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 2700 by tim, Tue Apr 11 03:38:09 2006 UTC vs.
Revision 2726 by tim, Fri Apr 21 15:37:53 2006 UTC

# Line 27 | Line 27 | $F_ij$ be the force that particle $i$ exerts on partic
27   \end{equation}
28   A point mass interacting with other bodies moves with the
29   acceleration along the direction of the force acting on it. Let
30 < $F_ij$ be the force that particle $i$ exerts on particle $j$, and
31 < $F_ji$ be the force that particle $j$ exerts on particle $i$.
30 > $F_{ij}$ be the force that particle $i$ exerts on particle $j$, and
31 > $F_{ji}$ be the force that particle $j$ exerts on particle $i$.
32   Newton¡¯s third law states that
33   \begin{equation}
34 < F_ij = -F_ji
34 > F_{ij} = -F_{ji}
35   \label{introEquation:newtonThirdLaw}
36   \end{equation}
37  
# 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 394 | Line 393 | distribution,
393   \begin{equation}
394   \rho  \propto e^{ - \beta H}
395   \label{introEquation:densityAndHamiltonian}
396 + \end{equation}
397 +
398 + \subsubsection{\label{introSection:phaseSpaceConservation}Conservation of Phase Space}
399 + Lets consider a region in the phase space,
400 + \begin{equation}
401 + \delta v = \int { \ldots \int {dq_1 } ...dq_f dp_1 } ..dp_f .
402 + \end{equation}
403 + If this region is small enough, the density $\rho$ can be regarded
404 + as uniform over the whole phase space. Thus, the number of phase
405 + points inside this region is given by,
406 + \begin{equation}
407 + \delta N = \rho \delta v = \rho \int { \ldots \int {dq_1 } ...dq_f
408 + dp_1 } ..dp_f.
409 + \end{equation}
410 +
411 + \begin{equation}
412 + \frac{{d(\delta N)}}{{dt}} = \frac{{d\rho }}{{dt}}\delta v + \rho
413 + \frac{d}{{dt}}(\delta v) = 0.
414 + \end{equation}
415 + With the help of stationary assumption
416 + (\ref{introEquation:stationary}), we obtain the principle of the
417 + \emph{conservation of extension in phase space},
418 + \begin{equation}
419 + \frac{d}{{dt}}(\delta v) = \frac{d}{{dt}}\int { \ldots \int {dq_1 }
420 + ...dq_f dp_1 } ..dp_f  = 0.
421 + \label{introEquation:volumePreserving}
422   \end{equation}
423  
424 + \subsubsection{\label{introSection:liouvilleInOtherForms}Liouville's Theorem in Other Forms}
425 +
426   Liouville's theorem can be expresses in a variety of different forms
427   which are convenient within different contexts. For any two function
428   $F$ and $G$ of the coordinates and momenta of a system, the Poisson
# Line 431 | Line 458 | expressed as
458   \label{introEquation:liouvilleTheoremInOperator}
459   \end{equation}
460  
434
461   \subsection{\label{introSection:ergodic}The Ergodic Hypothesis}
462  
463   Various thermodynamic properties can be calculated from Molecular
# Line 544 | Line 570 | The free rigid body is an example of Poisson system (a
570   \dot x = J(x)\nabla _x H \label{introEquation:poissonHamiltonian}
571   \end{equation}
572   The most obvious change being that matrix $J$ now depends on $x$.
547 The free rigid body is an example of Poisson system (actually a
548 Lie-Poisson system) with Hamiltonian function of angular kinetic
549 energy.
550 \begin{equation}
551 J(\pi ) = \left( {\begin{array}{*{20}c}
552   0 & {\pi _3 } & { - \pi _2 }  \\
553   { - \pi _3 } & 0 & {\pi _1 }  \\
554   {\pi _2 } & { - \pi _1 } & 0  \\
555 \end{array}} \right)
556 \end{equation}
573  
574 < \begin{equation}
559 < H = \frac{1}{2}\left( {\frac{{\pi _1^2 }}{{I_1 }} + \frac{{\pi _2^2
560 < }}{{I_2 }} + \frac{{\pi _3^2 }}{{I_3 }}} \right)
561 < \end{equation}
574 > \subsection{\label{introSection:exactFlow}Exact Flow}
575  
563 \subsection{\label{introSection:geometricProperties}Geometric Properties}
576   Let $x(t)$ be the exact solution of the ODE system,
577   \begin{equation}
578   \frac{{dx}}{{dt}} = f(x) \label{introEquation:ODE}
# Line 570 | Line 582 | space to itself. In most cases, it is not easy to find
582   x(t+\tau) =\varphi_\tau(x(t))
583   \]
584   where $\tau$ is a fixed time step and $\varphi$ is a map from phase
585 < space to itself. In most cases, it is not easy to find the exact
574 < flow $\varphi_\tau$. Instead, we use a approximate map, $\psi_\tau$,
575 < which is usually called integrator. The order of an integrator
576 < $\psi_\tau$ is $p$, if the Taylor series of $\psi_\tau$ agree to
577 < order $p$,
585 > space to itself. The flow has the continuous group property,
586   \begin{equation}
587 + \varphi _{\tau _1 }  \circ \varphi _{\tau _2 }  = \varphi _{\tau _1
588 + + \tau _2 } .
589 + \end{equation}
590 + In particular,
591 + \begin{equation}
592 + \varphi _\tau   \circ \varphi _{ - \tau }  = I
593 + \end{equation}
594 + Therefore, the exact flow is self-adjoint,
595 + \begin{equation}
596 + \varphi _\tau   = \varphi _{ - \tau }^{ - 1}.
597 + \end{equation}
598 + The exact flow can also be written in terms of the of an operator,
599 + \begin{equation}
600 + \varphi _\tau  (x) = e^{\tau \sum\limits_i {f_i (x)\frac{\partial
601 + }{{\partial x_i }}} } (x) \equiv \exp (\tau f)(x).
602 + \label{introEquation:exponentialOperator}
603 + \end{equation}
604 +
605 + In most cases, it is not easy to find the exact flow $\varphi_\tau$.
606 + Instead, we use a approximate map, $\psi_\tau$, which is usually
607 + called integrator. The order of an integrator $\psi_\tau$ is $p$, if
608 + the Taylor series of $\psi_\tau$ agree to order $p$,
609 + \begin{equation}
610   \psi_tau(x) = x + \tau f(x) + O(\tau^{p+1})
611   \end{equation}
612  
613 + \subsection{\label{introSection:geometricProperties}Geometric Properties}
614 +
615   The hidden geometric properties of ODE and its flow play important
616 < roles in numerical studies. Let $\varphi$ be the flow of Hamiltonian
617 < vector field, $\varphi$ is a \emph{symplectic} flow if it satisfies,
616 > roles in numerical studies. Many of them can be found in systems
617 > which occur naturally in applications.
618 >
619 > Let $\varphi$ be the flow of Hamiltonian vector field, $\varphi$ is
620 > a \emph{symplectic} flow if it satisfies,
621   \begin{equation}
622 < '\varphi^T J '\varphi = J.
622 > {\varphi '}^T J \varphi ' = J.
623   \end{equation}
624   According to Liouville's theorem, the symplectic volume is invariant
625   under a Hamiltonian flow, which is the basis for classical
# Line 591 | Line 627 | symplectomorphism. As to the Poisson system,
627   field on a symplectic manifold can be shown to be a
628   symplectomorphism. As to the Poisson system,
629   \begin{equation}
630 < '\varphi ^T J '\varphi  = J \circ \varphi
630 > {\varphi '}^T J \varphi ' = J \circ \varphi
631   \end{equation}
632 < is the property must be preserved by the integrator. It is possible
633 < to construct a \emph{volume-preserving} flow for a source free($
634 < \nabla \cdot f = 0 $) ODE, if the flow satisfies $ \det d\varphi  =
635 < 1$. Changing the variables $y = h(x)$ in a
636 < ODE\ref{introEquation:ODE} will result in a new system,
632 > is the property must be preserved by the integrator.
633 >
634 > It is possible to construct a \emph{volume-preserving} flow for a
635 > source free($ \nabla \cdot f = 0 $) ODE, if the flow satisfies $
636 > \det d\varphi  = 1$. One can show easily that a symplectic flow will
637 > be volume-preserving.
638 >
639 > Changing the variables $y = h(x)$ in a ODE\ref{introEquation:ODE}
640 > will result in a new system,
641   \[
642   \dot y = \tilde f(y) = ((dh \cdot f)h^{ - 1} )(y).
643   \]
644   The vector filed $f$ has reversing symmetry $h$ if $f = - \tilde f$.
645   In other words, the flow of this vector field is reversible if and
646 < only if $ h \circ \varphi ^{ - 1}  = \varphi  \circ h $. When
607 < designing any numerical methods, one should always try to preserve
608 < the structural properties of the original ODE and its flow.
646 > only if $ h \circ \varphi ^{ - 1}  = \varphi  \circ h $.
647  
648 + A \emph{first integral}, or conserved quantity of a general
649 + differential function is a function $ G:R^{2d}  \to R^d $ which is
650 + constant for all solutions of the ODE $\frac{{dx}}{{dt}} = f(x)$ ,
651 + \[
652 + \frac{{dG(x(t))}}{{dt}} = 0.
653 + \]
654 + Using chain rule, one may obtain,
655 + \[
656 + \sum\limits_i {\frac{{dG}}{{dx_i }}} f_i (x) = f \bullet \nabla G,
657 + \]
658 + which is the condition for conserving \emph{first integral}. For a
659 + canonical Hamiltonian system, the time evolution of an arbitrary
660 + smooth function $G$ is given by,
661 + \begin{equation}
662 + \begin{array}{c}
663 + \frac{{dG(x(t))}}{{dt}} = [\nabla _x G(x(t))]^T \dot x(t) \\
664 +  = [\nabla _x G(x(t))]^T J\nabla _x H(x(t)). \\
665 + \end{array}
666 + \label{introEquation:firstIntegral1}
667 + \end{equation}
668 + Using poisson bracket notion, Equation
669 + \ref{introEquation:firstIntegral1} can be rewritten as
670 + \[
671 + \frac{d}{{dt}}G(x(t)) = \left\{ {G,H} \right\}(x(t)).
672 + \]
673 + Therefore, the sufficient condition for $G$ to be the \emph{first
674 + integral} of a Hamiltonian system is
675 + \[
676 + \left\{ {G,H} \right\} = 0.
677 + \]
678 + As well known, the Hamiltonian (or energy) H of a Hamiltonian system
679 + is a \emph{first integral}, which is due to the fact $\{ H,H\}  =
680 + 0$.
681 +
682 +
683 + When designing any numerical methods, one should always try to
684 + preserve the structural properties of the original ODE and its flow.
685 +
686   \subsection{\label{introSection:constructionSymplectic}Construction of Symplectic Methods}
687   A lot of well established and very effective numerical methods have
688   been successful precisely because of their symplecticities even
# Line 622 | Line 698 | and difficult to use\cite{}. In dissipative systems, v
698   \end{enumerate}
699  
700   Generating function tends to lead to methods which are cumbersome
701 < and difficult to use\cite{}. In dissipative systems, variational
702 < methods can capture the decay of energy accurately\cite{}. Since
703 < their geometrically unstable nature against non-Hamiltonian
704 < perturbations, ordinary implicit Runge-Kutta methods are not
705 < suitable for Hamiltonian system. Recently, various high-order
706 < explicit Runge--Kutta methods have been developed to overcome this
707 < instability \cite{}. However, due to computational penalty involved
708 < in implementing the Runge-Kutta methods, they do not attract too
709 < much attention from Molecular Dynamics community. Instead, splitting
710 < have been widely accepted since they exploit natural decompositions
711 < of the system\cite{Tuckerman92}. The main idea behind splitting
712 < methods is to decompose the discrete $\varphi_h$ as a composition of
713 < simpler flows,
701 > and difficult to use. In dissipative systems, variational methods
702 > can capture the decay of energy accurately. Since their
703 > geometrically unstable nature against non-Hamiltonian perturbations,
704 > ordinary implicit Runge-Kutta methods are not suitable for
705 > Hamiltonian system. Recently, various high-order explicit
706 > Runge--Kutta methods have been developed to overcome this
707 > instability. However, due to computational penalty involved in
708 > implementing the Runge-Kutta methods, they do not attract too much
709 > attention from Molecular Dynamics community. Instead, splitting have
710 > been widely accepted since they exploit natural decompositions of
711 > the system\cite{Tuckerman92}.
