temperature of an
895 > Hamiltonian system of $N$ particle can be measured by
896 > \[
897 > T = \sum\limits_{i = 1}^N {\frac{{m_i v_i^2 }}{{fk_B }}}
898 > \]
899 > where $m_i$ and $v_i$ are the mass and velocity of $i$th particle
900 > respectively, $f$ is the number of degrees of freedom, and $k_B$ is
901 > the boltzman constant.
902  
903 < \subsection{\label{introSec:mdInit}Initialization}
903 > A typical molecular dynamics run consists of three essential steps:
904 > \begin{enumerate}
905 >  \item Initialization
906 >    \begin{enumerate}
907 >    \item Preliminary preparation
908 >    \item Minimization
909 >    \item Heating
910 >    \item Equilibration
911 >    \end{enumerate}
912 >  \item Production
913 >  \item Analysis
914 > \end{enumerate}
915 > These three individual steps will be covered in the following
916 > sections. Sec.~\ref{introSec:initialSystemSettings} deals with the
917 > initialization of a simulation. Sec.~\ref{introSection:production}
918 > will discusses issues in production run.
919 > Sec.~\ref{introSection:Analysis} provides the theoretical tools for
920 > trajectory analysis.
921  
922 < \subsection{\label{introSection:mdIntegration} Integration of the Equations of Motion}
922 > \subsection{\label{introSec:initialSystemSettings}Initialization}
923  
924 < \section{\label{introSection:rigidBody}Dynamics of Rigid Bodies}
924 > \subsubsection{Preliminary preparation}
925  
926 < A rigid body is a body in which the distance between any two given
927 < points of a rigid body remains constant regardless of external
928 < forces exerted on it. A rigid body therefore conserves its shape
929 < during its motion.
930 <
931 < Applications of dynamics of rigid bodies.
932 <
933 < \subsection{\label{introSection:lieAlgebra}Lie Algebra}
934 <
935 < \subsection{\label{introSection:DLMMotionEquation}The Euler Equations of Rigid Body Motion}
936 <
937 < \subsection{\label{introSection:otherRBMotionEquation}Other Formulations for Rigid Body Motion}
938 <
939 < \section{\label{introSection:correlationFunctions}Correlation Functions}
940 <
893 < \section{\label{introSection:langevinDynamics}Langevin Dynamics}
894 <
895 < \subsection{\label{introSection:LDIntroduction}Introduction and application of Langevin Dynamics}
926 > When selecting the starting structure of a molecule for molecular
927 > simulation, one may retrieve its Cartesian coordinates from public
928 > databases, such as RCSB Protein Data Bank \textit{etc}. Although
929 > thousands of crystal structures of molecules are discovered every
930 > year, many more remain unknown due to the difficulties of
931 > purification and crystallization. Even for the molecule with known
932 > structure, some important information is missing. For example, the
933 > missing hydrogen atom which acts as donor in hydrogen bonding must
934 > be added. Moreover, in order to include electrostatic interaction,
935 > one may need to specify the partial charges for individual atoms.
936 > Under some circumstances, we may even need to prepare the system in
937 > a special setup. For instance, when studying transport phenomenon in
938 > membrane system, we may prepare the lipids in bilayer structure
939 > instead of placing lipids randomly in solvent, since we are not
940 > interested in self-aggregation and it takes a long time to happen.
941  
942 < \subsection{\label{introSection:generalizedLangevinDynamics}Generalized Langevin Dynamics}
942 > \subsubsection{Minimization}
943  
944 + It is quite possible that some of molecules in the system from
945 + preliminary preparation may be overlapped with each other. This
946 + close proximity leads to high potential energy which consequently
947 + jeopardizes any molecular dynamics simulations. To remove these
948 + steric overlaps, one typically performs energy minimization to find
949 + a more reasonable conformation. Several energy minimization methods
950 + have been developed to exploit the energy surface and to locate the
951 + local minimum. While converging slowly near the minimum, steepest
952 + descent method is extremely robust when systems are far from
953 + harmonic. Thus, it is often used to refine structure from
954 + crystallographic data. Relied on the gradient or hessian, advanced
955 + methods like conjugate gradient and Newton-Raphson converge rapidly
956 + to a local minimum, while become unstable if the energy surface is
957 + far from quadratic. Another factor must be taken into account, when
958 + choosing energy minimization method, is the size of the system.
959 + Steepest descent and conjugate gradient can deal with models of any
960 + size. Because of the limit of computation power to calculate hessian
961 + matrix and insufficient storage capacity to store them, most
962 + Newton-Raphson methods can not be used with very large models.
963 +
964 + \subsubsection{Heating}
965 +
966 + Typically, Heating is performed by assigning random velocities
967 + according to a Gaussian distribution for a temperature. Beginning at
968 + a lower temperature and gradually increasing the temperature by
969 + assigning greater random velocities, we end up with setting the
970 + temperature of the system to a final temperature at which the
971 + simulation will be conducted. In heating phase, we should also keep
972 + the system from drifting or rotating as a whole. Equivalently, the
973 + net linear momentum and angular momentum of the system should be
974 + shifted to zero.
975 +
976 + \subsubsection{Equilibration}
977 +
978 + The purpose of equilibration is to allow the system to evolve
979 + spontaneously for a period of time and reach equilibrium. The
980 + procedure is continued until various statistical properties, such as
981 + temperature, pressure, energy, volume and other structural
982 + properties \textit{etc}, become independent of time. Strictly
983 + speaking, minimization and heating are not necessary, provided the
984 + equilibration process is long enough. However, these steps can serve
985 + as a means to arrive at an equilibrated structure in an effective
986 + way.
987 +
988 + \subsection{\label{introSection:production}Production}
989 +
990 + Production run is the most important step of the simulation, in
991 + which the equilibrated structure is used as a starting point and the
992 + motions of the molecules are collected for later analysis. In order
993 + to capture the macroscopic properties of the system, the molecular
994 + dynamics simulation must be performed in correct and efficient way.
995 +
996 + The most expensive part of a molecular dynamics simulation is the
997 + calculation of non-bonded forces, such as van der Waals force and
998 + Coulombic forces \textit{etc}. For a system of $N$ particles, the
999 + complexity of the algorithm for pair-wise interactions is $O(N^2 )$,
1000 + which making large simulations prohibitive in the absence of any
1001 + computation saving techniques.
1002 +
1003 + A natural approach to avoid system size issue is to represent the
1004 + bulk behavior by a finite number of the particles. However, this
1005 + approach will suffer from the surface effect. To offset this,
1006 + \textit{Periodic boundary condition} (see Fig.~\ref{introFig:pbc})
1007 + is developed to simulate bulk properties with a relatively small
1008 + number of particles. In this method, the simulation box is
1009 + replicated throughout space to form an infinite lattice. During the
1010 + simulation, when a particle moves in the primary cell, its image in
1011 + other cells move in exactly the same direction with exactly the same
1012 + orientation. Thus, as a particle leaves the primary cell, one of its
1013 + images will enter through the opposite face.
1014 + \begin{figure}
1015 + \centering
1016 + \includegraphics[width=\linewidth]{pbc.eps}
1017 + \caption[An illustration of periodic boundary conditions]{A 2-D
1018 + illustration of periodic boundary conditions. As one particle leaves
1019 + the left of the simulation box, an image of it enters the right.}
1020 + \label{introFig:pbc}
1021 + \end{figure}
1022 +
1023 + %cutoff and minimum image convention
1024 + Another important technique to improve the efficiency of force
1025 + evaluation is to apply cutoff where particles farther than a
1026 + predetermined distance, are not included in the calculation
1027 + \cite{Frenkel1996}. The use of a cutoff radius will cause a
1028 + discontinuity in the potential energy curve. Fortunately, one can
1029 + shift the potential to ensure the potential curve go smoothly to
1030 + zero at the cutoff radius. Cutoff strategy works pretty well for
1031 + Lennard-Jones interaction because of its short range nature.
1032 + However, simply truncating the electrostatic interaction with the
1033 + use of cutoff has been shown to lead to severe artifacts in
1034 + simulations. Ewald summation, in which the slowly conditionally
1035 + convergent Coulomb potential is transformed into direct and
1036 + reciprocal sums with rapid and absolute convergence, has proved to
1037 + minimize the periodicity artifacts in liquid simulations. Taking the
1038 + advantages of the fast Fourier transform (FFT) for calculating
1039 + discrete Fourier transforms, the particle mesh-based
1040 + methods\cite{Hockney1981,Shimada1993, Luty1994} are accelerated from
1041 + $O(N^{3/2})$ to $O(N logN)$. An alternative approach is \emph{fast
1042 + multipole method}\cite{Greengard1987, Greengard1994}, which treats
1043 + Coulombic interaction exactly at short range, and approximate the
1044 + potential at long range through multipolar expansion. In spite of
1045 + their wide acceptances at the molecular simulation community, these
1046 + two methods are hard to be implemented correctly and efficiently.
