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 2791 by tim, Mon Jun 5 21:21:27 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 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 396 | Line 395 | Liouville's theorem can be expresses in a variety of d
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 444 | 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 458 | 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 472 | 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 539 | 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 $x$.
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}
575  
576 < \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}
576 > \subsection{\label{introSection:exactFlow}Exact Flow}
577  
563 \subsection{\label{introSection:geometricProperties}Geometric Properties}
578   Let $x(t)$ be the exact solution of the ODE system,
579   \begin{equation}
580   \frac{{dx}}{{dt}} = f(x) \label{introEquation:ODE}
# Line 570 | Line 584 | space to itself. In most cases, it is not easy to find
584   x(t+\tau) =\varphi_\tau(x(t))
585   \]
586   where $\tau$ is a fixed time step and $\varphi$ is a map from phase
587 < 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 $p$,
587 > space to itself. The flow has the continuous group property,
588   \begin{equation}
589 + \varphi _{\tau _1 }  \circ \varphi _{\tau _2 }  = \varphi _{\tau _1
590 + + \tau _2 } .
591 + \end{equation}
592 + In particular,
593 + \begin{equation}
594 + \varphi _\tau   \circ \varphi _{ - \tau }  = I
595 + \end{equation}
596 + Therefore, the exact flow is self-adjoint,
597 + \begin{equation}
598 + \varphi _\tau   = \varphi _{ - \tau }^{ - 1}.
599 + \end{equation}
600 + The exact flow can also be written in terms of the of an operator,
601 + \begin{equation}
602 + \varphi _\tau  (x) = e^{\tau \sum\limits_i {f_i (x)\frac{\partial
603 + }{{\partial x_i }}} } (x) \equiv \exp (\tau f)(x).
604 + \label{introEquation:exponentialOperator}
605 + \end{equation}
606 +
607 + In most cases, it is not easy to find the exact flow $\varphi_\tau$.
608 + Instead, we use a approximate map, $\psi_\tau$, which is usually
609 + called integrator. The order of an integrator $\psi_\tau$ is $p$, if
610 + the Taylor series of $\psi_\tau$ agree to order $p$,
611 + \begin{equation}
612   \psi_tau(x) = x + \tau f(x) + O(\tau^{p+1})
613   \end{equation}
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,
615 > \subsection{\label{introSection:geometricProperties}Geometric Properties}
616 >
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 591 | 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. It is possible
635 < to construct a \emph{volume-preserving} flow for a source free($
636 < \nabla \cdot f = 0 $) ODE, if the flow satisfies $ \det d\varphi  =
637 < 1$. Changing the variables $y = h(x)$ in a
638 < ODE\ref{introEquation:ODE} will result in a new system,
634 > is the property must be preserved by the integrator.
635 >
636 > It is possible to construct a \emph{volume-preserving} flow for a
637 > source free($ \nabla \cdot f = 0 $) ODE, if the flow satisfies $
638 > \det d\varphi  = 1$. One can show easily that a symplectic flow will
639 > be volume-preserving.
640 >
641 > Changing the variables $y = h(x)$ in a ODE\ref{introEquation:ODE}
642 > will result in a new system,
643   \[
644   \dot y = \tilde f(y) = ((dh \cdot f)h^{ - 1} )(y).
645   \]
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 $. When
607 < designing any numerical methods, one should always try to preserve
608 < the structural properties of the original ODE and its flow.
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 +
688   \subsection{\label{introSection:constructionSymplectic}Construction of Symplectic Methods}
689   A lot of well established and very effective numerical methods have
690   been successful precisely because of their symplecticities even
# Line 621 | 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\cite{}. In dissipative systems, variational
704 < methods can capture the decay of energy accurately\cite{}. Since
705 < their geometrically unstable nature against non-Hamiltonian
706 < perturbations, ordinary implicit Runge-Kutta methods are not
707 < suitable for Hamiltonian system. Recently, various high-order
708 < explicit 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}. The main idea behind splitting
714 < methods is to decompose the discrete $\varphi_h$ as a composition of
715 < simpler flows,
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 >
718 > The main idea behind splitting methods is to decompose the discrete
719 > $\varphi_h$ as a composition of simpler flows,
720   \begin{equation}
721   \varphi _h  = \varphi _{h_1 }  \circ \varphi _{h_2 }  \ldots  \circ
722   \varphi _{h_n }
723   \label{introEquation:FlowDecomposition}
724   \end{equation}
725   where each of the sub-flow is chosen such that each represent a
726 < simpler integration of the system. Let $\phi$ and $\psi$ both be
727 < symplectic maps, it is easy to show that any composition of
728 < symplectic flows yields a symplectic map,
726 > simpler integration of the system.
727 >
728 > Suppose that a Hamiltonian system takes the form,
729 > \[
730 > H = H_1 + H_2.
731 > \]
732 > Here, $H_1$ and $H_2$ may represent different physical processes of
733 > the system. For instance, they may relate to kinetic and potential
734 > energy respectively, which is a natural decomposition of the
735 > problem. If $H_1$ and $H_2$ can be integrated using exact flows
736 > $\varphi_1(t)$ and $\varphi_2(t)$, respectively, a simple first
737 > order is then given by the Lie-Trotter formula
738   \begin{equation}
739 + \varphi _h  = \varphi _{1,h}  \circ \varphi _{2,h},
740 + \label{introEquation:firstOrderSplitting}
741 + \end{equation}
742 + where $\varphi _h$ is the result of applying the corresponding
743 + continuous $\varphi _i$ over a time $h$. By definition, as
744 + $\varphi_i(t)$ is the exact solution of a Hamiltonian system, it
745 + must follow that each operator $\varphi_i(t)$ is a symplectic map.