712 >
713 > \subsubsection{\label{introSection:splittingMethod}Splitting Method}
714 >
715 > The main idea behind splitting methods is to decompose the discrete
716 > $\varphi_h$ as a composition of simpler flows,
717   \begin{equation}
718   \varphi _h  = \varphi _{h_1 }  \circ \varphi _{h_2 }  \ldots  \circ
719   \varphi _{h_n }
720   \label{introEquation:FlowDecomposition}
721   \end{equation}
722   where each of the sub-flow is chosen such that each represent a
723 < simpler integration of the system. Let $\phi$ and $\psi$ both be
724 < symplectic maps, it is easy to show that any composition of
725 < symplectic flows yields a symplectic map,
723 > simpler integration of the system.
724 >
725 > Suppose that a Hamiltonian system takes the form,
726 > \[
727 > H = H_1 + H_2.
728 > \]
729 > Here, $H_1$ and $H_2$ may represent different physical processes of
730 > the system. For instance, they may relate to kinetic and potential
731 > energy respectively, which is a natural decomposition of the
732 > problem. If $H_1$ and $H_2$ can be integrated using exact flows
733 > $\varphi_1(t)$ and $\varphi_2(t)$, respectively, a simple first
734 > order is then given by the Lie-Trotter formula
735   \begin{equation}
736 + \varphi _h  = \varphi _{1,h}  \circ \varphi _{2,h},
737 + \label{introEquation:firstOrderSplitting}
738 + \end{equation}
739 + where $\varphi _h$ is the result of applying the corresponding
740 + continuous $\varphi _i$ over a time $h$. By definition, as
741 + $\varphi_i(t)$ is the exact solution of a Hamiltonian system, it
742 + must follow that each operator $\varphi_i(t)$ is a symplectic map.
743 + It is easy to show that any composition of symplectic flows yields a
744 + symplectic map,
745 + \begin{equation}
746   (\varphi '\phi ')^T J\varphi '\phi ' = \phi '^T \varphi '^T J\varphi
747 < '\phi ' = \phi '^T J\phi ' = J.
747 > '\phi ' = \phi '^T J\phi ' = J,
748   \label{introEquation:SymplecticFlowComposition}
749   \end{equation}
750 < Suppose that a Hamiltonian system has a form with $H = T + V$
750 > where $\phi$ and $\psi$ both are symplectic maps. Thus operator
751 > splitting in this context automatically generates a symplectic map.
752 >
753 > The Lie-Trotter splitting(\ref{introEquation:firstOrderSplitting})
754 > introduces local errors proportional to $h^2$, while Strang
755 > splitting gives a second-order decomposition,
756 > \begin{equation}
757 > \varphi _h  = \varphi _{1,h/2}  \circ \varphi _{2,h}  \circ \varphi
758 > _{1,h/2} , \label{introEquation:secondOrderSplitting}
759 > \end{equation}
760 > which has a local error proportional to $h^3$. Sprang splitting's
761 > popularity in molecular simulation community attribute to its
762 > symmetric property,
763 > \begin{equation}
764 > \varphi _h^{ - 1} = \varphi _{ - h}.
765 > \label{introEquation:timeReversible}
766 > \end{equation}
767  
768 < \section{\label{introSection:molecularDynamics}Molecular Dynamics}
768 > \subsubsection{\label{introSection:exampleSplittingMethod}Example of Splitting Method}
769 > The classical equation for a system consisting of interacting
770 > particles can be written in Hamiltonian form,
771 > \[
772 > H = T + V
773 > \]
774 > where $T$ is the kinetic energy and $V$ is the potential energy.
775 > Setting $H_1 = T, H_2 = V$ and applying Strang splitting, one
776 > obtains the following:
777 > \begin{align}
778 > q(\Delta t) &= q(0) + \dot{q}(0)\Delta t +
779 >    \frac{F[q(0)]}{m}\frac{\Delta t^2}{2}, %
780 > \label{introEquation:Lp10a} \\%
781 > %
782 > \dot{q}(\Delta t) &= \dot{q}(0) + \frac{\Delta t}{2m}
783 >    \biggl [F[q(0)] + F[q(\Delta t)] \biggr]. %
784 > \label{introEquation:Lp10b}
785 > \end{align}
786 > where $F(t)$ is the force at time $t$. This integration scheme is
787 > known as \emph{velocity verlet} which is
788 > symplectic(\ref{introEquation:SymplecticFlowComposition}),
789 > time-reversible(\ref{introEquation:timeReversible}) and
790 > volume-preserving (\ref{introEquation:volumePreserving}). These
791 > geometric properties attribute to its long-time stability and its
792 > popularity in the community. However, the most commonly used
793 > velocity verlet integration scheme is written as below,
794 > \begin{align}
795 > \dot{q}\biggl (\frac{\Delta t}{2}\biggr ) &=
796 >    \dot{q}(0) + \frac{\Delta t}{2m}\, F[q(0)], \label{introEquation:Lp9a}\\%
797 > %
798 > q(\Delta t) &= q(0) + \Delta t\, \dot{q}\biggl (\frac{\Delta t}{2}\biggr ),%
799 >    \label{introEquation:Lp9b}\\%
800 > %
801 > \dot{q}(\Delta t) &= \dot{q}\biggl (\frac{\Delta t}{2}\biggr ) +
802 >    \frac{\Delta t}{2m}\, F[q(0)]. \label{introEquation:Lp9c}
803 > \end{align}
804 > From the preceding splitting, one can see that the integration of
805 > the equations of motion would follow:
806 > \begin{enumerate}
807 > \item calculate the velocities at the half step, $\frac{\Delta t}{2}$, from the forces calculated at the initial position.
808  
809 < As a special discipline of molecular modeling, Molecular dynamics
657 < has proven to be a powerful tool for studying the functions of
658 < biological systems, providing structural, thermodynamic and
659 < dynamical information.
809 > \item Use the half step velocities to move positions one whole step, $\Delta t$.
810  
811 < \subsection{\label{introSec:mdInit}Initialization}
811 > \item Evaluate the forces at the new positions, $\mathbf{r}(\Delta t)$, and use the new forces to complete the velocity move.
812  
813 < \subsection{\label{introSection:mdIntegration} Integration of the Equations of Motion}
813 > \item Repeat from step 1 with the new position, velocities, and forces assuming the roles of the initial values.
814 > \end{enumerate}
815  
816 < \section{\label{introSection:rigidBody}Dynamics of Rigid Bodies}
816 > Simply switching the order of splitting and composing, a new
817 > integrator, the \emph{position verlet} integrator, can be generated,
818 > \begin{align}
819 > \dot q(\Delta t) &= \dot q(0) + \Delta tF(q(0))\left[ {q(0) +
820 > \frac{{\Delta t}}{{2m}}\dot q(0)} \right], %
821 > \label{introEquation:positionVerlet1} \\%
822 > %
823 > q(\Delta t) &= q(0) + \frac{{\Delta t}}{2}\left[ {\dot q(0) + \dot
824 > q(\Delta t)} \right]. %
825 > \label{introEquation:positionVerlet2}
826 > \end{align}
827  
828 < A rigid body is a body in which the distance between any two given
668 < points of a rigid body remains constant regardless of external
669 < forces exerted on it. A rigid body therefore conserves its shape
670 < during its motion.
828 > \subsubsection{\label{introSection:errorAnalysis}Error Analysis and Higher Order Methods}
829  
830 < Applications of dynamics of rigid bodies.
831 <
832 < \subsection{\label{introSection:lieAlgebra}Lie Algebra}
833 <
834 < \subsection{\label{introSection:DLMMotionEquation}The Euler Equations of Rigid Body Motion}
677 <
678 < \subsection{\label{introSection:otherRBMotionEquation}Other Formulations for Rigid Body Motion}
679 <
680 < %\subsection{\label{introSection:poissonBrackets}Poisson Brackets}
681 <
682 < \section{\label{introSection:correlationFunctions}Correlation Functions}
683 <
684 < \section{\label{introSection:langevinDynamics}Langevin Dynamics}
685 <
686 < \subsection{\label{introSection:LDIntroduction}Introduction and application of Langevin Dynamics}
687 <
688 < \subsection{\label{introSection:generalizedLangevinDynamics}Generalized Langevin Dynamics}
689 <
830 > Baker-Campbell-Hausdorff formula can be used to determine the local
831 > error of splitting method in terms of commutator of the
832 > operators(\ref{introEquation:exponentialOperator}) associated with
833 > the sub-flow. For operators $hX$ and $hY$ which are associate to
834 > $\varphi_1(t)$ and $\varphi_2(t)$ respectively , we have
835   \begin{equation}
836 < H = \frac{{p^2 }}{{2m}} + U(x) + H_B  + \Delta U(x,x_1 , \ldots x_N)
692 < \label{introEquation:bathGLE}
836 > \exp (hX + hY) = \exp (hZ)
837   \end{equation}
838 < where $H_B$ is harmonic bath Hamiltonian,
838 > where
839 > \begin{equation}
840 > hZ = hX + hY + \frac{{h^2 }}{2}[X,Y] + \frac{{h^3 }}{2}\left(
841 > {[X,[X,Y]] + [Y,[Y,X]]} \right) +  \ldots .
842 > \end{equation}
843 > Here, $[X,Y]$ is the commutators of operator $X$ and $Y$ given by
844   \[
845 < H_B  =\sum\limits_{\alpha  = 1}^N {\left\{ {\frac{{p_\alpha ^2
697 < }}{{2m_\alpha  }} + \frac{1}{2}m_\alpha  w_\alpha ^2 } \right\}}
845 > [X,Y] = XY - YX .
846   \]
847 < and $\Delta U$ is bilinear system-bath coupling,
847 > Applying Baker-Campbell-Hausdorff formula to Sprang splitting, we
848 > can obtain
849 > \begin{equation}
850 > \exp (h X/2)\exp (h Y)\exp (h X/2) & = & \exp (h X + h Y + h^2
851 > [X,Y]/4 + h^2 [Y,X]/4 \\ & & \mbox{} + h^2 [X,X]/8 + h^2 [Y,Y]/8 \\
852 > & & \mbox{} + h^3 [Y,[Y,X]]/12 - h^3 [X,[X,Y]]/24 & & \mbox{} +
853 > \ldots )
854 > \end{equation}
855 > Since \[ [X,Y] + [Y,X] = 0\] and \[ [X,X] = 0\], the dominant local
856 > error of Spring splitting is proportional to $h^3$. The same
857 > procedure can be applied to general splitting,  of the form
858 > \begin{equation}
859 > \varphi _{b_m h}^2  \circ \varphi _{a_m h}^1  \circ \varphi _{b_{m -
860 > 1} h}^2  \circ  \ldots  \circ \varphi _{a_1 h}^1 .
861 > \end{equation}
862 > Careful choice of coefficient $a_1 ,\ldot , b_m$ will lead to higher
863 > order method. Yoshida proposed an elegant way to compose higher
864 > order methods based on symmetric splitting. Given a symmetric second
865 > order base method $ \varphi _h^{(2)} $, a fourth-order symmetric
866 > method can be constructed by composing,
867   \[
868 < \Delta U =  - \sum\limits_{\alpha  = 1}^N {g_\alpha  x_\alpha  x}
868 > \varphi _h^{(4)}  = \varphi _{\alpha h}^{(2)}  \circ \varphi _{\beta
869 > h}^{(2)}  \circ \varphi _{\alpha h}^{(2)}
870   \]
871 < Completing the square,
871 > where $ \alpha  =  - \frac{{2^{1/3} }}{{2 - 2^{1/3} }}$ and $ \beta
872 > = \frac{{2^{1/3} }}{{2 - 2^{1/3} }}$. Moreover, a symmetric
873 > integrator $ \varphi _h^{(2n + 2)}$ can be composed by
874 > \begin{equation}
875 > \varphi _h^{(2n + 2)}  = \varphi _{\alpha h}^{(2n)}  \circ \varphi
876 > _{\beta h}^{(2n)}  \circ \varphi _{\alpha h}^{(2n)}
877 > \end{equation}
878 > , if the weights are chosen as
879   \[
880 < H_B  + \Delta U = \sum\limits_{\alpha  = 1}^N {\left\{
881 < {\frac{{p_\alpha ^2 }}{{2m_\alpha  }} + \frac{1}{2}m_\alpha
707 < w_\alpha ^2 \left( {x_\alpha   - \frac{{g_\alpha  }}{{m_\alpha
708 < w_\alpha ^2 }}x} \right)^2 } \right\}}  - \sum\limits_{\alpha  =
709 < 1}^N {\frac{{g_\alpha ^2 }}{{2m_\alpha  w_\alpha ^2 }}} x^2
880 > \alpha  =  - \frac{{2^{1/(2n + 1)} }}{{2 - 2^{1/(2n + 1)} }},\beta =
881 > \frac{{2^{1/(2n + 1)} }}{{2 - 2^{1/(2n + 1)} }} .