1047 + Instead, we use a damped and charge-neutralized Coulomb potential
1048 + method developed by Wolf and his coworkers\cite{Wolf1999}. The
1049 + shifted Coulomb potential for particle $i$ and particle $j$ at
1050 + distance $r_{rj}$ is given by:
1051   \begin{equation}
1052 < H = \frac{{p^2 }}{{2m}} + U(x) + H_B  + \Delta U(x,x_1 , \ldots x_N)
1053 < \label{introEquation:bathGLE}
1052 > V(r_{ij})= \frac{q_i q_j \textrm{erfc}(\alpha
1053 > r_{ij})}{r_{ij}}-\lim_{r_{ij}\rightarrow
1054 > R_\textrm{c}}\left\{\frac{q_iq_j \textrm{erfc}(\alpha
1055 > r_{ij})}{r_{ij}}\right\}. \label{introEquation:shiftedCoulomb}
1056   \end{equation}
1057 < where $H_B$ is harmonic bath Hamiltonian,
1058 < \[
1059 < H_B  =\sum\limits_{\alpha  = 1}^N {\left\{ {\frac{{p_\alpha ^2
1060 < }}{{2m_\alpha  }} + \frac{1}{2}m_\alpha  w_\alpha ^2 } \right\}}
1061 < \]
1062 < and $\Delta U$ is bilinear system-bath coupling,
1063 < \[
1064 < \Delta U =  - \sum\limits_{\alpha  = 1}^N {g_\alpha  x_\alpha  x}
1065 < \]
1066 < Completing the square,
913 < \[
914 < H_B  + \Delta U = \sum\limits_{\alpha  = 1}^N {\left\{
915 < {\frac{{p_\alpha ^2 }}{{2m_\alpha  }} + \frac{1}{2}m_\alpha
916 < w_\alpha ^2 \left( {x_\alpha   - \frac{{g_\alpha  }}{{m_\alpha
917 < w_\alpha ^2 }}x} \right)^2 } \right\}}  - \sum\limits_{\alpha  =
918 < 1}^N {\frac{{g_\alpha ^2 }}{{2m_\alpha  w_\alpha ^2 }}} x^2
919 < \]
920 < and putting it back into Eq.~\ref{introEquation:bathGLE},
921 < \[
922 < H = \frac{{p^2 }}{{2m}} + W(x) + \sum\limits_{\alpha  = 1}^N
923 < {\left\{ {\frac{{p_\alpha ^2 }}{{2m_\alpha  }} + \frac{1}{2}m_\alpha
924 < w_\alpha ^2 \left( {x_\alpha   - \frac{{g_\alpha  }}{{m_\alpha
925 < w_\alpha ^2 }}x} \right)^2 } \right\}}
926 < \]
927 < where
928 < \[
929 < W(x) = U(x) - \sum\limits_{\alpha  = 1}^N {\frac{{g_\alpha ^2
930 < }}{{2m_\alpha  w_\alpha ^2 }}} x^2
931 < \]
932 < Since the first two terms of the new Hamiltonian depend only on the
933 < system coordinates, we can get the equations of motion for
934 < Generalized Langevin Dynamics by Hamilton's equations
935 < \ref{introEquation:motionHamiltonianCoordinate,
936 < introEquation:motionHamiltonianMomentum},
937 < \begin{align}
938 < \dot p &=  - \frac{{\partial H}}{{\partial x}}
939 <       &= m\ddot x
940 <       &= - \frac{{\partial W(x)}}{{\partial x}} - \sum\limits_{\alpha  = 1}^N {g_\alpha  \left( {x_\alpha   - \frac{{g_\alpha  }}{{m_\alpha  w_\alpha ^2 }}x} \right)}
941 < \label{introEquation:Lp5}
942 < \end{align}
943 < , and
944 < \begin{align}
945 < \dot p_\alpha   &=  - \frac{{\partial H}}{{\partial x_\alpha  }}
946 <                &= m\ddot x_\alpha
947 <                &= \- m_\alpha  w_\alpha ^2 \left( {x_\alpha   - \frac{{g_\alpha}}{{m_\alpha  w_\alpha ^2 }}x} \right)
948 < \end{align}
1057 > where $\alpha$ is the convergence parameter. Due to the lack of
1058 > inherent periodicity and rapid convergence,this method is extremely
1059 > efficient and easy to implement.
1060 > \begin{figure}
1061 > \centering
1062 > \includegraphics[width=\linewidth]{shifted_coulomb.eps}
1063 > \caption[An illustration of shifted Coulomb potential]{An
1064 > illustration of shifted Coulomb potential.}
1065 > \label{introFigure:shiftedCoulomb}
1066 > \end{figure}
1067  
1068 < \subsection{\label{introSection:laplaceTransform}The Laplace Transform}
1068 > %multiple time step
1069  
1070 < \[
1071 < L(x) = \int_0^\infty  {x(t)e^{ - pt} dt}
1072 < \]
1073 <
1074 < \[
1075 < L(x + y) = L(x) + L(y)
1070 > \subsection{\label{introSection:Analysis} Analysis}
1071 >
1072 > Recently, advanced visualization technique are widely applied to
1073 > monitor the motions of molecules. Although the dynamics of the
1074 > system can be described qualitatively from animation, quantitative
1075 > trajectory analysis are more appreciable. According to the
1076 > principles of Statistical Mechanics,
1077 > Sec.~\ref{introSection:statisticalMechanics}, one can compute
1078 > thermodynamics properties, analyze fluctuations of structural
1079 > parameters, and investigate time-dependent processes of the molecule
1080 > from the trajectories.
1081 >
1082 > \subsubsection{\label{introSection:thermodynamicsProperties}Thermodynamics Properties}
1083 >
1084 > Thermodynamics properties, which can be expressed in terms of some
1085 > function of the coordinates and momenta of all particles in the
1086 > system, can be directly computed from molecular dynamics. The usual
1087 > way to measure the pressure is based on virial theorem of Clausius
1088 > which states that the virial is equal to $-3Nk_BT$. For a system
1089 > with forces between particles, the total virial, $W$, contains the
1090 > contribution from external pressure and interaction between the
1091 > particles:
1092 > \[
1093 > W =  - 3PV + \left\langle {\sum\limits_{i < j} {r{}_{ij} \cdot
1094 > f_{ij} } } \right\rangle
1095   \]
1096 + where $f_{ij}$ is the force between particle $i$ and $j$ at a
1097 + distance $r_{ij}$. Thus, the expression for the pressure is given
1098 + by:
1099 + \begin{equation}
1100 + P = \frac{{Nk_B T}}{V} - \frac{1}{{3V}}\left\langle {\sum\limits_{i
1101 + < j} {r{}_{ij} \cdot f_{ij} } } \right\rangle
1102 + \end{equation}
1103  
1104 + \subsubsection{\label{introSection:structuralProperties}Structural Properties}
1105 +
1106 + Structural Properties of a simple fluid can be described by a set of
1107 + distribution functions. Among these functions,\emph{pair
1108 + distribution function}, also known as \emph{radial distribution
1109 + function}, is of most fundamental importance to liquid-state theory.
1110 + Pair distribution function can be gathered by Fourier transforming
1111 + raw data from a series of neutron diffraction experiments and
1112 + integrating over the surface factor \cite{Powles1973}. The
1113 + experiment result can serve as a criterion to justify the
1114 + correctness of the theory. Moreover, various equilibrium
1115 + thermodynamic and structural properties can also be expressed in
1116 + terms of radial distribution function \cite{Allen1987}.
1117 +
1118 + A pair distribution functions $g(r)$ gives the probability that a
1119 + particle $i$ will be located at a distance $r$ from a another
1120 + particle $j$ in the system
1121   \[
1122 < L(ax) = aL(x)
1122 > g(r) = \frac{V}{{N^2 }}\left\langle {\sum\limits_i {\sum\limits_{j
1123 > \ne i} {\delta (r - r_{ij} )} } } \right\rangle.
1124   \]
1125 + Note that the delta function can be replaced by a histogram in
1126 + computer simulation. Figure
1127 + \ref{introFigure:pairDistributionFunction} shows a typical pair
1128 + distribution function for the liquid argon system. The occurrence of
1129 + several peaks in the plot of $g(r)$ suggests that it is more likely
1130 + to find particles at certain radial values than at others. This is a
1131 + result of the attractive interaction at such distances. Because of
1132 + the strong repulsive forces at short distance, the probability of
1133 + locating particles at distances less than about 2.5{\AA} from each
1134 + other is essentially zero.
1135  
1136 + %\begin{figure}
1137 + %\centering
1138 + %\includegraphics[width=\linewidth]{pdf.eps}
1139 + %\caption[Pair distribution function for the liquid argon
1140 + %]{Pair distribution function for the liquid argon}
1141 + %\label{introFigure:pairDistributionFunction}
1142 + %\end{figure}
1143 +
1144 + \subsubsection{\label{introSection:timeDependentProperties}Time-dependent
1145 + Properties}
1146 +
1147 + Time-dependent properties are usually calculated using \emph{time
1148 + correlation function}, which correlates random variables $A$ and $B$
1149 + at two different time
1150 + \begin{equation}
1151 + C_{AB} (t) = \left\langle {A(t)B(0)} \right\rangle.
1152 + \label{introEquation:timeCorrelationFunction}
1153 + \end{equation}
1154 + If $A$ and $B$ refer to same variable, this kind of correlation
1155 + function is called \emph{auto correlation function}. One example of
1156 + auto correlation function is velocity auto-correlation function
1157 + which is directly related to transport properties of molecular
1158 + liquids:
1159   \[
1160 < L(\dot x) = pL(x) - px(0)
1160 > D = \frac{1}{3}\int\limits_0^\infty  {\left\langle {v(t) \cdot v(0)}
1161 > \right\rangle } dt
1162   \]
1163 + where $D$ is diffusion constant. Unlike velocity autocorrelation
1164 + function which is averaging over time origins and over all the
1165 + atoms, dipole autocorrelation are calculated for the entire system.
1166 + The dipole autocorrelation function is given by:
1167 + \[
1168 + c_{dipole}  = \left\langle {u_{tot} (t) \cdot u_{tot} (t)}
1169 + \right\rangle
1170 + \]
1171 + Here $u_{tot}$ is the net dipole of the entire system and is given
1172 + by
1173 + \[
1174 + u_{tot} (t) = \sum\limits_i {u_i (t)}
1175 + \]
1176 + In principle, many time correlation functions can be related with
1177 + Fourier transforms of the infrared, Raman, and inelastic neutron
1178 + scattering spectra of molecular liquids. In practice, one can
1179 + extract the IR spectrum from the intensity of dipole fluctuation at
1180 + each frequency using the following relationship:
1181 + \[
1182 + \hat c_{dipole} (v) = \int_{ - \infty }^\infty  {c_{dipole} (t)e^{ -
1183 + i2\pi vt} dt}
1184 + \]
1185  
1186 + \section{\label{introSection:rigidBody}Dynamics of Rigid Bodies}
1187 +
1188 + Rigid bodies are frequently involved in the modeling of different
1189 + areas, from engineering, physics, to chemistry. For example,
1190 + missiles and vehicle are usually modeled by rigid bodies.  The
1191 + movement of the objects in 3D gaming engine or other physics
1192 + simulator is governed by the rigid body dynamics. In molecular
1193 + simulation, rigid body is used to simplify the model in
1194 + protein-protein docking study\cite{Gray2003}.
1195 +
1196 + It is very important to develop stable and efficient methods to
1197 + integrate the equations of motion of orientational degrees of
1198 + freedom. Euler angles are the nature choice to describe the
1199 + rotational degrees of freedom. However, due to its singularity, the
1200 + numerical integration of corresponding equations of motion is very
1201 + inefficient and inaccurate. Although an alternative integrator using
1202 + different sets of Euler angles can overcome this
1203 + difficulty\cite{Barojas1973}, the computational penalty and the lost
1204 + of angular momentum conservation still remain. A singularity free
1205 + representation utilizing quaternions was developed by Evans in
1206 + 1977\cite{Evans1977}. Unfortunately, this approach suffer from the
1207 + nonseparable Hamiltonian resulted from quaternion representation,
1208 + which prevents the symplectic algorithm to be utilized. Another
1209 + different approach is to apply holonomic constraints to the atoms
1210 + belonging to the rigid body. Each atom moves independently under the
1211 + normal forces deriving from potential energy and constraint forces
1212 + which are used to guarantee the rigidness. However, due to their
1213 + iterative nature, SHAKE and Rattle algorithm converge very slowly
1214 + when the number of constraint increases\cite{Ryckaert1977,
1215 + Andersen1983}.
1216 +
1217 + The break through in geometric literature suggests that, in order to
1218 + develop a long-term integration scheme, one should preserve the
1219 + symplectic structure of the flow. Introducing conjugate momentum to
1220 + rotation matrix $Q$ and re-formulating Hamiltonian's equation, a
1221 + symplectic integrator, RSHAKE\cite{Kol1997}, was proposed to evolve
1222 + the Hamiltonian system in a constraint manifold by iteratively
1223 + satisfying the orthogonality constraint $Q_T Q = 1$. An alternative
1224 + method using quaternion representation was developed by
1225 + Omelyan\cite{Omelyan1998}. However, both of these methods are
1226 + iterative and inefficient. In this section, we will present a
1227 + symplectic Lie-Poisson integrator for rigid body developed by
1228 + Dullweber and his coworkers\cite{Dullweber1997} in depth.
1229 +
1230 + \subsection{\label{introSection:constrainedHamiltonianRB}Constrained Hamiltonian for Rigid Body}
1231 + The motion of the rigid body is Hamiltonian with the Hamiltonian
1232 + function
1233 + \begin{equation}
1234 + H = \frac{1}{2}(p^T m^{ - 1} p) + \frac{1}{2}tr(PJ^{ - 1} P) +
1235 + V(q,Q) + \frac{1}{2}tr[(QQ^T  - 1)\Lambda ].
1236 + \label{introEquation:RBHamiltonian}
1237 + \end{equation}
1238 + Here, $q$ and $Q$  are the position and rotation matrix for the
1239 + rigid-body, $p$ and $P$  are conjugate momenta to $q$  and $Q$ , and
1240 + $J$, a diagonal matrix, is defined by
1241   \[
1242 < L(\ddot x) = p^2 L(x) - px(0) - \dot x(0)
1242 > I_{ii}^{ - 1}  = \frac{1}{2}\sum\limits_{i \ne j} {J_{jj}^{ - 1} }
1243   \]
1244 + where $I_{ii}$ is the diagonal element of the inertia tensor. This
1245 + constrained Hamiltonian equation subjects to a holonomic constraint,
1246 + \begin{equation}
1247 + Q^T Q = 1, \label{introEquation:orthogonalConstraint}
1248 + \end{equation}
1249 + which is used to ensure rotation matrix's orthogonality.
1250 + Differentiating \ref{introEquation:orthogonalConstraint} and using
1251 + Equation \ref{introEquation:RBMotionMomentum}, one may obtain,
1252 + \begin{equation}
1253 + Q^T PJ^{ - 1}  + J^{ - 1} P^T Q = 0 . \\
1254 + \label{introEquation:RBFirstOrderConstraint}
1255 + \end{equation}
1256  
1257 + Using Equation (\ref{introEquation:motionHamiltonianCoordinate},
1258 + \ref{introEquation:motionHamiltonianMomentum}), one can write down
1259 + the equations of motion,
1260 +
1261 + \begin{eqnarray}
1262 + \frac{{dq}}{{dt}} & = & \frac{p}{m} \label{introEquation:RBMotionPosition}\\
1263 + \frac{{dp}}{{dt}} & = & - \nabla _q V(q,Q) \label{introEquation:RBMotionMomentum}\\
1264 + \frac{{dQ}}{{dt}} & = & PJ^{ - 1}  \label{introEquation:RBMotionRotation}\\
1265 + \frac{{dP}}{{dt}} & = & - \nabla _Q V(q,Q) - 2Q\Lambda . \label{introEquation:RBMotionP}
1266 + \end{eqnarray}
1267 +
1268 + In general, there are two ways to satisfy the holonomic constraints.
1269 + We can use constraint force provided by lagrange multiplier on the
1270 + normal manifold to keep the motion on constraint space. Or we can
1271 + simply evolve the system in constraint manifold. These two methods
1272 + are proved to be equivalent. The holonomic constraint and equations
1273 + of motions define a constraint manifold for rigid body
1274   \[
1275 < L\left( {\int_0^t {g(t - \tau )h(\tau )d\tau } } \right) = G(p)H(p)
1275 > M = \left\{ {(Q,P):Q^T Q = 1,Q^T PJ^{ - 1}  + J^{ - 1} P^T Q = 0}
1276 > \right\}.
1277   \]
1278  
1279 < Some relatively important transformation,
1279 > Unfortunately, this constraint manifold is not the cotangent bundle
1280 > $T_{\star}SO(3)$. However, it turns out that under symplectic
1281 > transformation, the cotangent space and the phase space are
1282 > diffeomorphic. Introducing
1283   \[
1284 < L(\cos at) = \frac{p}{{p^2  + a^2 }}
1284 > \tilde Q = Q,\tilde P = \frac{1}{2}\left( {P - QP^T Q} \right),
1285   \]
1286 + the mechanical system subject to a holonomic constraint manifold $M$
1287 + can be re-formulated as a Hamiltonian system on the cotangent space
1288 + \[
1289 + T^* SO(3) = \left\{ {(\tilde Q,\tilde P):\tilde Q^T \tilde Q =
1290 + 1,\tilde Q^T \tilde PJ^{ - 1}  + J^{ - 1} P^T \tilde Q = 0} \right\}
1291 + \]
1292  
1293 + For a body fixed vector $X_i$ with respect to the center of mass of
1294 + the rigid body, its corresponding lab fixed vector $X_0^{lab}$  is
1295 + given as
1296 + \begin{equation}
1297 + X_i^{lab} = Q X_i + q.
1298 + \end{equation}
1299 + Therefore, potential energy $V(q,Q)$ is defined by
1300   \[
1301 < L(\sin at) = \frac{a}{{p^2  + a^2 }}
1301 > V(q,Q) = V(Q X_0 + q).
1302   \]
1303 + Hence, the force and torque are given by
1304 + \[
1305 + \nabla _q V(q,Q) = F(q,Q) = \sum\limits_i {F_i (q,Q)},
1306 + \]
1307 + and
1308 + \[
1309 + \nabla _Q V(q,Q) = F(q,Q)X_i^t
1310 + \]
1311 + respectively.
1312  
1313 + As a common choice to describe the rotation dynamics of the rigid
1314 + body, angular momentum on body frame $\Pi  = Q^t P$ is introduced to
1315 + rewrite the equations of motion,
1316 + \begin{equation}
1317 + \begin{array}{l}
1318 + \mathop \Pi \limits^ \bullet   = J^{ - 1} \Pi ^T \Pi  + Q^T \sum\limits_i {F_i (q,Q)X_i^T }  - \Lambda  \\
1319 + \mathop Q\limits^{{\rm{   }} \bullet }  = Q\Pi {\rm{ }}J^{ - 1}  \\
1320 + \end{array}
1321 + \label{introEqaution:RBMotionPI}
1322 + \end{equation}
1323 + , as well as holonomic constraints,
1324   \[
1325 < L(1) = \frac{1}{p}
1325 > \begin{array}{l}
1326 > \Pi J^{ - 1}  + J^{ - 1} \Pi ^t  = 0 \\
1327 > Q^T Q = 1 \\
1328 > \end{array}
1329   \]
1330  
1331 < First, the bath coordinates,
1331 > For a vector $v(v_1 ,v_2 ,v_3 ) \in R^3$ and a matrix $\hat v \in
1332 > so(3)^ \star$, the hat-map isomorphism,
1333 > \begin{equation}
1334 > v(v_1 ,v_2 ,v_3 ) \Leftrightarrow \hat v = \left(
1335 > {\begin{array}{*{20}c}
1336 >   0 & { - v_3 } & {v_2 }  \\
1337 >   {v_3 } & 0 & { - v_1 }  \\
1338 >   { - v_2 } & {v_1 } & 0  \\
1339 > \end{array}} \right),
1340 > \label{introEquation:hatmapIsomorphism}
1341 > \end{equation}
1342 > will let us associate the matrix products with traditional vector
1343 > operations
1344   \[
1345 < p^2 L(x_\alpha  ) - px_\alpha  (0) - \dot x_\alpha  (0) =  - \omega
992 < _\alpha ^2 L(x_\alpha  ) + \frac{{g_\alpha  }}{{\omega _\alpha
993 < }}L(x)
1345 > \hat vu = v \times u
1346   \]
1347 + Using \ref{introEqaution:RBMotionPI}, one can construct a skew
1348 + matrix,
1349 + \begin{equation}
1350 + (\mathop \Pi \limits^ \bullet   - \mathop \Pi \limits^ {\bullet  ^T}
1351 + ){\rm{ }} = {\rm{ }}(\Pi  - \Pi ^T ){\rm{ }}(J^{ - 1} \Pi  + \Pi J^{
1352 + - 1} ) + \sum\limits_i {[Q^T F_i (r,Q)X_i^T  - X_i F_i (r,Q)^T Q]} -
1353 + (\Lambda  - \Lambda ^T ) . \label{introEquation:skewMatrixPI}
1354 + \end{equation}
1355 + Since $\Lambda$ is symmetric, the last term of Equation
1356 + \ref{introEquation:skewMatrixPI} is zero, which implies the Lagrange
1357 + multiplier $\Lambda$ is absent from the equations of motion. This
1358 + unique property eliminate the requirement of iterations which can
1359 + not be avoided in other methods\cite{Kol1997, Omelyan1998}.