882   \]
711 and putting it back into Eq.~\ref{introEquation:bathGLE},
712 \[
713 H = \frac{{p^2 }}{{2m}} + W(x) + \sum\limits_{\alpha  = 1}^N
714 {\left\{ {\frac{{p_\alpha ^2 }}{{2m_\alpha  }} + \frac{1}{2}m_\alpha
715 w_\alpha ^2 \left( {x_\alpha   - \frac{{g_\alpha  }}{{m_\alpha
716 w_\alpha ^2 }}x} \right)^2 } \right\}}
717 \]
718 where
719 \[
720 W(x) = U(x) - \sum\limits_{\alpha  = 1}^N {\frac{{g_\alpha ^2
721 }}{{2m_\alpha  w_\alpha ^2 }}} x^2
722 \]
723 Since the first two terms of the new Hamiltonian depend only on the
724 system coordinates, we can get the equations of motion for
725 Generalized Langevin Dynamics by Hamilton's equations
726 \ref{introEquation:motionHamiltonianCoordinate,
727 introEquation:motionHamiltonianMomentum},
728 \begin{align}
729 \dot p &=  - \frac{{\partial H}}{{\partial x}}
730       &= m\ddot x
731       &= - \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)}
732 \label{introEq:Lp5}
733 \end{align}
734 , and
735 \begin{align}
736 \dot p_\alpha   &=  - \frac{{\partial H}}{{\partial x_\alpha  }}
737                &= m\ddot x_\alpha
738                &= \- m_\alpha  w_\alpha ^2 \left( {x_\alpha   - \frac{{g_\alpha}}{{m_\alpha  w_\alpha ^2 }}x} \right)
739 \end{align}
883  
884 < \subsection{\label{introSection:laplaceTransform}The Laplace Transform}
884 > \section{\label{introSection:molecularDynamics}Molecular Dynamics}
885  
886 < \[
887 < L(x) = \int_0^\infty  {x(t)e^{ - pt} dt}
886 > As one of the principal tools of molecular modeling, Molecular
887 > dynamics has proven to be a powerful tool for studying the functions
888 > of biological systems, providing structural, thermodynamic and
889 > dynamical information. The basic idea of molecular dynamics is that
890 > macroscopic properties are related to microscopic behavior and
891 > microscopic behavior can be calculated from the trajectories in
892 > simulations. For instance, instantaneous temperature of an
893 > Hamiltonian system of $N$ particle can be measured by
894 > \[
895 > T = \sum\limits_{i = 1}^N {\frac{{m_i v_i^2 }}{{fk_B }}}
896   \]
897 + where $m_i$ and $v_i$ are the mass and velocity of $i$th particle
898 + respectively, $f$ is the number of degrees of freedom, and $k_B$ is
899 + the boltzman constant.
900  
901 + A typical molecular dynamics run consists of three essential steps:
902 + \begin{enumerate}
903 +  \item Initialization
904 +    \begin{enumerate}
905 +    \item Preliminary preparation
906 +    \item Minimization
907 +    \item Heating
908 +    \item Equilibration
909 +    \end{enumerate}
910 +  \item Production
911 +  \item Analysis
912 + \end{enumerate}
913 + These three individual steps will be covered in the following
914 + sections. Sec.~\ref{introSec:initialSystemSettings} deals with the
915 + initialization of a simulation. Sec.~\ref{introSec:production} will
916 + discusses issues in production run. Sec.~\ref{introSection:Analysis}
917 + provides the theoretical tools for trajectory analysis.
918 +
919 + \subsection{\label{introSec:initialSystemSettings}Initialization}
920 +
921 + \subsubsection{Preliminary preparation}
922 +
923 + When selecting the starting structure of a molecule for molecular
924 + simulation, one may retrieve its Cartesian coordinates from public
925 + databases, such as RCSB Protein Data Bank \textit{etc}. Although
926 + thousands of crystal structures of molecules are discovered every
927 + year, many more remain unknown due to the difficulties of
928 + purification and crystallization. Even for the molecule with known
929 + structure, some important information is missing. For example, the
930 + missing hydrogen atom which acts as donor in hydrogen bonding must
931 + be added. Moreover, in order to include electrostatic interaction,
932 + one may need to specify the partial charges for individual atoms.
933 + Under some circumstances, we may even need to prepare the system in
934 + a special setup. For instance, when studying transport phenomenon in
935 + membrane system, we may prepare the lipids in bilayer structure
936 + instead of placing lipids randomly in solvent, since we are not
937 + interested in self-aggregation and it takes a long time to happen.
938 +
939 + \subsubsection{Minimization}
940 +
941 + It is quite possible that some of molecules in the system from
942 + preliminary preparation may be overlapped with each other. This
943 + close proximity leads to high potential energy which consequently
944 + jeopardizes any molecular dynamics simulations. To remove these
945 + steric overlaps, one typically performs energy minimization to find
946 + a more reasonable conformation. Several energy minimization methods
947 + have been developed to exploit the energy surface and to locate the
948 + local minimum. While converging slowly near the minimum, steepest
949 + descent method is extremely robust when systems are far from
950 + harmonic. Thus, it is often used to refine structure from
951 + crystallographic data. Relied on the gradient or hessian, advanced
952 + methods like conjugate gradient and Newton-Raphson converge rapidly
953 + to a local minimum, while become unstable if the energy surface is
954 + far from quadratic. Another factor must be taken into account, when
955 + choosing energy minimization method, is the size of the system.
956 + Steepest descent and conjugate gradient can deal with models of any
957 + size. Because of the limit of computation power to calculate hessian
958 + matrix and insufficient storage capacity to store them, most
959 + Newton-Raphson methods can not be used with very large models.
960 +
961 + \subsubsection{Heating}
962 +
963 + Typically, Heating is performed by assigning random velocities
964 + according to a Gaussian distribution for a temperature. Beginning at
965 + a lower temperature and gradually increasing the temperature by
966 + assigning greater random velocities, we end up with setting the
967 + temperature of the system to a final temperature at which the
968 + simulation will be conducted. In heating phase, we should also keep
969 + the system from drifting or rotating as a whole. Equivalently, the
970 + net linear momentum and angular momentum of the system should be
971 + shifted to zero.
972 +
973 + \subsubsection{Equilibration}
974 +
975 + The purpose of equilibration is to allow the system to evolve
976 + spontaneously for a period of time and reach equilibrium. The
977 + procedure is continued until various statistical properties, such as
978 + temperature, pressure, energy, volume and other structural
979 + properties \textit{etc}, become independent of time. Strictly
980 + speaking, minimization and heating are not necessary, provided the
981 + equilibration process is long enough. However, these steps can serve
982 + as a means to arrive at an equilibrated structure in an effective
983 + way.
984 +
985 + \subsection{\label{introSection:production}Production}
986 +
987 + Production run is the most important steps of the simulation, in
988 + which the equilibrated structure is used as a starting point and the
989 + motions of the molecules are collected for later analysis. In order
990 + to capture the macroscopic properties of the system, the molecular
991 + dynamics simulation must be performed in correct and efficient way.
992 +
993 + The most expensive part of a molecular dynamics simulation is the
994 + calculation of non-bonded forces, such as van der Waals force and
995 + Coulombic forces \textit{etc}. For a system of $N$ particles, the
996 + complexity of the algorithm for pair-wise interactions is $O(N^2 )$,
997 + which making large simulations prohibitive in the absence of any
998 + computation saving techniques.
999 +
1000 + A natural approach to avoid system size issue is to represent the
1001 + bulk behavior by a finite number of the particles. However, this
1002 + approach will suffer from the surface effect. To offset this,
1003 + \textit{Periodic boundary condition} is developed to simulate bulk
1004 + properties with a relatively small number of particles. In this
1005 + method, the simulation box is replicated throughout space to form an
1006 + infinite lattice. During the simulation, when a particle moves in
1007 + the primary cell, its image in other cells move in exactly the same
1008 + direction with exactly the same orientation. Thus, as a particle
1009 + leaves the primary cell, one of its images will enter through the
1010 + opposite face.
1011 + %\begin{figure}
1012 + %\centering
1013 + %\includegraphics[width=\linewidth]{pbcFig.eps}
1014 + %\caption[An illustration of periodic boundary conditions]{A 2-D
1015 + %illustration of periodic boundary conditions. As one particle leaves
1016 + %the right of the simulation box, an image of it enters the left.}
1017 + %\label{introFig:pbc}
1018 + %\end{figure}
1019 +
1020 + %cutoff and minimum image convention
1021 + Another important technique to improve the efficiency of force
1022 + evaluation is to apply cutoff where particles farther than a
1023 + predetermined distance, are not included in the calculation
1024 + \cite{Frenkel1996}. The use of a cutoff radius will cause a
1025 + discontinuity in the potential energy curve
1026 + (Fig.~\ref{introFig:shiftPot}). Fortunately, one can shift the
1027 + potential to ensure the potential curve go smoothly to zero at the
1028 + cutoff radius. Cutoff strategy works pretty well for Lennard-Jones
1029 + interaction because of its short range nature. However, simply
1030 + truncating the electrostatic interaction with the use of cutoff has
1031 + been shown to lead to severe artifacts in simulations. Ewald
1032 + summation, in which the slowly conditionally convergent Coulomb
1033 + potential is transformed into direct and reciprocal sums with rapid
1034 + and absolute convergence, has proved to minimize the periodicity
1035 + artifacts in liquid simulations. Taking the advantages of the fast
1036 + Fourier transform (FFT) for calculating discrete Fourier transforms,
1037 + the particle mesh-based methods are accelerated from $O(N^{3/2})$ to
1038 + $O(N logN)$. An alternative approach is \emph{fast multipole
1039 + method}, which treats Coulombic interaction exactly at short range,
1040 + and approximate the potential at long range through multipolar
1041 + expansion. In spite of their wide acceptances at the molecular
1042 + simulation community, these two methods are hard to be implemented
1043 + correctly and efficiently. Instead, we use a damped and
1044 + charge-neutralized Coulomb potential method developed by Wolf and
1045 + his coworkers. The shifted Coulomb potential for particle $i$ and
1046 + particle $j$ at distance $r_{rj}$ is given by:
1047 + \begin{equation}
1048 + V(r_{ij})= \frac{q_i q_j \textrm{erfc}(\alpha
1049 + r_{ij})}{r_{ij}}-\lim_{r_{ij}\rightarrow
1050 + R_\textrm{c}}\left\{\frac{q_iq_j \textrm{erfc}(\alpha
1051 + r_{ij})}{r_{ij}}\right\}. \label{introEquation:shiftedCoulomb}
1052 + \end{equation}
1053 + where $\alpha$ is the convergence parameter. Due to the lack of
1054 + inherent periodicity and rapid convergence,this method is extremely
1055 + efficient and easy to implement.
1056 + %\begin{figure}
1057 + %\centering
1058 + %\includegraphics[width=\linewidth]{pbcFig.eps}
1059 + %\caption[An illustration of shifted Coulomb potential]{An illustration of shifted Coulomb potential.}
1060 + %\label{introFigure:shiftedCoulomb}
1061 + %\end{figure}
1062 +
1063 + %multiple time step
1064 +
1065 + \subsection{\label{introSection:Analysis} Analysis}
1066 +
1067 + Recently, advanced visualization technique are widely applied to
1068 + monitor the motions of molecules. Although the dynamics of the
1069 + system can be described qualitatively from animation, quantitative
1070 + trajectory analysis are more appreciable. According to the
1071 + principles of Statistical Mechanics,
1072 + Sec.~\ref{introSection:statisticalMechanics}, one can compute
1073 + thermodynamics properties, analyze fluctuations of structural
1074 + parameters, and investigate time-dependent processes of the molecule
1075 + from the trajectories.