1360 +
1361 + Applying hat-map isomorphism, we obtain the equation of motion for
1362 + angular momentum on body frame
1363 + \begin{equation}
1364 + \dot \pi  = \pi  \times I^{ - 1} \pi  + \sum\limits_i {\left( {Q^T
1365 + F_i (r,Q)} \right) \times X_i }.
1366 + \label{introEquation:bodyAngularMotion}
1367 + \end{equation}
1368 + In the same manner, the equation of motion for rotation matrix is
1369 + given by
1370   \[
1371 < L(x_\alpha  ) = \frac{{\frac{{g_\alpha  }}{{\omega _\alpha  }}L(x) +
997 < px_\alpha  (0) + \dot x_\alpha  (0)}}{{p^2  + \omega _\alpha ^2 }}
1371 > \dot Q = Qskew(I^{ - 1} \pi )
1372   \]
1373 < Then, the system coordinates,
1374 < \begin{align}
1375 < mL(\ddot x) &=  - \frac{1}{p}\frac{{\partial W(x)}}{{\partial x}} -
1376 < \sum\limits_{\alpha  = 1}^N {\left\{ {\frac{{\frac{{g_\alpha
1377 < }}{{\omega _\alpha  }}L(x) + px_\alpha  (0) + \dot x_\alpha
1378 < (0)}}{{p^2  + \omega _\alpha ^2 }} - \frac{{g_\alpha ^2 }}{{m_\alpha
1379 < }}\omega _\alpha ^2 L(x)} \right\}}
1380 < %
1381 < &= - \frac{1}{p}\frac{{\partial W(x)}}{{\partial x}} -
1382 < \sum\limits_{\alpha  = 1}^N {\left\{ { - \frac{{g_\alpha ^2 }}{{m_\alpha  \omega _\alpha ^2 }}\frac{p}{{p^2  + \omega _\alpha ^2 }}pL(x)
1383 < - \frac{p}{{p^2  + \omega _\alpha ^2 }}g_\alpha  x_\alpha  (0)
1384 < - \frac{1}{{p^2  + \omega _\alpha ^2 }}g_\alpha  \dot x_\alpha  (0)} \right\}}
1385 < \end{align}
1386 < Then, the inverse transform,
1373 >
1374 > \subsection{\label{introSection:SymplecticFreeRB}Symplectic
1375 > Lie-Poisson Integrator for Free Rigid Body}
1376 >
1377 > If there is not external forces exerted on the rigid body, the only
1378 > contribution to the rotational is from the kinetic potential (the
1379 > first term of \ref{introEquation:bodyAngularMotion}). The free rigid
1380 > body is an example of Lie-Poisson system with Hamiltonian function
1381 > \begin{equation}
1382 > T^r (\pi ) = T_1 ^r (\pi _1 ) + T_2^r (\pi _2 ) + T_3^r (\pi _3 )
1383 > \label{introEquation:rotationalKineticRB}
1384 > \end{equation}
1385 > where $T_i^r (\pi _i ) = \frac{{\pi _i ^2 }}{{2I_i }}$ and
1386 > Lie-Poisson structure matrix,
1387 > \begin{equation}
1388 > J(\pi ) = \left( {\begin{array}{*{20}c}
1389 >   0 & {\pi _3 } & { - \pi _2 }  \\
1390 >   { - \pi _3 } & 0 & {\pi _1 }  \\
1391 >   {\pi _2 } & { - \pi _1 } & 0  \\
1392 > \end{array}} \right)
1393 > \end{equation}
1394 > Thus, the dynamics of free rigid body is governed by
1395 > \begin{equation}
1396 > \frac{d}{{dt}}\pi  = J(\pi )\nabla _\pi  T^r (\pi )
1397 > \end{equation}
1398 >
1399 > One may notice that each $T_i^r$ in Equation
1400 > \ref{introEquation:rotationalKineticRB} can be solved exactly. For
1401 > instance, the equations of motion due to $T_1^r$ are given by
1402 > \begin{equation}
1403 > \frac{d}{{dt}}\pi  = R_1 \pi ,\frac{d}{{dt}}Q = QR_1
1404 > \label{introEqaution:RBMotionSingleTerm}
1405 > \end{equation}
1406 > where
1407 > \[ R_1  = \left( {\begin{array}{*{20}c}
1408 >   0 & 0 & 0  \\
1409 >   0 & 0 & {\pi _1 }  \\
1410 >   0 & { - \pi _1 } & 0  \\
1411 > \end{array}} \right).
1412 > \]
1413 > The solutions of Equation \ref{introEqaution:RBMotionSingleTerm} is
1414 > \[
1415 > \pi (\Delta t) = e^{\Delta tR_1 } \pi (0),Q(\Delta t) =
1416 > Q(0)e^{\Delta tR_1 }
1417 > \]
1418 > with
1419 > \[
1420 > e^{\Delta tR_1 }  = \left( {\begin{array}{*{20}c}
1421 >   0 & 0 & 0  \\
1422 >   0 & {\cos \theta _1 } & {\sin \theta _1 }  \\
1423 >   0 & { - \sin \theta _1 } & {\cos \theta _1 }  \\
1424 > \end{array}} \right),\theta _1  = \frac{{\pi _1 }}{{I_1 }}\Delta t.
1425 > \]
1426 > To reduce the cost of computing expensive functions in $e^{\Delta
1427 > tR_1 }$, we can use Cayley transformation,
1428 > \[
1429 > e^{\Delta tR_1 }  \approx (1 - \Delta tR_1 )^{ - 1} (1 + \Delta tR_1
1430 > )
1431 > \]
1432 > The flow maps for $T_2^r$ and $T_3^r$ can be found in the same
1433 > manner.
1434  
1435 < \begin{align}
1436 < m\ddot x &=  - \frac{{\partial W(x)}}{{\partial x}} -
1435 > In order to construct a second-order symplectic method, we split the
1436 > angular kinetic Hamiltonian function can into five terms
1437 > \[
1438 > T^r (\pi ) = \frac{1}{2}T_1 ^r (\pi _1 ) + \frac{1}{2}T_2^r (\pi _2
1439 > ) + T_3^r (\pi _3 ) + \frac{1}{2}T_2^r (\pi _2 ) + \frac{1}{2}T_1 ^r
1440 > (\pi _1 )
1441 > \].
1442 > Concatenating flows corresponding to these five terms, we can obtain
1443 > an symplectic integrator,
1444 > \[
1445 > \varphi _{\Delta t,T^r }  = \varphi _{\Delta t/2,\pi _1 }  \circ
1446 > \varphi _{\Delta t/2,\pi _2 }  \circ \varphi _{\Delta t,\pi _3 }
1447 > \circ \varphi _{\Delta t/2,\pi _2 }  \circ \varphi _{\Delta t/2,\pi
1448 > _1 }.
1449 > \]
1450 >
1451 > The non-canonical Lie-Poisson bracket ${F, G}$ of two function
1452 > $F(\pi )$ and $G(\pi )$ is defined by
1453 > \[
1454 > \{ F,G\} (\pi ) = [\nabla _\pi  F(\pi )]^T J(\pi )\nabla _\pi  G(\pi
1455 > )
1456 > \]
1457 > If the Poisson bracket of a function $F$ with an arbitrary smooth
1458 > function $G$ is zero, $F$ is a \emph{Casimir}, which is the
1459 > conserved quantity in Poisson system. We can easily verify that the
1460 > norm of the angular momentum, $\parallel \pi
1461 > \parallel$, is a \emph{Casimir}. Let$ F(\pi ) = S(\frac{{\parallel
1462 > \pi \parallel ^2 }}{2})$ for an arbitrary function $ S:R \to R$ ,
1463 > then by the chain rule
1464 > \[
1465 > \nabla _\pi  F(\pi ) = S'(\frac{{\parallel \pi \parallel ^2
1466 > }}{2})\pi
1467 > \]
1468 > Thus $ [\nabla _\pi  F(\pi )]^T J(\pi ) =  - S'(\frac{{\parallel \pi
1469 > \parallel ^2 }}{2})\pi  \times \pi  = 0 $. This explicit
1470 > Lie-Poisson integrator is found to be extremely efficient and stable
1471 > which can be explained by the fact the small angle approximation is
1472 > used and the norm of the angular momentum is conserved.
1473 >
1474 > \subsection{\label{introSection:RBHamiltonianSplitting} Hamiltonian
1475 > Splitting for Rigid Body}
1476 >
1477 > The Hamiltonian of rigid body can be separated in terms of kinetic
1478 > energy and potential energy,
1479 > \[
1480 > H = T(p,\pi ) + V(q,Q)
1481 > \]
1482 > The equations of motion corresponding to potential energy and
1483 > kinetic energy are listed in the below table,
1484 > \begin{table}
1485 > \caption{Equations of motion due to Potential and Kinetic Energies}
1486 > \begin{center}
1487 > \begin{tabular}{|l|l|}
1488 >  \hline
1489 >  % after \\: \hline or \cline{col1-col2} \cline{col3-col4} ...
1490 >  Potential & Kinetic \\
1491 >  $\frac{{dq}}{{dt}} = \frac{p}{m}$ & $\frac{d}{{dt}}q = p$ \\
1492 >  $\frac{d}{{dt}}p =  - \frac{{\partial V}}{{\partial q}}$ & $ \frac{d}{{dt}}p = 0$ \\
1493 >  $\frac{d}{{dt}}Q = 0$ & $ \frac{d}{{dt}}Q = Qskew(I^{ - 1} j)$ \\
1494 >  $ \frac{d}{{dt}}\pi  = \sum\limits_i {\left( {Q^T F_i (r,Q)} \right) \times X_i }$ & $\frac{d}{{dt}}\pi  = \pi  \times I^{ - 1} \pi$\\
1495 >  \hline
1496 > \end{tabular}
1497 > \end{center}
1498 > \end{table}
1499 > A second-order symplectic method is now obtained by the
1500 > composition of the flow maps,
1501 > \[
1502 > \varphi _{\Delta t}  = \varphi _{\Delta t/2,V}  \circ \varphi
1503 > _{\Delta t,T}  \circ \varphi _{\Delta t/2,V}.
1504 > \]
1505 > Moreover, $\varphi _{\Delta t/2,V}$ can be divided into two
1506 > sub-flows which corresponding to force and torque respectively,
1507 > \[
1508 > \varphi _{\Delta t/2,V}  = \varphi _{\Delta t/2,F}  \circ \varphi
1509 > _{\Delta t/2,\tau }.
1510 > \]
1511 > Since the associated operators of $\varphi _{\Delta t/2,F} $ and
1512 > $\circ \varphi _{\Delta t/2,\tau }$ are commuted, the composition
1513 > order inside $\varphi _{\Delta t/2,V}$ does not matter.
1514 >
1515 > Furthermore, kinetic potential can be separated to translational
1516 > kinetic term, $T^t (p)$, and rotational kinetic term, $T^r (\pi )$,
1517 > \begin{equation}
1518 > T(p,\pi ) =T^t (p) + T^r (\pi ).
1519 > \end{equation}
1520 > where $ T^t (p) = \frac{1}{2}p^T m^{ - 1} p $ and $T^r (\pi )$ is
1521 > defined by \ref{introEquation:rotationalKineticRB}. Therefore, the
1522 > corresponding flow maps are given by
1523 > \[
1524 > \varphi _{\Delta t,T}  = \varphi _{\Delta t,T^t }  \circ \varphi
1525 > _{\Delta t,T^r }.
1526 > \]
1527 > Finally, we obtain the overall symplectic flow maps for free moving
1528 > rigid body
1529 > \begin{equation}
1530 > \begin{array}{c}
1531 > \varphi _{\Delta t}  = \varphi _{\Delta t/2,F}  \circ \varphi _{\Delta t/2,\tau }  \\
1532 >  \circ \varphi _{\Delta t,T^t }  \circ \varphi _{\Delta t/2,\pi _1 }  \circ \varphi _{\Delta t/2,\pi _2 }  \circ \varphi _{\Delta t,\pi _3 }  \circ \varphi _{\Delta t/2,\pi _2 }  \circ \varphi _{\Delta t/2,\pi _1 }  \\
1533 >  \circ \varphi _{\Delta t/2,\tau }  \circ \varphi _{\Delta t/2,F}  .\\
1534 > \end{array}
1535 > \label{introEquation:overallRBFlowMaps}
1536 > \end{equation}
1537 >
1538 > \section{\label{introSection:langevinDynamics}Langevin Dynamics}
1539 > As an alternative to newtonian dynamics, Langevin dynamics, which
1540 > mimics a simple heat bath with stochastic and dissipative forces,
1541 > has been applied in a variety of studies. This section will review
1542 > the theory of Langevin dynamics simulation. A brief derivation of
1543 > generalized Langevin equation will be given first. Follow that, we
1544 > will discuss the physical meaning of the terms appearing in the
1545 > equation as well as the calculation of friction tensor from
1546 > hydrodynamics theory.
1547 >
1548 > \subsection{\label{introSection:generalizedLangevinDynamics}Derivation of Generalized Langevin Equation}
1549 >
1550 > Harmonic bath model, in which an effective set of harmonic
1551 > oscillators are used to mimic the effect of a linearly responding
1552 > environment, has been widely used in quantum chemistry and
1553 > statistical mechanics. One of the successful applications of
1554 > Harmonic bath model is the derivation of Deriving Generalized
1555 > Langevin Dynamics. Lets consider a system, in which the degree of
1556 > freedom $x$ is assumed to couple to the bath linearly, giving a
1557 > Hamiltonian of the form
1558 > \begin{equation}
1559 > H = \frac{{p^2 }}{{2m}} + U(x) + H_B  + \Delta U(x,x_1 , \ldots x_N)
1560 > \label{introEquation:bathGLE}.
1561 > \end{equation}
1562 > Here $p$ is a momentum conjugate to $q$, $m$ is the mass associated
1563 > with this degree of freedom, $H_B$ is harmonic bath Hamiltonian,
1564 > \[
1565 > H_B  = \sum\limits_{\alpha  = 1}^N {\left\{ {\frac{{p_\alpha ^2
1566 > }}{{2m_\alpha  }} + \frac{1}{2}m_\alpha  \omega _\alpha ^2 }
1567 > \right\}}
1568 > \]
1569 > where the index $\alpha$ runs over all the bath degrees of freedom,
1570 > $\omega _\alpha$ are the harmonic bath frequencies, $m_\alpha$ are
1571 > the harmonic bath masses, and $\Delta U$ is bilinear system-bath
1572 > coupling,
1573 > \[
1574 > \Delta U =  - \sum\limits_{\alpha  = 1}^N {g_\alpha  x_\alpha  x}
1575 > \]
1576 > where $g_\alpha$ are the coupling constants between the bath and the
1577 > coordinate $x$. Introducing
1578 > \[
1579 > W(x) = U(x) - \sum\limits_{\alpha  = 1}^N {\frac{{g_\alpha ^2
1580 > }}{{2m_\alpha  w_\alpha ^2 }}} x^2
1581 > \] and combining the last two terms in Equation
1582 > \ref{introEquation:bathGLE}, we may rewrite the Harmonic bath
1583 > Hamiltonian as
1584 > \[
1585 > H = \frac{{p^2 }}{{2m}} + W(x) + \sum\limits_{\alpha  = 1}^N
1586 > {\left\{ {\frac{{p_\alpha ^2 }}{{2m_\alpha  }} + \frac{1}{2}m_\alpha
1587 > w_\alpha ^2 \left( {x_\alpha   - \frac{{g_\alpha  }}{{m_\alpha
1588 > w_\alpha ^2 }}x} \right)^2 } \right\}}
1589 > \]
1590 > Since the first two terms of the new Hamiltonian depend only on the
1591 > system coordinates, we can get the equations of motion for
1592 > Generalized Langevin Dynamics by Hamilton's equations
1593 > \ref{introEquation:motionHamiltonianCoordinate,
1594 > introEquation:motionHamiltonianMomentum},
1595 > \begin{equation}
1596 > m\ddot x =  - \frac{{\partial W(x)}}{{\partial x}} -
1597 > \sum\limits_{\alpha  = 1}^N {g_\alpha  \left( {x_\alpha   -
1598 > \frac{{g_\alpha  }}{{m_\alpha  w_\alpha ^2 }}x} \right)},
1599 > \label{introEquation:coorMotionGLE}
1600 > \end{equation}
1601 > and
1602 > \begin{equation}
1603 > m\ddot x_\alpha   =  - m_\alpha  w_\alpha ^2 \left( {x_\alpha   -
1604 > \frac{{g_\alpha  }}{{m_\alpha  w_\alpha ^2 }}x} \right).
1605 > \label{introEquation:bathMotionGLE}
1606 > \end{equation}
1607 >
1608 > In order to derive an equation for $x$, the dynamics of the bath
1609 > variables $x_\alpha$ must be solved exactly first. As an integral
1610 > transform which is particularly useful in solving linear ordinary
1611 > differential equations, Laplace transform is the appropriate tool to
1612 > solve this problem. The basic idea is to transform the difficult
1613 > differential equations into simple algebra problems which can be
1614 > solved easily. Then applying inverse Laplace transform, also known
1615 > as the Bromwich integral, we can retrieve the solutions of the
1616 > original problems.