1076 +
1077 + \subsubsection{\label{introSection:thermodynamicsProperties}Thermodynamics Properties}
1078 +
1079 + Thermodynamics properties, which can be expressed in terms of some
1080 + function of the coordinates and momenta of all particles in the
1081 + system, can be directly computed from molecular dynamics. The usual
1082 + way to measure the pressure is based on virial theorem of Clausius
1083 + which states that the virial is equal to $-3Nk_BT$. For a system
1084 + with forces between particles, the total virial, $W$, contains the
1085 + contribution from external pressure and interaction between the
1086 + particles:
1087   \[
1088 < L(x + y) = L(x) + L(y)
1088 > W =  - 3PV + \left\langle {\sum\limits_{i < j} {r{}_{ij} \cdot
1089 > f_{ij} } } \right\rangle
1090   \]
1091 + where $f_{ij}$ is the force between particle $i$ and $j$ at a
1092 + distance $r_{ij}$. Thus, the expression for the pressure is given
1093 + by:
1094 + \begin{equation}
1095 + P = \frac{{Nk_B T}}{V} - \frac{1}{{3V}}\left\langle {\sum\limits_{i
1096 + < j} {r{}_{ij} \cdot f_{ij} } } \right\rangle
1097 + \end{equation}
1098  
1099 + \subsubsection{\label{introSection:structuralProperties}Structural Properties}
1100 +
1101 + Structural Properties of a simple fluid can be described by a set of
1102 + distribution functions. Among these functions,\emph{pair
1103 + distribution function}, also known as \emph{radial distribution
1104 + function}, is of most fundamental importance to liquid-state theory.
1105 + Pair distribution function can be gathered by Fourier transforming
1106 + raw data from a series of neutron diffraction experiments and
1107 + integrating over the surface factor \cite{Powles73}. The experiment
1108 + result can serve as a criterion to justify the correctness of the
1109 + theory. Moreover, various equilibrium thermodynamic and structural
1110 + properties can also be expressed in terms of radial distribution
1111 + function \cite{allen87:csl}.
1112 +
1113 + A pair distribution functions $g(r)$ gives the probability that a
1114 + particle $i$ will be located at a distance $r$ from a another
1115 + particle $j$ in the system
1116   \[
1117 < L(ax) = aL(x)
1117 > g(r) = \frac{V}{{N^2 }}\left\langle {\sum\limits_i {\sum\limits_{j
1118 > \ne i} {\delta (r - r_{ij} )} } } \right\rangle.
1119   \]
1120 + Note that the delta function can be replaced by a histogram in
1121 + computer simulation. Figure
1122 + \ref{introFigure:pairDistributionFunction} shows a typical pair
1123 + distribution function for the liquid argon system. The occurrence of
1124 + several peaks in the plot of $g(r)$ suggests that it is more likely
1125 + to find particles at certain radial values than at others. This is a
1126 + result of the attractive interaction at such distances. Because of
1127 + the strong repulsive forces at short distance, the probability of
1128 + locating particles at distances less than about 2.5{\AA} from each
1129 + other is essentially zero.
1130  
1131 + %\begin{figure}
1132 + %\centering
1133 + %\includegraphics[width=\linewidth]{pdf.eps}
1134 + %\caption[Pair distribution function for the liquid argon
1135 + %]{Pair distribution function for the liquid argon}
1136 + %\label{introFigure:pairDistributionFunction}
1137 + %\end{figure}
1138 +
1139 + \subsubsection{\label{introSection:timeDependentProperties}Time-dependent
1140 + Properties}
1141 +
1142 + Time-dependent properties are usually calculated using \emph{time
1143 + correlation function}, which correlates random variables $A$ and $B$
1144 + at two different time
1145 + \begin{equation}
1146 + C_{AB} (t) = \left\langle {A(t)B(0)} \right\rangle.
1147 + \label{introEquation:timeCorrelationFunction}
1148 + \end{equation}
1149 + If $A$ and $B$ refer to same variable, this kind of correlation
1150 + function is called \emph{auto correlation function}. One example of
1151 + auto correlation function is velocity auto-correlation function
1152 + which is directly related to transport properties of molecular
1153 + liquids:
1154   \[
1155 < L(\dot x) = pL(x) - px(0)
1155 > D = \frac{1}{3}\int\limits_0^\infty  {\left\langle {v(t) \cdot v(0)}
1156 > \right\rangle } dt
1157   \]
1158 + where $D$ is diffusion constant. Unlike velocity autocorrelation
1159 + function which is averaging over time origins and over all the
1160 + atoms, dipole autocorrelation are calculated for the entire system.
1161 + The dipole autocorrelation function is given by:
1162 + \[
1163 + c_{dipole}  = \left\langle {u_{tot} (t) \cdot u_{tot} (t)}
1164 + \right\rangle
1165 + \]
1166 + Here $u_{tot}$ is the net dipole of the entire system and is given
1167 + by
1168 + \[
1169 + u_{tot} (t) = \sum\limits_i {u_i (t)}
1170 + \]
1171 + In principle, many time correlation functions can be related with
1172 + Fourier transforms of the infrared, Raman, and inelastic neutron
1173 + scattering spectra of molecular liquids. In practice, one can
1174 + extract the IR spectrum from the intensity of dipole fluctuation at
1175 + each frequency using the following relationship:
1176 + \[
1177 + \hat c_{dipole} (v) = \int_{ - \infty }^\infty  {c_{dipole} (t)e^{ -
1178 + i2\pi vt} dt}
1179 + \]
1180  
1181 + \section{\label{introSection:rigidBody}Dynamics of Rigid Bodies}
1182 +
1183 + Rigid bodies are frequently involved in the modeling of different
1184 + areas, from engineering, physics, to chemistry. For example,
1185 + missiles and vehicle are usually modeled by rigid bodies.  The
1186 + movement of the objects in 3D gaming engine or other physics
1187 + simulator is governed by the rigid body dynamics. In molecular
1188 + simulation, rigid body is used to simplify the model in
1189 + protein-protein docking study{\cite{Gray03}}.
1190 +
1191 + It is very important to develop stable and efficient methods to
1192 + integrate the equations of motion of orientational degrees of
1193 + freedom. Euler angles are the nature choice to describe the
1194 + rotational degrees of freedom. However, due to its singularity, the
1195 + numerical integration of corresponding equations of motion is very
1196 + inefficient and inaccurate. Although an alternative integrator using
1197 + different sets of Euler angles can overcome this difficulty\cite{},
1198 + the computational penalty and the lost of angular momentum
1199 + conservation still remain. A singularity free representation
1200 + utilizing quaternions was developed by Evans in 1977. Unfortunately,
1201 + this approach suffer from the nonseparable Hamiltonian resulted from
1202 + quaternion representation, which prevents the symplectic algorithm
1203 + to be utilized. Another different approach is to apply holonomic
1204 + constraints to the atoms belonging to the rigid body. Each atom
1205 + moves independently under the normal forces deriving from potential
1206 + energy and constraint forces which are used to guarantee the
1207 + rigidness. However, due to their iterative nature, SHAKE and Rattle
1208 + algorithm converge very slowly when the number of constraint
1209 + increases.
1210 +
1211 + The break through in geometric literature suggests that, in order to
1212 + develop a long-term integration scheme, one should preserve the
1213 + symplectic structure of the flow. Introducing conjugate momentum to
1214 + rotation matrix $Q$ and re-formulating Hamiltonian's equation, a
1215 + symplectic integrator, RSHAKE, was proposed to evolve the
1216 + Hamiltonian system in a constraint manifold by iteratively
1217 + satisfying the orthogonality constraint $Q_T Q = 1$. An alternative
1218 + method using quaternion representation was developed by Omelyan.
1219 + However, both of these methods are iterative and inefficient. In
1220 + this section, we will present a symplectic Lie-Poisson integrator
1221 + for rigid body developed by Dullweber and his
1222 + coworkers\cite{Dullweber1997} in depth.
1223 +
1224 + \subsection{\label{introSection:constrainedHamiltonianRB}Constrained Hamiltonian for Rigid Body}
1225 + The motion of the rigid body is Hamiltonian with the Hamiltonian
1226 + function
1227 + \begin{equation}
1228 + H = \frac{1}{2}(p^T m^{ - 1} p) + \frac{1}{2}tr(PJ^{ - 1} P) +
1229 + V(q,Q) + \frac{1}{2}tr[(QQ^T  - 1)\Lambda ].
1230 + \label{introEquation:RBHamiltonian}
1231 + \end{equation}
1232 + Here, $q$ and $Q$  are the position and rotation matrix for the
1233 + rigid-body, $p$ and $P$  are conjugate momenta to $q$  and $Q$ , and
1234 + $J$, a diagonal matrix, is defined by
1235   \[
1236 < L(\ddot x) = p^2 L(x) - px(0) - \dot x(0)
1236 > I_{ii}^{ - 1}  = \frac{1}{2}\sum\limits_{i \ne j} {J_{jj}^{ - 1} }
1237   \]
1238 + where $I_{ii}$ is the diagonal element of the inertia tensor. This
1239 + constrained Hamiltonian equation subjects to a holonomic constraint,
1240 + \begin{equation}
1241 + Q^T Q = 1, \label{introEquation:orthogonalConstraint}
1242 + \end{equation}
1243 + which is used to ensure rotation matrix's orthogonality.
1244 + Differentiating \ref{introEquation:orthogonalConstraint} and using
1245 + Equation \ref{introEquation:RBMotionMomentum}, one may obtain,
1246 + \begin{equation}
1247 + Q^T PJ^{ - 1}  + J^{ - 1} P^T Q = 0 . \\
1248 + \label{introEquation:RBFirstOrderConstraint}
1249 + \end{equation}
1250  
1251 + Using Equation (\ref{introEquation:motionHamiltonianCoordinate},
1252 + \ref{introEquation:motionHamiltonianMomentum}), one can write down
1253 + the equations of motion,
1254   \[
1255 < L\left( {\int_0^t {g(t - \tau )h(\tau )d\tau } } \right) = G(p)H(p)
1255 > \begin{array}{c}
1256 > \frac{{dq}}{{dt}} = \frac{p}{m} \label{introEquation:RBMotionPosition}\\
1257 > \frac{{dp}}{{dt}} =  - \nabla _q V(q,Q) \label{introEquation:RBMotionMomentum}\\
1258 > \frac{{dQ}}{{dt}} = PJ^{ - 1}  \label{introEquation:RBMotionRotation}\\
1259 > \frac{{dP}}{{dt}} =  - \nabla _Q V(q,Q) - 2Q\Lambda . \label{introEquation:RBMotionP}\\
1260 > \end{array}
1261   \]
1262  
1263 < Some relatively important transformation,
1263 > In general, there are two ways to satisfy the holonomic constraints.
1264 > We can use constraint force provided by lagrange multiplier on the
1265 > normal manifold to keep the motion on constraint space. Or we can
1266 > simply evolve the system in constraint manifold. The two method are
1267 > proved to be equivalent. The holonomic constraint and equations of
1268 > motions define a constraint manifold for rigid body
1269   \[
1270 < L(\cos at) = \frac{p}{{p^2  + a^2 }}
1270 > M = \left\{ {(Q,P):Q^T Q = 1,Q^T PJ^{ - 1}  + J^{ - 1} P^T Q = 0}
1271 > \right\}.
1272   \]
1273 +
1274 + Unfortunately, this constraint manifold is not the cotangent bundle
1275 + $T_{\star}SO(3)$. However, it turns out that under symplectic
1276 + transformation, the cotangent space and the phase space are
1277 + diffeomorphic. Introducing
1278 + \[
1279 + \tilde Q = Q,\tilde P = \frac{1}{2}\left( {P - QP^T Q} \right),
1280 + \]
1281 + the mechanical system subject to a holonomic constraint manifold $M$
1282 + can be re-formulated as a Hamiltonian system on the cotangent space
1283 + \[
1284 + T^* SO(3) = \left\{ {(\tilde Q,\tilde P):\tilde Q^T \tilde Q =
1285 + 1,\tilde Q^T \tilde PJ^{ - 1}  + J^{ - 1} P^T \tilde Q = 0} \right\}
1286 + \]
1287 +
1288 + For a body fixed vector $X_i$ with respect to the center of mass of
1289 + the rigid body, its corresponding lab fixed vector $X_0^{lab}$  is
1290 + given as
1291 + \begin{equation}
1292 + X_i^{lab} = Q X_i + q.