1617 >
1618 > Let $f(t)$ be a function defined on $ [0,\infty ) $. The Laplace
1619 > transform of f(t) is a new function defined as
1620 > \[
1621 > L(f(t)) \equiv F(p) = \int_0^\infty  {f(t)e^{ - pt} dt}
1622 > \]
1623 > where  $p$ is real and  $L$ is called the Laplace Transform
1624 > Operator. Below are some important properties of Laplace transform
1625 >
1626 > \begin{eqnarray*}
1627 > L(x + y)  & = & L(x) + L(y) \\
1628 > L(ax)     & = & aL(x) \\
1629 > L(\dot x) & = & pL(x) - px(0) \\
1630 > L(\ddot x)& = & p^2 L(x) - px(0) - \dot x(0) \\
1631 > L\left( {\int_0^t {g(t - \tau )h(\tau )d\tau } } \right)& = & G(p)H(p) \\
1632 > \end{eqnarray*}
1633 >
1634 >
1635 > Applying Laplace transform to the bath coordinates, we obtain
1636 > \begin{eqnarray*}
1637 > p^2 L(x_\alpha  ) - px_\alpha  (0) - \dot x_\alpha  (0) & = & - \omega _\alpha ^2 L(x_\alpha  ) + \frac{{g_\alpha  }}{{\omega _\alpha  }}L(x) \\
1638 > L(x_\alpha  ) & = & \frac{{\frac{{g_\alpha  }}{{\omega _\alpha  }}L(x) + px_\alpha  (0) + \dot x_\alpha  (0)}}{{p^2  + \omega _\alpha ^2 }} \\
1639 > \end{eqnarray*}
1640 >
1641 > By the same way, the system coordinates become
1642 > \begin{eqnarray*}
1643 > mL(\ddot x) & = & - \frac{1}{p}\frac{{\partial W(x)}}{{\partial x}} \\
1644 >  & & \mbox{} - \sum\limits_{\alpha  = 1}^N {\left\{ { - \frac{{g_\alpha ^2 }}{{m_\alpha  \omega _\alpha ^2 }}\frac{p}{{p^2  + \omega _\alpha ^2 }}pL(x) - \frac{p}{{p^2  + \omega _\alpha ^2 }}g_\alpha  x_\alpha  (0) - \frac{1}{{p^2  + \omega _\alpha ^2 }}g_\alpha  \dot x_\alpha  (0)} \right\}}  \\
1645 > \end{eqnarray*}
1646 >
1647 > With the help of some relatively important inverse Laplace
1648 > transformations:
1649 > \[
1650 > \begin{array}{c}
1651 > L(\cos at) = \frac{p}{{p^2  + a^2 }} \\
1652 > L(\sin at) = \frac{a}{{p^2  + a^2 }} \\
1653 > L(1) = \frac{1}{p} \\
1654 > \end{array}
1655 > \]
1656 > , we obtain
1657 > \begin{eqnarray*}
1658 > m\ddot x & =  & - \frac{{\partial W(x)}}{{\partial x}} -
1659   \sum\limits_{\alpha  = 1}^N {\left\{ {\left( { - \frac{{g_\alpha ^2
1660   }}{{m_\alpha  \omega _\alpha ^2 }}} \right)\int_0^t {\cos (\omega
1661 < _\alpha  t)\dot x(t - \tau )d\tau  - \left[ {g_\alpha  x_\alpha  (0)
1662 < - \frac{{g_\alpha  }}{{m_\alpha  \omega _\alpha  }}} \right]\cos
1663 < (\omega _\alpha  t) - \frac{{g_\alpha  \dot x_\alpha  (0)}}{{\omega
1664 < _\alpha  }}\sin (\omega _\alpha  t)} } \right\}}
1665 < %
1666 < &= - \frac{{\partial W(x)}}{{\partial x}} - \int_0^t
1661 > _\alpha  t)\dot x(t - \tau )d\tau } } \right\}}  \\
1662 > & & + \sum\limits_{\alpha  = 1}^N {\left\{ {\left[ {g_\alpha
1663 > x_\alpha (0) - \frac{{g_\alpha  }}{{m_\alpha  \omega _\alpha  }}}
1664 > \right]\cos (\omega _\alpha  t) + \frac{{g_\alpha  \dot x_\alpha
1665 > (0)}}{{\omega _\alpha  }}\sin (\omega _\alpha  t)} \right\}}
1666 > \end{eqnarray*}
1667 > \begin{eqnarray*}
1668 > m\ddot x & = & - \frac{{\partial W(x)}}{{\partial x}} - \int_0^t
1669   {\sum\limits_{\alpha  = 1}^N {\left( { - \frac{{g_\alpha ^2
1670   }}{{m_\alpha  \omega _\alpha ^2 }}} \right)\cos (\omega _\alpha
1671 < t)\dot x(t - \tau )d} \tau }  + \sum\limits_{\alpha  = 1}^N {\left\{
1672 < {\left[ {g_\alpha  x_\alpha  (0) - \frac{{g_\alpha  }}{{m_\alpha
1673 < \omega _\alpha  }}} \right]\cos (\omega _\alpha  t) +
1674 < \frac{{g_\alpha  \dot x_\alpha  (0)}}{{\omega _\alpha  }}\sin
1675 < (\omega _\alpha  t)} \right\}}
1676 < \end{align}
1677 <
1671 > t)\dot x(t - \tau )d} \tau }  \\
1672 > & & + \sum\limits_{\alpha  = 1}^N {\left\{ {\left[ {g_\alpha
1673 > x_\alpha (0) - \frac{{g_\alpha }}{{m_\alpha \omega _\alpha  }}}
1674 > \right]\cos (\omega _\alpha  t) + \frac{{g_\alpha  \dot x_\alpha
1675 > (0)}}{{\omega _\alpha  }}\sin (\omega _\alpha  t)} \right\}}
1676 > \end{eqnarray*}
1677 > Introducing a \emph{dynamic friction kernel}
1678   \begin{equation}
1679 + \xi (t) = \sum\limits_{\alpha  = 1}^N {\left( { - \frac{{g_\alpha ^2
1680 + }}{{m_\alpha  \omega _\alpha ^2 }}} \right)\cos (\omega _\alpha  t)}
1681 + \label{introEquation:dynamicFrictionKernelDefinition}
1682 + \end{equation}
1683 + and \emph{a random force}
1684 + \begin{equation}
1685 + R(t) = \sum\limits_{\alpha  = 1}^N {\left( {g_\alpha  x_\alpha  (0)
1686 + - \frac{{g_\alpha ^2 }}{{m_\alpha  \omega _\alpha ^2 }}x(0)}
1687 + \right)\cos (\omega _\alpha  t)}  + \frac{{\dot x_\alpha
1688 + (0)}}{{\omega _\alpha  }}\sin (\omega _\alpha  t),
1689 + \label{introEquation:randomForceDefinition}
1690 + \end{equation}
1691 + the equation of motion can be rewritten as
1692 + \begin{equation}
1693   m\ddot x =  - \frac{{\partial W}}{{\partial x}} - \int_0^t {\xi
1694   (t)\dot x(t - \tau )d\tau }  + R(t)
1695   \label{introEuqation:GeneralizedLangevinDynamics}
1696   \end{equation}
1697 < %where $ {\xi (t)}$ is friction kernel, $R(t)$ is random force and
1698 < %$W$ is the potential of mean force. $W(x) =  - kT\ln p(x)$
1697 > which is known as the \emph{generalized Langevin equation}.
1698 >
1699 > \subsubsection{\label{introSection:randomForceDynamicFrictionKernel}Random Force and Dynamic Friction Kernel}
1700 >
1701 > One may notice that $R(t)$ depends only on initial conditions, which
1702 > implies it is completely deterministic within the context of a
1703 > harmonic bath. However, it is easy to verify that $R(t)$ is totally
1704 > uncorrelated to $x$ and $\dot x$,
1705   \[
1706 < \xi (t) = \sum\limits_{\alpha  = 1}^N {\left( { - \frac{{g_\alpha ^2
1707 < }}{{m_\alpha  \omega _\alpha ^2 }}} \right)\cos (\omega _\alpha  t)}
1706 > \begin{array}{l}
1707 > \left\langle {x(t)R(t)} \right\rangle  = 0, \\
1708 > \left\langle {\dot x(t)R(t)} \right\rangle  = 0. \\
1709 > \end{array}
1710   \]
1711 < For an infinite harmonic bath, we can use the spectral density and
1712 < an integral over frequencies.
1711 > This property is what we expect from a truly random process. As long
1712 > as the model, which is gaussian distribution in general, chosen for
1713 > $R(t)$ is a truly random process, the stochastic nature of the GLE
1714 > still remains.
1715  
1716 + %dynamic friction kernel
1717 + The convolution integral
1718   \[
1719 < R(t) = \sum\limits_{\alpha  = 1}^N {\left( {g_\alpha  x_\alpha  (0)
1049 < - \frac{{g_\alpha ^2 }}{{m_\alpha  \omega _\alpha ^2 }}x(0)}
1050 < \right)\cos (\omega _\alpha  t)}  + \frac{{\dot x_\alpha
1051 < (0)}}{{\omega _\alpha  }}\sin (\omega _\alpha  t)
1719 > \int_0^t {\xi (t)\dot x(t - \tau )d\tau }
1720   \]
1721 < The random forces depend only on initial conditions.
1721 > depends on the entire history of the evolution of $x$, which implies
1722 > that the bath retains memory of previous motions. In other words,
1723 > the bath requires a finite time to respond to change in the motion
1724 > of the system. For a sluggish bath which responds slowly to changes
1725 > in the system coordinate, we may regard $\xi(t)$ as a constant
1726 > $\xi(t) = \Xi_0$. Hence, the convolution integral becomes
1727 > \[
1728 > \int_0^t {\xi (t)\dot x(t - \tau )d\tau }  = \xi _0 (x(t) - x(0))
1729 > \]
1730 > and Equation \ref{introEuqation:GeneralizedLangevinDynamics} becomes
1731 > \[
1732 > m\ddot x =  - \frac{\partial }{{\partial x}}\left( {W(x) +
1733 > \frac{1}{2}\xi _0 (x - x_0 )^2 } \right) + R(t),
1734 > \]
1735 > which can be used to describe dynamic caging effect. The other
1736 > extreme is the bath that responds infinitely quickly to motions in
1737 > the system. Thus, $\xi (t)$ can be taken as a $delta$ function in
1738 > time:
1739 > \[
1740 > \xi (t) = 2\xi _0 \delta (t)
1741 > \]
1742 > Hence, the convolution integral becomes
1743 > \[
1744 > \int_0^t {\xi (t)\dot x(t - \tau )d\tau }  = 2\xi _0 \int_0^t
1745 > {\delta (t)\dot x(t - \tau )d\tau }  = \xi _0 \dot x(t),
1746 > \]
1747 > and Equation \ref{introEuqation:GeneralizedLangevinDynamics} becomes
1748 > \begin{equation}
1749 > m\ddot x =  - \frac{{\partial W(x)}}{{\partial x}} - \xi _0 \dot
1750 > x(t) + R(t) \label{introEquation:LangevinEquation}
1751 > \end{equation}
1752 > which is known as the Langevin equation. The static friction
1753 > coefficient $\xi _0$ can either be calculated from spectral density
1754 > or be determined by Stokes' law for regular shaped particles.A
1755 > briefly review on calculating friction tensor for arbitrary shaped
1756 > particles is given in Sec.~\ref{introSection:frictionTensor}.
1757  
1758   \subsubsection{\label{introSection:secondFluctuationDissipation}The Second Fluctuation Dissipation Theorem}
1759 < So we can define a new set of coordinates,
1759 >
1760 > Defining a new set of coordinates,
1761   \[
1762   q_\alpha  (t) = x_\alpha  (t) - \frac{1}{{m_\alpha  \omega _\alpha
1763   ^2 }}x(0)
1764 < \]
1765 < This makes
1764 > \],
1765 > we can rewrite $R(T)$ as
1766   \[
1767 < R(t) = \sum\limits_{\alpha  = 1}^N {g_\alpha  q_\alpha  (t)}
1767 > R(t) = \sum\limits_{\alpha  = 1}^N {g_\alpha  q_\alpha  (t)}.