1293 + \end{equation}
1294 + Therefore, potential energy $V(q,Q)$ is defined by
1295 + \[
1296 + V(q,Q) = V(Q X_0 + q).
1297 + \]
1298 + Hence, the force and torque are given by
1299 + \[
1300 + \nabla _q V(q,Q) = F(q,Q) = \sum\limits_i {F_i (q,Q)},
1301 + \]
1302 + and
1303 + \[
1304 + \nabla _Q V(q,Q) = F(q,Q)X_i^t
1305 + \]
1306 + respectively.
1307  
1308 + As a common choice to describe the rotation dynamics of the rigid
1309 + body, angular momentum on body frame $\Pi  = Q^t P$ is introduced to
1310 + rewrite the equations of motion,
1311 + \begin{equation}
1312 + \begin{array}{l}
1313 + \mathop \Pi \limits^ \bullet   = J^{ - 1} \Pi ^T \Pi  + Q^T \sum\limits_i {F_i (q,Q)X_i^T }  - \Lambda  \\
1314 + \mathop Q\limits^{{\rm{   }} \bullet }  = Q\Pi {\rm{ }}J^{ - 1}  \\
1315 + \end{array}
1316 + \label{introEqaution:RBMotionPI}
1317 + \end{equation}
1318 + , as well as holonomic constraints,
1319   \[
1320 < L(\sin at) = \frac{a}{{p^2  + a^2 }}
1320 > \begin{array}{l}
1321 > \Pi J^{ - 1}  + J^{ - 1} \Pi ^t  = 0 \\
1322 > Q^T Q = 1 \\
1323 > \end{array}
1324   \]
1325  
1326 + For a vector $v(v_1 ,v_2 ,v_3 ) \in R^3$ and a matrix $\hat v \in
1327 + so(3)^ \star$, the hat-map isomorphism,
1328 + \begin{equation}
1329 + v(v_1 ,v_2 ,v_3 ) \Leftrightarrow \hat v = \left(
1330 + {\begin{array}{*{20}c}
1331 +   0 & { - v_3 } & {v_2 }  \\
1332 +   {v_3 } & 0 & { - v_1 }  \\
1333 +   { - v_2 } & {v_1 } & 0  \\
1334 + \end{array}} \right),
1335 + \label{introEquation:hatmapIsomorphism}
1336 + \end{equation}
1337 + will let us associate the matrix products with traditional vector
1338 + operations
1339   \[
1340 < L(1) = \frac{1}{p}
1340 > \hat vu = v \times u
1341   \]
1342  
1343 < First, the bath coordinates,
1343 > Using \ref{introEqaution:RBMotionPI}, one can construct a skew
1344 > matrix,
1345 > \begin{equation}
1346 > (\mathop \Pi \limits^ \bullet   - \mathop \Pi \limits^ \bullet  ^T
1347 > ){\rm{ }} = {\rm{ }}(\Pi  - \Pi ^T ){\rm{ }}(J^{ - 1} \Pi  + \Pi J^{
1348 > - 1} ) + \sum\limits_i {[Q^T F_i (r,Q)X_i^T  - X_i F_i (r,Q)^T Q]} -
1349 > (\Lambda  - \Lambda ^T ) . \label{introEquation:skewMatrixPI}
1350 > \end{equation}
1351 > Since $\Lambda$ is symmetric, the last term of Equation
1352 > \ref{introEquation:skewMatrixPI} is zero, which implies the Lagrange
1353 > multiplier $\Lambda$ is absent from the equations of motion. This
1354 > unique property eliminate the requirement of iterations which can
1355 > not be avoided in other methods\cite{}.
1356 >
1357 > Applying hat-map isomorphism, we obtain the equation of motion for
1358 > angular momentum on body frame
1359 > \begin{equation}
1360 > \dot \pi  = \pi  \times I^{ - 1} \pi  + \sum\limits_i {\left( {Q^T
1361 > F_i (r,Q)} \right) \times X_i }.
1362 > \label{introEquation:bodyAngularMotion}
1363 > \end{equation}
1364 > In the same manner, the equation of motion for rotation matrix is
1365 > given by
1366   \[
1367 < p^2 L(x_\alpha  ) - px_\alpha  (0) - \dot x_\alpha  (0) =  - \omega
783 < _\alpha ^2 L(x_\alpha  ) + \frac{{g_\alpha  }}{{\omega _\alpha
784 < }}L(x)
1367 > \dot Q = Qskew(I^{ - 1} \pi )
1368   \]
1369 +
1370 + \subsection{\label{introSection:SymplecticFreeRB}Symplectic
1371 + Lie-Poisson Integrator for Free Rigid Body}
1372 +
1373 + If there is not external forces exerted on the rigid body, the only
1374 + contribution to the rotational is from the kinetic potential (the
1375 + first term of \ref{ introEquation:bodyAngularMotion}). The free
1376 + rigid body is an example of Lie-Poisson system with Hamiltonian
1377 + function
1378 + \begin{equation}
1379 + T^r (\pi ) = T_1 ^r (\pi _1 ) + T_2^r (\pi _2 ) + T_3^r (\pi _3 )
1380 + \label{introEquation:rotationalKineticRB}
1381 + \end{equation}
1382 + where $T_i^r (\pi _i ) = \frac{{\pi _i ^2 }}{{2I_i }}$ and
1383 + Lie-Poisson structure matrix,
1384 + \begin{equation}
1385 + J(\pi ) = \left( {\begin{array}{*{20}c}
1386 +   0 & {\pi _3 } & { - \pi _2 }  \\
1387 +   { - \pi _3 } & 0 & {\pi _1 }  \\
1388 +   {\pi _2 } & { - \pi _1 } & 0  \\
1389 + \end{array}} \right)
1390 + \end{equation}
1391 + Thus, the dynamics of free rigid body is governed by
1392 + \begin{equation}
1393 + \frac{d}{{dt}}\pi  = J(\pi )\nabla _\pi  T^r (\pi )
1394 + \end{equation}
1395 +
1396 + One may notice that each $T_i^r$ in Equation
1397 + \ref{introEquation:rotationalKineticRB} can be solved exactly. For
1398 + instance, the equations of motion due to $T_1^r$ are given by
1399 + \begin{equation}
1400 + \frac{d}{{dt}}\pi  = R_1 \pi ,\frac{d}{{dt}}Q = QR_1
1401 + \label{introEqaution:RBMotionSingleTerm}
1402 + \end{equation}
1403 + where
1404 + \[ R_1  = \left( {\begin{array}{*{20}c}
1405 +   0 & 0 & 0  \\
1406 +   0 & 0 & {\pi _1 }  \\
1407 +   0 & { - \pi _1 } & 0  \\
1408 + \end{array}} \right).
1409 + \]
1410 + The solutions of Equation \ref{introEqaution:RBMotionSingleTerm} is
1411   \[
1412 < L(x_\alpha  ) = \frac{{\frac{{g_\alpha  }}{{\omega _\alpha  }}L(x) +
1413 < px_\alpha  (0) + \dot x_\alpha  (0)}}{{p^2  + \omega _\alpha ^2 }}
1412 > \pi (\Delta t) = e^{\Delta tR_1 } \pi (0),Q(\Delta t) =
1413 > Q(0)e^{\Delta tR_1 }
1414   \]
1415 < Then, the system coordinates,
1416 < \begin{align}
1417 < mL(\ddot x) &=  - \frac{1}{p}\frac{{\partial W(x)}}{{\partial x}} -
1418 < \sum\limits_{\alpha  = 1}^N {\left\{ {\frac{{\frac{{g_\alpha
1419 < }}{{\omega _\alpha  }}L(x) + px_\alpha  (0) + \dot x_\alpha
1420 < (0)}}{{p^2  + \omega _\alpha ^2 }} - \frac{{g_\alpha ^2 }}{{m_\alpha
1421 < }}\omega _\alpha ^2 L(x)} \right\}}
1422 < %
1423 < &= - \frac{1}{p}\frac{{\partial W(x)}}{{\partial x}} -
1424 < \sum\limits_{\alpha  = 1}^N {\left\{ { - \frac{{g_\alpha ^2 }}{{m_\alpha  \omega _\alpha ^2 }}\frac{p}{{p^2  + \omega _\alpha ^2 }}pL(x)
1425 < - \frac{p}{{p^2  + \omega _\alpha ^2 }}g_\alpha  x_\alpha  (0)
1426 < - \frac{1}{{p^2  + \omega _\alpha ^2 }}g_\alpha  \dot x_\alpha  (0)} \right\}}
1427 < \end{align}
1428 < Then, the inverse transform,
1415 > with
1416 > \[
1417 > e^{\Delta tR_1 }  = \left( {\begin{array}{*{20}c}
1418 >   0 & 0 & 0  \\
1419 >   0 & {\cos \theta _1 } & {\sin \theta _1 }  \\
1420 >   0 & { - \sin \theta _1 } & {\cos \theta _1 }  \\
1421 > \end{array}} \right),\theta _1  = \frac{{\pi _1 }}{{I_1 }}\Delta t.
1422 > \]
1423 > To reduce the cost of computing expensive functions in $e^{\Delta
1424 > tR_1 }$, we can use Cayley transformation,
1425 > \[
1426 > e^{\Delta tR_1 }  \approx (1 - \Delta tR_1 )^{ - 1} (1 + \Delta tR_1
1427 > )
1428 > \]
1429 > The flow maps for $T_2^r$ and $T_3^r$ can be found in the same
1430 > manner.
1431  
1432 + In order to construct a second-order symplectic method, we split the
1433 + angular kinetic Hamiltonian function can into five terms
1434 + \[
1435 + T^r (\pi ) = \frac{1}{2}T_1 ^r (\pi _1 ) + \frac{1}{2}T_2^r (\pi _2
1436 + ) + T_3^r (\pi _3 ) + \frac{1}{2}T_2^r (\pi _2 ) + \frac{1}{2}T_1 ^r
1437 + (\pi _1 )
1438 + \].
1439 + Concatenating flows corresponding to these five terms, we can obtain
1440 + an symplectic integrator,
1441 + \[
1442 + \varphi _{\Delta t,T^r }  = \varphi _{\Delta t/2,\pi _1 }  \circ
1443 + \varphi _{\Delta t/2,\pi _2 }  \circ \varphi _{\Delta t,\pi _3 }
1444 + \circ \varphi _{\Delta t/2,\pi _2 }  \circ \varphi _{\Delta t/2,\pi
1445 + _1 }.
1446 + \]
1447 +
1448 + The non-canonical Lie-Poisson bracket ${F, G}$ of two function
1449 + $F(\pi )$ and $G(\pi )$ is defined by
1450 + \[
1451 + \{ F,G\} (\pi ) = [\nabla _\pi  F(\pi )]^T J(\pi )\nabla _\pi  G(\pi
1452 + )
1453 + \]
1454 + If the Poisson bracket of a function $F$ with an arbitrary smooth
1455 + function $G$ is zero, $F$ is a \emph{Casimir}, which is the
1456 + conserved quantity in Poisson system. We can easily verify that the
1457 + norm of the angular momentum, $\parallel \pi
1458 + \parallel$, is a \emph{Casimir}. Let$ F(\pi ) = S(\frac{{\parallel
1459 + \pi \parallel ^2 }}{2})$ for an arbitrary function $ S:R \to R$ ,
1460 + then by the chain rule
1461 + \[
1462 + \nabla _\pi  F(\pi ) = S'(\frac{{\parallel \pi \parallel ^2
1463 + }}{2})\pi
1464 + \]
1465 + Thus $ [\nabla _\pi  F(\pi )]^T J(\pi ) =  - S'(\frac{{\parallel \pi
1466 + \parallel ^2 }}{2})\pi  \times \pi  = 0 $. This explicit
1467 + Lie-Poisson integrator is found to be extremely efficient and stable
1468 + which can be explained by the fact the small angle approximation is
1469 + used and the norm of the angular momentum is conserved.
1470 +
1471 + \subsection{\label{introSection:RBHamiltonianSplitting} Hamiltonian
1472 + Splitting for Rigid Body}
1473 +
1474 + The Hamiltonian of rigid body can be separated in terms of kinetic
1475 + energy and potential energy,
1476 + \[
1477 + H = T(p,\pi ) + V(q,Q)
1478 + \]
1479 + The equations of motion corresponding to potential energy and
1480 + kinetic energy are listed in the below table,
1481 + \begin{center}
1482 + \begin{tabular}{|l|l|}
1483 +  \hline
1484 +  % after \\: \hline or \cline{col1-col2} \cline{col3-col4} ...
1485 +  Potential & Kinetic \\
1486 +  $\frac{{dq}}{{dt}} = \frac{p}{m}$ & $\frac{d}{{dt}}q = p$ \\
1487 +  $\frac{d}{{dt}}p =  - \frac{{\partial V}}{{\partial q}}$ & $ \frac{d}{{dt}}p = 0$ \\
1488 +  $\frac{d}{{dt}}Q = 0$ & $ \frac{d}{{dt}}Q = Qskew(I^{ - 1} j)$ \\
1489 +  $ \frac{d}{{dt}}\pi  = \sum\limits_i {\left( {Q^T F_i (r,Q)} \right) \times X_i }$ & $\frac{d}{{dt}}\pi  = \pi  \times I^{ - 1} \pi$\\
1490 +  \hline
1491 + \end{tabular}
1492 + \end{center}
1493 + A second-order symplectic method is now obtained by the composition
1494 + of the flow maps,
1495 + \[
1496 + \varphi _{\Delta t}  = \varphi _{\Delta t/2,V}  \circ \varphi
1497 + _{\Delta t,T}  \circ \varphi _{\Delta t/2,V}.
1498 + \]
1499 + Moreover, $\varphi _{\Delta t/2,V}$ can be divided into two
1500 + sub-flows which corresponding to force and torque respectively,
1501 + \[
1502 + \varphi _{\Delta t/2,V}  = \varphi _{\Delta t/2,F}  \circ \varphi
1503 + _{\Delta t/2,\tau }.
1504 + \]
1505 + Since the associated operators of $\varphi _{\Delta t/2,F} $ and
1506 + $\circ \varphi _{\Delta t/2,\tau }$ are commuted, the composition
1507 + order inside $\varphi _{\Delta t/2,V}$ does not matter.
1508 +
1509 + Furthermore, kinetic potential can be separated to translational
1510 + kinetic term, $T^t (p)$, and rotational kinetic term, $T^r (\pi )$,
1511 + \begin{equation}
1512 + T(p,\pi ) =T^t (p) + T^r (\pi ).
1513 + \end{equation}
1514 + where $ T^t (p) = \frac{1}{2}p^T m^{ - 1} p $ and $T^r (\pi )$ is
1515 + defined by \ref{introEquation:rotationalKineticRB}. Therefore, the
1516 + corresponding flow maps are given by
1517 + \[
1518 + \varphi _{\Delta t,T}  = \varphi _{\Delta t,T^t }  \circ \varphi
1519 + _{\Delta t,T^r }.
1520 + \]
1521 + Finally, we obtain the overall symplectic flow maps for free moving
1522 + rigid body
1523 + \begin{equation}
1524 + \begin{array}{c}
1525 + \varphi _{\Delta t}  = \varphi _{\Delta t/2,F}  \circ \varphi _{\Delta t/2,\tau }  \\
1526 +  \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 }  \\
1527 +  \circ \varphi _{\Delta t/2,\tau }  \circ \varphi _{\Delta t/2,F}  .\\
1528 + \end{array}
1529 + \label{introEquation:overallRBFlowMaps}
1530 + \end{equation}
1531 +
1532 + \section{\label{introSection:langevinDynamics}Langevin Dynamics}
1533 + As an alternative to newtonian dynamics, Langevin dynamics, which
1534 + mimics a simple heat bath with stochastic and dissipative forces,
1535 + has been applied in a variety of studies. This section will review
1536 + the theory of Langevin dynamics simulation. A brief derivation of
1537 + generalized Langevin equation will be given first. Follow that, we
1538 + will discuss the physical meaning of the terms appearing in the
1539 + equation as well as the calculation of friction tensor from
1540 + hydrodynamics theory.
1541 +
1542 + \subsection{\label{introSection:generalizedLangevinDynamics}Derivation of Generalized Langevin Equation}
1543 +
1544 + Harmonic bath model, in which an effective set of harmonic
1545 + oscillators are used to mimic the effect of a linearly responding
1546 + environment, has been widely used in quantum chemistry and
1547 + statistical mechanics. One of the successful applications of
1548 + Harmonic bath model is the derivation of Deriving Generalized
1549 + Langevin Dynamics. Lets consider a system, in which the degree of
1550 + freedom $x$ is assumed to couple to the bath linearly, giving a
1551 + Hamiltonian of the form
1552 + \begin{equation}
1553 + H = \frac{{p^2 }}{{2m}} + U(x) + H_B  + \Delta U(x,x_1 , \ldots x_N)
1554 + \label{introEquation:bathGLE}.
1555 + \end{equation}
1556 + Here $p$ is a momentum conjugate to $q$, $m$ is the mass associated
1557 + with this degree of freedom, $H_B$ is harmonic bath Hamiltonian,
1558 + \[
1559 + H_B  = \sum\limits_{\alpha  = 1}^N {\left\{ {\frac{{p_\alpha ^2
1560 + }}{{2m_\alpha  }} + \frac{1}{2}m_\alpha  \omega _\alpha ^2 }
1561 + \right\}}
1562 + \]
1563 + where the index $\alpha$ runs over all the bath degrees of freedom,
1564 + $\omega _\alpha$ are the harmonic bath frequencies, $m_\alpha$ are
1565 + the harmonic bath masses, and $\Delta U$ is bilinear system-bath
1566 + coupling,
1567 + \[
1568 + \Delta U =  - \sum\limits_{\alpha  = 1}^N {g_\alpha  x_\alpha  x}
1569 + \]
1570 + where $g_\alpha$ are the coupling constants between the bath and the
1571 + coordinate $x$. Introducing
1572 + \[
1573 + W(x) = U(x) - \sum\limits_{\alpha  = 1}^N {\frac{{g_\alpha ^2
1574 + }}{{2m_\alpha  w_\alpha ^2 }}} x^2
1575 + \] and combining the last two terms in Equation
1576 + \ref{introEquation:bathGLE}, we may rewrite the Harmonic bath
1577 + Hamiltonian as
1578 + \[
1579 + H = \frac{{p^2 }}{{2m}} + W(x) + \sum\limits_{\alpha  = 1}^N
1580 + {\left\{ {\frac{{p_\alpha ^2 }}{{2m_\alpha  }} + \frac{1}{2}m_\alpha
1581 + w_\alpha ^2 \left( {x_\alpha   - \frac{{g_\alpha  }}{{m_\alpha
1582 + w_\alpha ^2 }}x} \right)^2 } \right\}}
1583 + \]
1584 + Since the first two terms of the new Hamiltonian depend only on the
1585 + system coordinates, we can get the equations of motion for
1586 + Generalized Langevin Dynamics by Hamilton's equations
1587 + \ref{introEquation:motionHamiltonianCoordinate,
1588 + introEquation:motionHamiltonianMomentum},
1589 + \begin{equation}
1590 + m\ddot x =  - \frac{{\partial W(x)}}{{\partial x}} -
1591 + \sum\limits_{\alpha  = 1}^N {g_\alpha  \left( {x_\alpha   -
1592 + \frac{{g_\alpha  }}{{m_\alpha  w_\alpha ^2 }}x} \right)},
1593 + \label{introEquation:coorMotionGLE}
1594 + \end{equation}
1595 + and
1596 + \begin{equation}
1597 + m\ddot x_\alpha   =  - m_\alpha  w_\alpha ^2 \left( {x_\alpha   -
1598 + \frac{{g_\alpha  }}{{m_\alpha  w_\alpha ^2 }}x} \right).
1599 + \label{introEquation:bathMotionGLE}
1600 + \end{equation}
1601 +
1602 + In order to derive an equation for $x$, the dynamics of the bath
1603 + variables $x_\alpha$ must be solved exactly first. As an integral
1604 + transform which is particularly useful in solving linear ordinary
1605 + differential equations, Laplace transform is the appropriate tool to
1606 + solve this problem. The basic idea is to transform the difficult
1607 + differential equations into simple algebra problems which can be
1608 + solved easily. Then applying inverse Laplace transform, also known
1609 + as the Bromwich integral, we can retrieve the solutions of the
1610 + original problems.
1611 +
1612 + Let $f(t)$ be a function defined on $ [0,\infty ) $. The Laplace
1613 + transform of f(t) is a new function defined as
1614 + \[
1615 + L(f(t)) \equiv F(p) = \int_0^\infty  {f(t)e^{ - pt} dt}
1616 + \]
1617 + where  $p$ is real and  $L$ is called the Laplace Transform
1618 + Operator. Below are some important properties of Laplace transform
1619 + \begin{equation}
1620 + \begin{array}{c}
1621 + L(x + y) = L(x) + L(y) \\
1622 + L(ax) = aL(x) \\
1623 + L(\dot x) = pL(x) - px(0) \\
1624 + L(\ddot x) = p^2 L(x) - px(0) - \dot x(0) \\
1625 + L\left( {\int_0^t {g(t - \tau )h(\tau )d\tau } } \right) = G(p)H(p) \\
1626 + \end{array}
1627 + \end{equation}
1628 +
1629 + Applying Laplace transform to the bath coordinates, we obtain
1630 + \[
1631 + \begin{array}{c}
1632 + p^2 L(x_\alpha  ) - px_\alpha  (0) - \dot x_\alpha  (0) =  - \omega _\alpha ^2 L(x_\alpha  ) + \frac{{g_\alpha  }}{{\omega _\alpha  }}L(x) \\
1633 + L(x_\alpha  ) = \frac{{\frac{{g_\alpha  }}{{\omega _\alpha  }}L(x) + px_\alpha  (0) + \dot x_\alpha  (0)}}{{p^2  + \omega _\alpha ^2 }} \\
1634 + \end{array}
1635 + \]
1636 + By the same way, the system coordinates become
1637 + \[
1638 + \begin{array}{c}
1639 + mL(\ddot x) =  - \frac{1}{p}\frac{{\partial W(x)}}{{\partial x}} \\
1640 +  - \sum\limits_{\alpha  = 1}^N {\left\{ { - \frac{{g_\alpha ^2 }}{{m_\alpha  \omega _\alpha ^2 }}\frac{p}{{p^2  + \omega _\alpha ^2 }}pL(x) - \frac{p}{{p^2  + \omega _\alpha ^2 }}g_\alpha  x_\alpha  (0) - \frac{1}{{p^2  + \omega _\alpha ^2 }}g_\alpha  \dot x_\alpha  (0)} \right\}}  \\
1641 + \end{array}
1642 + \]
1643 +
1644 + With the help of some relatively important inverse Laplace
1645 + transformations:
1646 + \[
1647 + \begin{array}{c}
1648 + L(\cos at) = \frac{p}{{p^2  + a^2 }} \\
1649 + L(\sin at) = \frac{a}{{p^2  + a^2 }} \\
1650 + L(1) = \frac{1}{p} \\
1651 + \end{array}
1652 + \]
1653 + , we obtain
1654   \begin{align}
1655   m\ddot x &=  - \frac{{\partial W(x)}}{{\partial x}} -
1656   \sum\limits_{\alpha  = 1}^N {\left\{ {\left( { - \frac{{g_\alpha ^2
# Line 821 | Line 1670 | t)\dot x(t - \tau )d} \tau }  + \sum\limits_{\alpha  =
1670   (\omega _\alpha  t)} \right\}}
1671   \end{align}
1672  
1673 + Introducing a \emph{dynamic friction kernel}
1674   \begin{equation}
825 m\ddot x =  - \frac{{\partial W}}{{\partial x}} - \int_0^t {\xi
826 (t)\dot x(t - \tau )d\tau }  + R(t)
827 \label{introEuqation:GeneralizedLangevinDynamics}
828 \end{equation}
829 %where $ {\xi (t)}$ is friction kernel, $R(t)$ is random force and
830 %$W$ is the potential of mean force. $W(x) =  - kT\ln p(x)$
831 \[
1675   \xi (t) = \sum\limits_{\alpha  = 1}^N {\left( { - \frac{{g_\alpha ^2
1676   }}{{m_\alpha  \omega _\alpha ^2 }}} \right)\cos (\omega _\alpha  t)}
1677 < \]
1678 < For an infinite harmonic bath, we can use the spectral density and
1679 < an integral over frequencies.