1768   \]
1769   And since the $q$ coordinates are harmonic oscillators,
1770 +
1771 + \begin{eqnarray*}
1772 + \left\langle {q_\alpha ^2 } \right\rangle  & = & \frac{{kT}}{{m_\alpha  \omega _\alpha ^2 }} \\
1773 + \left\langle {q_\alpha  (t)q_\alpha  (0)} \right\rangle & = & \left\langle {q_\alpha ^2 (0)} \right\rangle \cos (\omega _\alpha  t) \\
1774 + \left\langle {q_\alpha  (t)q_\beta  (0)} \right\rangle & = &\delta _{\alpha \beta } \left\langle {q_\alpha  (t)q_\alpha  (0)} \right\rangle  \\
1775 + \left\langle {R(t)R(0)} \right\rangle & = & \sum\limits_\alpha  {\sum\limits_\beta  {g_\alpha  g_\beta  \left\langle {q_\alpha  (t)q_\beta  (0)} \right\rangle } }  \\
1776 +  & = &\sum\limits_\alpha  {g_\alpha ^2 \left\langle {q_\alpha ^2 (0)} \right\rangle \cos (\omega _\alpha  t)}  \\
1777 +  & = &kT\xi (t) \\
1778 + \end{eqnarray*}
1779 +
1780 + Thus, we recover the \emph{second fluctuation dissipation theorem}
1781 + \begin{equation}
1782 + \xi (t) = \left\langle {R(t)R(0)} \right\rangle
1783 + \label{introEquation:secondFluctuationDissipation}.
1784 + \end{equation}
1785 + In effect, it acts as a constraint on the possible ways in which one
1786 + can model the random force and friction kernel.
1787 +
1788 + \subsection{\label{introSection:frictionTensor} Friction Tensor}
1789 + Theoretically, the friction kernel can be determined using velocity
1790 + autocorrelation function. However, this approach become impractical
1791 + when the system become more and more complicate. Instead, various
1792 + approaches based on hydrodynamics have been developed to calculate
1793 + the friction coefficients. The friction effect is isotropic in
1794 + Equation, $\zeta$ can be taken as a scalar. In general, friction
1795 + tensor $\Xi$ is a $6\times 6$ matrix given by
1796   \[
1797 + \Xi  = \left( {\begin{array}{*{20}c}
1798 +   {\Xi _{}^{tt} } & {\Xi _{}^{rt} }  \\
1799 +   {\Xi _{}^{tr} } & {\Xi _{}^{rr} }  \\
1800 + \end{array}} \right).
1801 + \]
1802 + Here, $ {\Xi^{tt} }$ and $ {\Xi^{rr} }$ are translational friction
1803 + tensor and rotational resistance (friction) tensor respectively,
1804 + while ${\Xi^{tr} }$ is translation-rotation coupling tensor and $
1805 + {\Xi^{rt} }$ is rotation-translation coupling tensor. When a
1806 + particle moves in a fluid, it may experience friction force or
1807 + torque along the opposite direction of the velocity or angular
1808 + velocity,
1809 + \[
1810 + \left( \begin{array}{l}
1811 + F_R  \\
1812 + \tau _R  \\
1813 + \end{array} \right) =  - \left( {\begin{array}{*{20}c}
1814 +   {\Xi ^{tt} } & {\Xi ^{rt} }  \\
1815 +   {\Xi ^{tr} } & {\Xi ^{rr} }  \\
1816 + \end{array}} \right)\left( \begin{array}{l}
1817 + v \\
1818 + w \\
1819 + \end{array} \right)
1820 + \]
1821 + where $F_r$ is the friction force and $\tau _R$ is the friction
1822 + toque.
1823 +
1824 + \subsubsection{\label{introSection:resistanceTensorRegular}The Resistance Tensor for Regular Shape}
1825 +
1826 + For a spherical particle, the translational and rotational friction
1827 + constant can be calculated from Stoke's law,
1828 + \[
1829 + \Xi ^{tt}  = \left( {\begin{array}{*{20}c}
1830 +   {6\pi \eta R} & 0 & 0  \\
1831 +   0 & {6\pi \eta R} & 0  \\
1832 +   0 & 0 & {6\pi \eta R}  \\
1833 + \end{array}} \right)
1834 + \]
1835 + and
1836 + \[
1837 + \Xi ^{rr}  = \left( {\begin{array}{*{20}c}
1838 +   {8\pi \eta R^3 } & 0 & 0  \\
1839 +   0 & {8\pi \eta R^3 } & 0  \\
1840 +   0 & 0 & {8\pi \eta R^3 }  \\
1841 + \end{array}} \right)
1842 + \]
1843 + where $\eta$ is the viscosity of the solvent and $R$ is the
1844 + hydrodynamics radius.
1845 +
1846 + Other non-spherical shape, such as cylinder and ellipsoid
1847 + \textit{etc}, are widely used as reference for developing new
1848 + hydrodynamics theory, because their properties can be calculated
1849 + exactly. In 1936, Perrin extended Stokes's law to general ellipsoid,
1850 + also called a triaxial ellipsoid, which is given in Cartesian
1851 + coordinates by\cite{Perrin1934, Perrin1936}
1852 + \[
1853 + \frac{{x^2 }}{{a^2 }} + \frac{{y^2 }}{{b^2 }} + \frac{{z^2 }}{{c^2
1854 + }} = 1
1855 + \]
1856 + where the semi-axes are of lengths $a$, $b$, and $c$. Unfortunately,
1857 + due to the complexity of the elliptic integral, only the ellipsoid
1858 + with the restriction of two axes having to be equal, \textit{i.e.}
1859 + prolate($ a \ge b = c$) and oblate ($ a < b = c $), can be solved
1860 + exactly. Introducing an elliptic integral parameter $S$ for prolate,
1861 + \[
1862 + S = \frac{2}{{\sqrt {a^2  - b^2 } }}\ln \frac{{a + \sqrt {a^2  - b^2
1863 + } }}{b},
1864 + \]
1865 + and oblate,
1866 + \[
1867 + S = \frac{2}{{\sqrt {b^2  - a^2 } }}arctg\frac{{\sqrt {b^2  - a^2 }
1868 + }}{a}
1869 + \],
1870 + one can write down the translational and rotational resistance
1871 + tensors
1872 + \[
1873   \begin{array}{l}
1874 < \left\langle {q_\alpha  (t)q_\alpha  (0)} \right\rangle  = \left\langle {q_\alpha ^2 (0)} \right\rangle \cos (\omega _\alpha  t) \\
1875 < \left\langle {q_\alpha  (t)q_\beta  (0)} \right\rangle  = \delta _{\alpha \beta } \left\langle {q_\alpha  (t)q_\alpha  (0)} \right\rangle  \\
1876 < \end{array}
1874 > \Xi _a^{tt}  = 16\pi \eta \frac{{a^2  - b^2 }}{{(2a^2  - b^2 )S - 2a}} \\
1875 > \Xi _b^{tt}  = \Xi _c^{tt}  = 32\pi \eta \frac{{a^2  - b^2 }}{{(2a^2  - 3b^2 )S + 2a}} \\
1876 > \end{array},
1877   \]
1878 + and
1879 + \[
1880 + \begin{array}{l}
1881 + \Xi _a^{rr}  = \frac{{32\pi }}{3}\eta \frac{{(a^2  - b^2 )b^2 }}{{2a - b^2 S}} \\
1882 + \Xi _b^{rr}  = \Xi _c^{rr}  = \frac{{32\pi }}{3}\eta \frac{{(a^4  - b^4 )}}{{(2a^2  - b^2 )S - 2a}} \\
1883 + \end{array}.
1884 + \]
1885  
1886 < \begin{align}
1074 < \left\langle {R(t)R(0)} \right\rangle  &= \sum\limits_\alpha
1075 < {\sum\limits_\beta  {g_\alpha  g_\beta  \left\langle {q_\alpha
1076 < (t)q_\beta  (0)} \right\rangle } }
1077 < %
1078 < &= \sum\limits_\alpha  {g_\alpha ^2 \left\langle {q_\alpha ^2 (0)}
1079 < \right\rangle \cos (\omega _\alpha  t)}
1080 < %
1081 < &= kT\xi (t)
1082 < \end{align}
1886 > \subsubsection{\label{introSection:resistanceTensorRegularArbitrary}The Resistance Tensor for Arbitrary Shape}
1887  
1888 + Unlike spherical and other regular shaped molecules, there is not
1889 + analytical solution for friction tensor of any arbitrary shaped
1890 + rigid molecules. The ellipsoid of revolution model and general
1891 + triaxial ellipsoid model have been used to approximate the
1892 + hydrodynamic properties of rigid bodies. However, since the mapping
1893 + from all possible ellipsoidal space, $r$-space, to all possible
1894 + combination of rotational diffusion coefficients, $D$-space is not
1895 + unique\cite{Wegener1979} as well as the intrinsic coupling between
1896 + translational and rotational motion of rigid body, general ellipsoid
1897 + is not always suitable for modeling arbitrarily shaped rigid
1898 + molecule. A number of studies have been devoted to determine the
1899 + friction tensor for irregularly shaped rigid bodies using more
1900 + advanced method where the molecule of interest was modeled by
1901 + combinations of spheres(beads)\cite{Carrasco1999} and the
1902 + hydrodynamics properties of the molecule can be calculated using the
1903 + hydrodynamic interaction tensor. Let us consider a rigid assembly of
1904 + $N$ beads immersed in a continuous medium. Due to hydrodynamics
1905 + interaction, the ``net'' velocity of $i$th bead, $v'_i$ is different
1906 + than its unperturbed velocity $v_i$,
1907 + \[
1908 + v'_i  = v_i  - \sum\limits_{j \ne i} {T_{ij} F_j }
1909 + \]
1910 + where $F_i$ is the frictional force, and $T_{ij}$ is the
1911 + hydrodynamic interaction tensor. The friction force of $i$th bead is
1912 + proportional to its ``net'' velocity
1913   \begin{equation}
1914 < \xi (t) = \left\langle {R(t)R(0)} \right\rangle
1915 < \label{introEquation:secondFluctuationDissipation}
1914 > F_i  = \zeta _i v_i  - \zeta _i \sum\limits_{j \ne i} {T_{ij} F_j }.