1680 <
838 < \[
1677 > \label{introEquation:dynamicFrictionKernelDefinition}
1678 > \end{equation}
1679 > and \emph{a random force}
1680 > \begin{equation}
1681   R(t) = \sum\limits_{\alpha  = 1}^N {\left( {g_\alpha  x_\alpha  (0)
1682   - \frac{{g_\alpha ^2 }}{{m_\alpha  \omega _\alpha ^2 }}x(0)}
1683   \right)\cos (\omega _\alpha  t)}  + \frac{{\dot x_\alpha
1684 < (0)}}{{\omega _\alpha  }}\sin (\omega _\alpha  t)
1684 > (0)}}{{\omega _\alpha  }}\sin (\omega _\alpha  t),
1685 > \label{introEquation:randomForceDefinition}
1686 > \end{equation}
1687 > the equation of motion can be rewritten as
1688 > \begin{equation}
1689 > m\ddot x =  - \frac{{\partial W}}{{\partial x}} - \int_0^t {\xi
1690 > (t)\dot x(t - \tau )d\tau }  + R(t)
1691 > \label{introEuqation:GeneralizedLangevinDynamics}
1692 > \end{equation}
1693 > which is known as the \emph{generalized Langevin equation}.
1694 >
1695 > \subsubsection{\label{introSection:randomForceDynamicFrictionKernel}Random Force and Dynamic Friction Kernel}
1696 >
1697 > One may notice that $R(t)$ depends only on initial conditions, which
1698 > implies it is completely deterministic within the context of a
1699 > harmonic bath. However, it is easy to verify that $R(t)$ is totally
1700 > uncorrelated to $x$ and $\dot x$,
1701 > \[
1702 > \begin{array}{l}
1703 > \left\langle {x(t)R(t)} \right\rangle  = 0, \\
1704 > \left\langle {\dot x(t)R(t)} \right\rangle  = 0. \\
1705 > \end{array}
1706   \]
1707 < The random forces depend only on initial conditions.
1707 > This property is what we expect from a truly random process. As long
1708 > as the model, which is gaussian distribution in general, chosen for
1709 > $R(t)$ is a truly random process, the stochastic nature of the GLE
1710 > still remains.
1711  
1712 + %dynamic friction kernel
1713 + The convolution integral
1714 + \[
1715 + \int_0^t {\xi (t)\dot x(t - \tau )d\tau }
1716 + \]
1717 + depends on the entire history of the evolution of $x$, which implies
1718 + that the bath retains memory of previous motions. In other words,
1719 + the bath requires a finite time to respond to change in the motion
1720 + of the system. For a sluggish bath which responds slowly to changes
1721 + in the system coordinate, we may regard $\xi(t)$ as a constant
1722 + $\xi(t) = \Xi_0$. Hence, the convolution integral becomes
1723 + \[
1724 + \int_0^t {\xi (t)\dot x(t - \tau )d\tau }  = \xi _0 (x(t) - x(0))
1725 + \]
1726 + and Equation \ref{introEuqation:GeneralizedLangevinDynamics} becomes
1727 + \[
1728 + m\ddot x =  - \frac{\partial }{{\partial x}}\left( {W(x) +
1729 + \frac{1}{2}\xi _0 (x - x_0 )^2 } \right) + R(t),
1730 + \]
1731 + which can be used to describe dynamic caging effect. The other
1732 + extreme is the bath that responds infinitely quickly to motions in
1733 + the system. Thus, $\xi (t)$ can be taken as a $delta$ function in
1734 + time:
1735 + \[
1736 + \xi (t) = 2\xi _0 \delta (t)
1737 + \]
1738 + Hence, the convolution integral becomes
1739 + \[
1740 + \int_0^t {\xi (t)\dot x(t - \tau )d\tau }  = 2\xi _0 \int_0^t
1741 + {\delta (t)\dot x(t - \tau )d\tau }  = \xi _0 \dot x(t),
1742 + \]
1743 + and Equation \ref{introEuqation:GeneralizedLangevinDynamics} becomes
1744 + \begin{equation}
1745 + m\ddot x =  - \frac{{\partial W(x)}}{{\partial x}} - \xi _0 \dot
1746 + x(t) + R(t) \label{introEquation:LangevinEquation}
1747 + \end{equation}
1748 + which is known as the Langevin equation. The static friction
1749 + coefficient $\xi _0$ can either be calculated from spectral density
1750 + or be determined by Stokes' law for regular shaped particles.A
1751 + briefly review on calculating friction tensor for arbitrary shaped
1752 + particles is given in Sec.~\ref{introSection:frictionTensor}.
1753 +
1754   \subsubsection{\label{introSection:secondFluctuationDissipation}The Second Fluctuation Dissipation Theorem}
1755 < So we can define a new set of coordinates,
1755 >
1756 > Defining a new set of coordinates,
1757   \[
1758   q_\alpha  (t) = x_\alpha  (t) - \frac{1}{{m_\alpha  \omega _\alpha
1759   ^2 }}x(0)
1760 < \]
1761 < This makes
1760 > \],
1761 > we can rewrite $R(T)$ as
1762   \[
1763 < R(t) = \sum\limits_{\alpha  = 1}^N {g_\alpha  q_\alpha  (t)}
1763 > R(t) = \sum\limits_{\alpha  = 1}^N {g_\alpha  q_\alpha  (t)}.
1764   \]
1765   And since the $q$ coordinates are harmonic oscillators,
1766   \[
1767 < \begin{array}{l}
1767 > \begin{array}{c}
1768 > \left\langle {q_\alpha ^2 } \right\rangle  = \frac{{kT}}{{m_\alpha  \omega _\alpha ^2 }} \\
1769   \left\langle {q_\alpha  (t)q_\alpha  (0)} \right\rangle  = \left\langle {q_\alpha ^2 (0)} \right\rangle \cos (\omega _\alpha  t) \\
1770   \left\langle {q_\alpha  (t)q_\beta  (0)} \right\rangle  = \delta _{\alpha \beta } \left\langle {q_\alpha  (t)q_\alpha  (0)} \right\rangle  \\
1771 + \left\langle {R(t)R(0)} \right\rangle  = \sum\limits_\alpha  {\sum\limits_\beta  {g_\alpha  g_\beta  \left\langle {q_\alpha  (t)q_\beta  (0)} \right\rangle } }  \\
1772 +  = \sum\limits_\alpha  {g_\alpha ^2 \left\langle {q_\alpha ^2 (0)} \right\rangle \cos (\omega _\alpha  t)}  \\
1773 +  = kT\xi (t) \\
1774   \end{array}
1775   \]
1776 <
864 < \begin{align}
865 < \left\langle {R(t)R(0)} \right\rangle  &= \sum\limits_\alpha
866 < {\sum\limits_\beta  {g_\alpha  g_\beta  \left\langle {q_\alpha
867 < (t)q_\beta  (0)} \right\rangle } }
868 < %
869 < &= \sum\limits_\alpha  {g_\alpha ^2 \left\langle {q_\alpha ^2 (0)}
870 < \right\rangle \cos (\omega _\alpha  t)}
871 < %
872 < &= kT\xi (t)
873 < \end{align}
874 <
1776 > Thus, we recover the \emph{second fluctuation dissipation theorem}
1777   \begin{equation}
1778   \xi (t) = \left\langle {R(t)R(0)} \right\rangle
1779 < \label{introEquation:secondFluctuationDissipation}
1779 > \label{introEquation:secondFluctuationDissipation}.
1780   \end{equation}
1781 + In effect, it acts as a constraint on the possible ways in which one
1782 + can model the random force and friction kernel.
1783  
880 \section{\label{introSection:hydroynamics}Hydrodynamics}
881
1784   \subsection{\label{introSection:frictionTensor} Friction Tensor}
1785 < \subsection{\label{introSection:analyticalApproach}Analytical
1786 < Approach}
1785 > Theoretically, the friction kernel can be determined using velocity
1786 > autocorrelation function. However, this approach become impractical
1787 > when the system become more and more complicate. Instead, various
1788 > approaches based on hydrodynamics have been developed to calculate
1789 > the friction coefficients. The friction effect is isotropic in
1790 > Equation, \zeta can be taken as a scalar. In general, friction
1791 > tensor \Xi is a $6\times 6$ matrix given by
1792 > \[
1793 > \Xi  = \left( {\begin{array}{*{20}c}
1794 >   {\Xi _{}^{tt} } & {\Xi _{}^{rt} }  \\
1795 >   {\Xi _{}^{tr} } & {\Xi _{}^{rr} }  \\
1796 > \end{array}} \right).
1797 > \]
1798 > Here, $ {\Xi^{tt} }$ and $ {\Xi^{rr} }$ are translational friction
1799 > tensor and rotational resistance (friction) tensor respectively,
1800 > while ${\Xi^{tr} }$ is translation-rotation coupling tensor and $
1801 > {\Xi^{rt} }$ is rotation-translation coupling tensor. When a
1802 > particle moves in a fluid, it may experience friction force or
1803 > torque along the opposite direction of the velocity or angular
1804 > velocity,
1805 > \[
1806 > \left( \begin{array}{l}
1807 > F_R  \\
1808 > \tau _R  \\
1809 > \end{array} \right) =  - \left( {\begin{array}{*{20}c}
1810 >   {\Xi ^{tt} } & {\Xi ^{rt} }  \\
1811 >   {\Xi ^{tr} } & {\Xi ^{rr} }  \\
1812 > \end{array}} \right)\left( \begin{array}{l}
1813 > v \\
1814 > w \\
1815 > \end{array} \right)
1816 > \]
1817 > where $F_r$ is the friction force and $\tau _R$ is the friction
1818 > toque.
1819  
1820 < \subsection{\label{introSection:approximationApproach}Approximation
887 < Approach}
1820 > \subsubsection{\label{introSection:resistanceTensorRegular}The Resistance Tensor for Regular Shape}
1821  
1822 < \subsection{\label{introSection:centersRigidBody}Centers of Rigid
1823 < Body}
1822 > For a spherical particle, the translational and rotational friction
1823 > constant can be calculated from Stoke's law,
1824 > \[
1825 > \Xi ^{tt}  = \left( {\begin{array}{*{20}c}
1826 >   {6\pi \eta R} & 0 & 0  \\
1827 >   0 & {6\pi \eta R} & 0  \\
1828 >   0 & 0 & {6\pi \eta R}  \\
1829 > \end{array}} \right)
1830 > \]
1831 > and
1832 > \[
1833 > \Xi ^{rr}  = \left( {\begin{array}{*{20}c}
1834 >   {8\pi \eta R^3 } & 0 & 0  \\
1835 >   0 & {8\pi \eta R^3 } & 0  \\
1836 >   0 & 0 & {8\pi \eta R^3 }  \\
1837 > \end{array}} \right)
1838 > \]
1839 > where $\eta$ is the viscosity of the solvent and $R$ is the
1840 > hydrodynamics radius.
1841 >
1842 > Other non-spherical shape, such as cylinder and ellipsoid
1843 > \textit{etc}, are widely used as reference for developing new
1844 > hydrodynamics theory, because their properties can be calculated
1845 > exactly. In 1936, Perrin extended Stokes's law to general ellipsoid,
1846 > also called a triaxial ellipsoid, which is given in Cartesian
1847 > coordinates by
1848 > \[
1849 > \frac{{x^2 }}{{a^2 }} + \frac{{y^2 }}{{b^2 }} + \frac{{z^2 }}{{c^2
1850 > }} = 1
1851 > \]
1852 > where the semi-axes are of lengths $a$, $b$, and $c$. Unfortunately,
1853 > due to the complexity of the elliptic integral, only the ellipsoid
1854 > with the restriction of two axes having to be equal, \textit{i.e.}
1855 > prolate($ a \ge b = c$) and oblate ($ a < b = c $), can be solved
1856 > exactly. Introducing an elliptic integral parameter $S$ for prolate,
1857 > \[
1858 > S = \frac{2}{{\sqrt {a^2  - b^2 } }}\ln \frac{{a + \sqrt {a^2  - b^2
1859 > } }}{b},
1860 > \]
1861 > and oblate,
1862 > \[
1863 > S = \frac{2}{{\sqrt {b^2  - a^2 } }}arctg\frac{{\sqrt {b^2  - a^2 }
1864 > }}{a}
1865 > \],
1866 > one can write down the translational and rotational resistance
1867 > tensors
1868 > \[
1869 > \begin{array}{l}
1870 > \Xi _a^{tt}  = 16\pi \eta \frac{{a^2  - b^2 }}{{(2a^2  - b^2 )S - 2a}} \\
1871 > \Xi _b^{tt}  = \Xi _c^{tt}  = 32\pi \eta \frac{{a^2  - b^2 }}{{(2a^2  - 3b^2 )S + 2a}} \\
1872 > \end{array},
1873 > \]
1874 > and
1875 > \[
1876 > \begin{array}{l}
1877 > \Xi _a^{rr}  = \frac{{32\pi }}{3}\eta \frac{{(a^2  - b^2 )b^2 }}{{2a - b^2 S}} \\
1878 > \Xi _b^{rr}  = \Xi _c^{rr}  = \frac{{32\pi }}{3}\eta \frac{{(a^4  - b^4 )}}{{(2a^2  - b^2 )S - 2a}} \\
1879 > \end{array}.
1880 > \]
1881 >
1882 > \subsubsection{\label{introSection:resistanceTensorRegularArbitrary}The Resistance Tensor for Arbitrary Shape}
1883 >
1884 > Unlike spherical and other regular shaped molecules, there is not
1885 > analytical solution for friction tensor of any arbitrary shaped
1886 > rigid molecules. The ellipsoid of revolution model and general
1887 > triaxial ellipsoid model have been used to approximate the
1888 > hydrodynamic properties of rigid bodies. However, since the mapping
1889 > from all possible ellipsoidal space, $r$-space, to all possible
1890 > combination of rotational diffusion coefficients, $D$-space is not
1891 > unique\cite{Wegener79} as well as the intrinsic coupling between
1892 > translational and rotational motion of rigid body\cite{}, general
1893 > ellipsoid is not always suitable for modeling arbitrarily shaped
1894 > rigid molecule. A number of studies have been devoted to determine
1895 > the friction tensor for irregularly shaped rigid bodies using more
1896 > advanced method\cite{} where the molecule of interest was modeled by
1897 > combinations of spheres(beads)\cite{} and the hydrodynamics
1898 > properties of the molecule can be calculated using the hydrodynamic
1899 > interaction tensor. Let us consider a rigid assembly of $N$ beads
1900 > immersed in a continuous medium. Due to hydrodynamics interaction,
1901 > the ``net'' velocity of $i$th bead, $v'_i$ is different than its
1902 > unperturbed velocity $v_i$,
1903 > \[
1904 > v'_i  = v_i  - \sum\limits_{j \ne i} {T_{ij} F_j }
1905 > \]
1906 > where $F_i$ is the frictional force, and $T_{ij}$ is the
1907 > hydrodynamic interaction tensor. The friction force of $i$th bead is
1908 > proportional to its ``net'' velocity
1909 > \begin{equation}
1910 > F_i  = \zeta _i v_i  - \zeta _i \sum\limits_{j \ne i} {T_{ij} F_j }.
1911 > \label{introEquation:tensorExpression}
1912 > \end{equation}
1913 > This equation is the basis for deriving the hydrodynamic tensor. In
1914 > 1930, Oseen and Burgers gave a simple solution to Equation
1915 > \ref{introEquation:tensorExpression}
1916 > \begin{equation}
1917 > T_{ij}  = \frac{1}{{8\pi \eta r_{ij} }}\left( {I + \frac{{R_{ij}
1918 > R_{ij}^T }}{{R_{ij}^2 }}} \right).
1919 > \label{introEquation:oseenTensor}
1920 > \end{equation}
1921 > Here $R_{ij}$ is the distance vector between bead $i$ and bead $j$.
1922 > A second order expression for element of different size was
1923 > introduced by Rotne and Prager\cite{} and improved by Garc\'{i}a de
1924 > la Torre and Bloomfield,
1925 > \begin{equation}
1926 > T_{ij}  = \frac{1}{{8\pi \eta R_{ij} }}\left[ {\left( {I +
1927 > \frac{{R_{ij} R_{ij}^T }}{{R_{ij}^2 }}} \right) + R\frac{{\sigma
1928 > _i^2  + \sigma _j^2 }}{{r_{ij}^2 }}\left( {\frac{I}{3} -
1929 > \frac{{R_{ij} R_{ij}^T }}{{R_{ij}^2 }}} \right)} \right].
1930 > \label{introEquation:RPTensorNonOverlapped}
1931 > \end{equation}
1932 > Both of the Equation \ref{introEquation:oseenTensor} and Equation
1933 > \ref{introEquation:RPTensorNonOverlapped} have an assumption $R_{ij}
1934 > \ge \sigma _i  + \sigma _j$. An alternative expression for
1935 > overlapping beads with the same radius, $\sigma$, is given by
1936 > \begin{equation}
1937 > T_{ij}  = \frac{1}{{6\pi \eta R_{ij} }}\left[ {\left( {1 -
1938 > \frac{2}{{32}}\frac{{R_{ij} }}{\sigma }} \right)I +
1939 > \frac{2}{{32}}\frac{{R_{ij} R_{ij}^T }}{{R_{ij} \sigma }}} \right]
1940 > \label{introEquation:RPTensorOverlapped}
1941 > \end{equation}
1942 >
1943 > To calculate the resistance tensor at an arbitrary origin $O$, we
1944 > construct a $3N \times 3N$ matrix consisting of $N \times N$
1945 > $B_{ij}$ blocks
1946 > \begin{equation}
1947 > B = \left( {\begin{array}{*{20}c}
1948 >   {B_{11} } &  \ldots  & {B_{1N} }  \\
1949 >    \vdots  &  \ddots  &  \vdots   \\
1950 >   {B_{N1} } &  \cdots  & {B_{NN} }  \\
1951 > \end{array}} \right),
1952 > \end{equation}
1953 > where $B_{ij}$ is given by
1954 > \[
1955 > B_{ij}  = \delta _{ij} \frac{I}{{6\pi \eta R}} + (1 - \delta _{ij}
1956 > )T_{ij}
1957 > \]
1958 > where $\delta _{ij}$ is Kronecker delta function. Inverting matrix
1959 > $B$, we obtain
1960 >
1961 > \[
1962 > C = B^{ - 1}  = \left( {\begin{array}{*{20}c}
1963 >   {C_{11} } &  \ldots  & {C_{1N} }  \\
1964 >    \vdots  &  \ddots  &  \vdots   \\
1965 >   {C_{N1} } &  \cdots  & {C_{NN} }  \\
1966 > \end{array}} \right)
1967 > \]
1968 > , which can be partitioned into $N \times N$ $3 \times 3$ block
1969 > $C_{ij}$. With the help of $C_{ij}$ and skew matrix $U_i$
1970 > \[
1971 > U_i  = \left( {\begin{array}{*{20}c}
1972 >   0 & { - z_i } & {y_i }  \\
1973 >   {z_i } & 0 & { - x_i }  \\
1974 >   { - y_i } & {x_i } & 0  \\
1975 > \end{array}} \right)
1976 > \]
1977 > where $x_i$, $y_i$, $z_i$ are the components of the vector joining
1978 > bead $i$ and origin $O$. Hence, the elements of resistance tensor at
1979 > arbitrary origin $O$ can be written as
1980 > \begin{equation}
1981 > \begin{array}{l}
1982 > \Xi _{}^{tt}  = \sum\limits_i {\sum\limits_j {C_{ij} } } , \\
1983 > \Xi _{}^{tr}  = \Xi _{}^{rt}  = \sum\limits_i {\sum\limits_j {U_i C_{ij} } } , \\
1984 > \Xi _{}^{rr}  =  - \sum\limits_i {\sum\limits_j {U_i C_{ij} } } U_j  \\
1985 > \end{array}
1986 > \label{introEquation:ResistanceTensorArbitraryOrigin}
1987 > \end{equation}
1988 >
1989 > The resistance tensor depends on the origin to which they refer. The
1990 > proper location for applying friction force is the center of
1991 > resistance (reaction), at which the trace of rotational resistance
1992 > tensor, $ \Xi ^{rr}$ reaches minimum. Mathematically, the center of
1993 > resistance is defined as an unique point of the rigid body at which
1994 > the translation-rotation coupling tensor are symmetric,
1995 > \begin{equation}
1996 > \Xi^{tr}  = \left( {\Xi^{tr} } \right)^T
1997 > \label{introEquation:definitionCR}
1998 > \end{equation}
1999 > Form Equation \ref{introEquation:ResistanceTensorArbitraryOrigin},
2000 > we can easily find out that the translational resistance tensor is
2001 > origin independent, while the rotational resistance tensor and
2002 > translation-rotation coupling resistance tensor depend on the
2003 > origin. Given resistance tensor at an arbitrary origin $O$, and a
2004 > vector ,$r_{OP}(x_{OP}, y_{OP}, z_{OP})$, from $O$ to $P$, we can
2005 > obtain the resistance tensor at $P$ by
2006 > \begin{equation}
2007 > \begin{array}{l}
2008 > \Xi _P^{tt}  = \Xi _O^{tt}  \\
2009 > \Xi _P^{tr}  = \Xi _P^{rt}  = \Xi _O^{tr}  - U_{OP} \Xi _O^{tt}  \\
2010 > \Xi _P^{rr}  = \Xi _O^{rr}  - U_{OP} \Xi _O^{tt} U_{OP}  + \Xi _O^{tr} U_{OP}  - U_{OP} \Xi _O^{tr} ^{^T }  \\
2011 > \end{array}
2012 > \label{introEquation:resistanceTensorTransformation}
2013 > \end{equation}
2014 > where
2015 > \[
2016 > U_{OP}  = \left( {\begin{array}{*{20}c}
2017 >   0 & { - z_{OP} } & {y_{OP} }  \\
2018 >   {z_i } & 0 & { - x_{OP} }  \\
2019 >   { - y_{OP} } & {x_{OP} } & 0  \\
2020 > \end{array}} \right)
2021 > \]
2022 > Using Equations \ref{introEquation:definitionCR} and
2023 > \ref{introEquation:resistanceTensorTransformation}, one can locate
2024 > the position of center of resistance,
2025 > \[
2026 > \left( \begin{array}{l}
2027 > x_{OR}  \\
2028 > y_{OR}  \\
2029 > z_{OR}  \\
2030 > \end{array} \right) = \left( {\begin{array}{*{20}c}
2031 >   {(\Xi _O^{rr} )_{yy}  + (\Xi _O^{rr} )_{zz} } & { - (\Xi _O^{rr} )_{xy} } & { - (\Xi _O^{rr} )_{xz} }  \\
2032 >   { - (\Xi _O^{rr} )_{xy} } & {(\Xi _O^{rr} )_{zz}  + (\Xi _O^{rr} )_{xx} } & { - (\Xi _O^{rr} )_{yz} }  \\
2033 >   { - (\Xi _O^{rr} )_{xz} } & { - (\Xi _O^{rr} )_{yz} } & {(\Xi _O^{rr} )_{xx}  + (\Xi _O^{rr} )_{yy} }  \\
2034 > \end{array}} \right)^{ - 1} \left( \begin{array}{l}
2035 > (\Xi _O^{tr} )_{yz}  - (\Xi _O^{tr} )_{zy}  \\
2036 > (\Xi _O^{tr} )_{zx}  - (\Xi _O^{tr} )_{xz}  \\
2037 > (\Xi _O^{tr} )_{xy}  - (\Xi _O^{tr} )_{yx}  \\
2038 > \end{array} \right).
2039 > \]
2040 > where $x_OR$, $y_OR$, $z_OR$ are the components of the vector
2041 > joining center of resistance $R$ and origin $O$.

Diff Legend

Removed lines
+ Added lines
< Changed lines
> Changed lines