1915 > \label{introEquation:tensorExpression}
1916   \end{equation}
1917 + This equation is the basis for deriving the hydrodynamic tensor. In
1918 + 1930, Oseen and Burgers gave a simple solution to Equation
1919 + \ref{introEquation:tensorExpression}
1920 + \begin{equation}
1921 + T_{ij}  = \frac{1}{{8\pi \eta r_{ij} }}\left( {I + \frac{{R_{ij}
1922 + R_{ij}^T }}{{R_{ij}^2 }}} \right).
1923 + \label{introEquation:oseenTensor}
1924 + \end{equation}
1925 + Here $R_{ij}$ is the distance vector between bead $i$ and bead $j$.
1926 + A second order expression for element of different size was
1927 + introduced by Rotne and Prager\cite{Rotne1969} and improved by
1928 + Garc\'{i}a de la Torre and Bloomfield\cite{Torre1977},
1929 + \begin{equation}
1930 + T_{ij}  = \frac{1}{{8\pi \eta R_{ij} }}\left[ {\left( {I +
1931 + \frac{{R_{ij} R_{ij}^T }}{{R_{ij}^2 }}} \right) + R\frac{{\sigma
1932 + _i^2  + \sigma _j^2 }}{{r_{ij}^2 }}\left( {\frac{I}{3} -
1933 + \frac{{R_{ij} R_{ij}^T }}{{R_{ij}^2 }}} \right)} \right].
1934 + \label{introEquation:RPTensorNonOverlapped}
1935 + \end{equation}
1936 + Both of the Equation \ref{introEquation:oseenTensor} and Equation
1937 + \ref{introEquation:RPTensorNonOverlapped} have an assumption $R_{ij}
1938 + \ge \sigma _i  + \sigma _j$. An alternative expression for
1939 + overlapping beads with the same radius, $\sigma$, is given by
1940 + \begin{equation}
1941 + T_{ij}  = \frac{1}{{6\pi \eta R_{ij} }}\left[ {\left( {1 -
1942 + \frac{2}{{32}}\frac{{R_{ij} }}{\sigma }} \right)I +
1943 + \frac{2}{{32}}\frac{{R_{ij} R_{ij}^T }}{{R_{ij} \sigma }}} \right]
1944 + \label{introEquation:RPTensorOverlapped}
1945 + \end{equation}
1946  
1947 < \section{\label{introSection:hydroynamics}Hydrodynamics}
1947 > To calculate the resistance tensor at an arbitrary origin $O$, we
1948 > construct a $3N \times 3N$ matrix consisting of $N \times N$
1949 > $B_{ij}$ blocks
1950 > \begin{equation}
1951 > B = \left( {\begin{array}{*{20}c}
1952 >   {B_{11} } &  \ldots  & {B_{1N} }  \\
1953 >    \vdots  &  \ddots  &  \vdots   \\
1954 >   {B_{N1} } &  \cdots  & {B_{NN} }  \\
1955 > \end{array}} \right),
1956 > \end{equation}
1957 > where $B_{ij}$ is given by
1958 > \[
1959 > B_{ij}  = \delta _{ij} \frac{I}{{6\pi \eta R}} + (1 - \delta _{ij}
1960 > )T_{ij}
1961 > \]
1962 > where $\delta _{ij}$ is Kronecker delta function. Inverting matrix
1963 > $B$, we obtain
1964  
1965 < \subsection{\label{introSection:frictionTensor} Friction Tensor}
1966 < \subsection{\label{introSection:analyticalApproach}Analytical
1967 < Approach}
1965 > \[
1966 > C = B^{ - 1}  = \left( {\begin{array}{*{20}c}
1967 >   {C_{11} } &  \ldots  & {C_{1N} }  \\
1968 >    \vdots  &  \ddots  &  \vdots   \\
1969 >   {C_{N1} } &  \cdots  & {C_{NN} }  \\
1970 > \end{array}} \right)
1971 > \]
1972 > , which can be partitioned into $N \times N$ $3 \times 3$ block
1973 > $C_{ij}$. With the help of $C_{ij}$ and skew matrix $U_i$
1974 > \[
1975 > U_i  = \left( {\begin{array}{*{20}c}
1976 >   0 & { - z_i } & {y_i }  \\
1977 >   {z_i } & 0 & { - x_i }  \\
1978 >   { - y_i } & {x_i } & 0  \\
1979 > \end{array}} \right)
1980 > \]
1981 > where $x_i$, $y_i$, $z_i$ are the components of the vector joining
1982 > bead $i$ and origin $O$. Hence, the elements of resistance tensor at
1983 > arbitrary origin $O$ can be written as
1984 > \begin{equation}
1985 > \begin{array}{l}
1986 > \Xi _{}^{tt}  = \sum\limits_i {\sum\limits_j {C_{ij} } } , \\
1987 > \Xi _{}^{tr}  = \Xi _{}^{rt}  = \sum\limits_i {\sum\limits_j {U_i C_{ij} } } , \\
1988 > \Xi _{}^{rr}  =  - \sum\limits_i {\sum\limits_j {U_i C_{ij} } } U_j  \\
1989 > \end{array}
1990 > \label{introEquation:ResistanceTensorArbitraryOrigin}
1991 > \end{equation}
1992  
1993 < \subsection{\label{introSection:approximationApproach}Approximation
1994 < Approach}
1993 > The resistance tensor depends on the origin to which they refer. The
1994 > proper location for applying friction force is the center of
1995 > resistance (reaction), at which the trace of rotational resistance
1996 > tensor, $ \Xi ^{rr}$ reaches minimum. Mathematically, the center of
1997 > resistance is defined as an unique point of the rigid body at which
1998 > the translation-rotation coupling tensor are symmetric,
1999 > \begin{equation}
2000 > \Xi^{tr}  = \left( {\Xi^{tr} } \right)^T
2001 > \label{introEquation:definitionCR}
2002 > \end{equation}
2003 > Form Equation \ref{introEquation:ResistanceTensorArbitraryOrigin},
2004 > we can easily find out that the translational resistance tensor is
2005 > origin independent, while the rotational resistance tensor and
2006 > translation-rotation coupling resistance tensor depend on the
2007 > origin. Given resistance tensor at an arbitrary origin $O$, and a
2008 > vector ,$r_{OP}(x_{OP}, y_{OP}, z_{OP})$, from $O$ to $P$, we can
2009 > obtain the resistance tensor at $P$ by
2010 > \begin{equation}
2011 > \begin{array}{l}
2012 > \Xi _P^{tt}  = \Xi _O^{tt}  \\
2013 > \Xi _P^{tr}  = \Xi _P^{rt}  = \Xi _O^{tr}  - U_{OP} \Xi _O^{tt}  \\
2014 > \Xi _P^{rr}  = \Xi _O^{rr}  - U_{OP} \Xi _O^{tt} U_{OP}  + \Xi _O^{tr} U_{OP}  - U_{OP} \Xi _O^{{tr} ^{^T }}  \\
2015 > \end{array}
2016 > \label{introEquation:resistanceTensorTransformation}
2017 > \end{equation}
2018 > where
2019 > \[
2020 > U_{OP}  = \left( {\begin{array}{*{20}c}
2021 >   0 & { - z_{OP} } & {y_{OP} }  \\
2022 >   {z_i } & 0 & { - x_{OP} }  \\
2023 >   { - y_{OP} } & {x_{OP} } & 0  \\
2024 > \end{array}} \right)
2025 > \]
2026 > Using Equations \ref{introEquation:definitionCR} and
2027 > \ref{introEquation:resistanceTensorTransformation}, one can locate
2028 > the position of center of resistance,
2029 > \begin{eqnarray*}
2030 > \left( \begin{array}{l}
2031 > x_{OR}  \\
2032 > y_{OR}  \\
2033 > z_{OR}  \\
2034 > \end{array} \right) & = &\left( {\begin{array}{*{20}c}
2035 >   {(\Xi _O^{rr} )_{yy}  + (\Xi _O^{rr} )_{zz} } & { - (\Xi _O^{rr} )_{xy} } & { - (\Xi _O^{rr} )_{xz} }  \\
2036 >   { - (\Xi _O^{rr} )_{xy} } & {(\Xi _O^{rr} )_{zz}  + (\Xi _O^{rr} )_{xx} } & { - (\Xi _O^{rr} )_{yz} }  \\
2037 >   { - (\Xi _O^{rr} )_{xz} } & { - (\Xi _O^{rr} )_{yz} } & {(\Xi _O^{rr} )_{xx}  + (\Xi _O^{rr} )_{yy} }  \\
2038 > \end{array}} \right)^{ - 1}  \\
2039 >  & & \left( \begin{array}{l}
2040 > (\Xi _O^{tr} )_{yz}  - (\Xi _O^{tr} )_{zy}  \\
2041 > (\Xi _O^{tr} )_{zx}  - (\Xi _O^{tr} )_{xz}  \\
2042 > (\Xi _O^{tr} )_{xy}  - (\Xi _O^{tr} )_{yx}  \\
2043 > \end{array} \right) \\
2044 > \end{eqnarray*}
2045  
2046 < \subsection{\label{introSection:centersRigidBody}Centers of Rigid
2047 < Body}
2046 >
2047 >
2048 > where $x_OR$, $y_OR$, $z_OR$ are the components of the vector
2049 > joining center of resistance $R$ and origin $O$.

Diff Legend

Removed lines
+ Added lines
< Changed lines
> Changed lines