746 + It is easy to show that any composition of symplectic flows yields a
747 + symplectic map,
748 + \begin{equation}
749   (\varphi '\phi ')^T J\varphi '\phi ' = \phi '^T \varphi '^T J\varphi
750 < '\phi ' = \phi '^T J\phi ' = J.
750 > '\phi ' = \phi '^T J\phi ' = J,
751   \label{introEquation:SymplecticFlowComposition}
752   \end{equation}
753 < Suppose that a Hamiltonian system has a form with $H = T + V$
753 > where $\phi$ and $\psi$ both are symplectic maps. Thus operator
754 > splitting in this context automatically generates a symplectic map.
755  
756 < \section{\label{introSection:molecularDynamics}Molecular Dynamics}
756 > The Lie-Trotter splitting(\ref{introEquation:firstOrderSplitting})
757 > introduces local errors proportional to $h^2$, while Strang
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} , \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 > \label{introEquation:timeReversible}
769 > \end{equation}
770  
771 < As a special discipline of molecular modeling, Molecular dynamics
772 < has proven to be a powerful tool for studying the functions of
773 < biological systems, providing structural, thermodynamic and
774 < dynamical information.
771 > \subsubsection{\label{introSection:exampleSplittingMethod}Example of Splitting Method}
772 > The classical equation for a system consisting of interacting
773 > particles can be written in Hamiltonian form,
774 > \[
775 > H = T + V
776 > \]
777 > where $T$ is the kinetic energy and $V$ is the potential energy.
778 > Setting $H_1 = T, H_2 = V$ and applying Strang splitting, one
779 > obtains the following:
780 > \begin{align}
781 > q(\Delta t) &= q(0) + \dot{q}(0)\Delta t +
782 >    \frac{F[q(0)]}{m}\frac{\Delta t^2}{2}, %
783 > \label{introEquation:Lp10a} \\%
784 > %
785 > \dot{q}(\Delta t) &= \dot{q}(0) + \frac{\Delta t}{2m}
786 >    \biggl [F[q(0)] + F[q(\Delta t)] \biggr]. %
787 > \label{introEquation:Lp10b}
788 > \end{align}
789 > where $F(t)$ is the force at time $t$. This integration scheme is
790 > known as \emph{velocity verlet} which is
791 > symplectic(\ref{introEquation:SymplecticFlowComposition}),
792 > time-reversible(\ref{introEquation:timeReversible}) and
793 > volume-preserving (\ref{introEquation:volumePreserving}). These
794 > geometric properties attribute to its long-time stability and its
795 > popularity in the community. However, the most commonly used
796 > velocity verlet integration scheme is written as below,
797 > \begin{align}
798 > \dot{q}\biggl (\frac{\Delta t}{2}\biggr ) &=
799 >    \dot{q}(0) + \frac{\Delta t}{2m}\, F[q(0)], \label{introEquation:Lp9a}\\%
800 > %
801 > q(\Delta t) &= q(0) + \Delta t\, \dot{q}\biggl (\frac{\Delta t}{2}\biggr ),%
802 >    \label{introEquation:Lp9b}\\%
803 > %
804 > \dot{q}(\Delta t) &= \dot{q}\biggl (\frac{\Delta t}{2}\biggr ) +
805 >    \frac{\Delta t}{2m}\, F[q(0)]. \label{introEquation:Lp9c}
806 > \end{align}
807 > From the preceding splitting, one can see that the integration of
808 > the equations of motion would follow:
809 > \begin{enumerate}
810 > \item calculate the velocities at the half step, $\frac{\Delta t}{2}$, from the forces calculated at the initial position.
811  
812 < \subsection{\label{introSec:mdInit}Initialization}
812 > \item Use the half step velocities to move positions one whole step, $\Delta t$.
813  
814 < \subsection{\label{introSection:mdIntegration} Integration of the Equations of Motion}
814 > \item Evaluate the forces at the new positions, $\mathbf{r}(\Delta t)$, and use the new forces to complete the velocity move.
815  
816 < \section{\label{introSection:rigidBody}Dynamics of Rigid Bodies}
816 > \item Repeat from step 1 with the new position, velocities, and forces assuming the roles of the initial values.
817 > \end{enumerate}
818  
819 < A rigid body is a body in which the distance between any two given
820 < points of a rigid body remains constant regardless of external
821 < forces exerted on it. A rigid body therefore conserves its shape
822 < during its motion.
819 > Simply switching the order of splitting and composing, a new
820 > integrator, the \emph{position verlet} integrator, can be generated,
821 > \begin{align}
822 > \dot q(\Delta t) &= \dot q(0) + \Delta tF(q(0))\left[ {q(0) +
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
827 > q(\Delta t)} \right]. %
828 > \label{introEquation:positionVerlet2}
829 > \end{align}
830  
831 < Applications of dynamics of rigid bodies.
831 > \subsubsection{\label{introSection:errorAnalysis}Error Analysis and Higher Order Methods}
832  
833 < \subsection{\label{introSection:lieAlgebra}Lie Algebra}
834 <
835 < \subsection{\label{introSection:DLMMotionEquation}The Euler Equations of Rigid Body Motion}
836 <
837 < \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 <
833 > Baker-Campbell-Hausdorff formula can be used to determine the local
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
838   \begin{equation}
839 < H = \frac{{p^2 }}{{2m}} + U(x) + H_B  + \Delta U(x,x_1 , \ldots x_N)
692 < \label{introEquation:bathGLE}
839 > \exp (hX + hY) = \exp (hZ)
840   \end{equation}
841 < where $H_B$ is harmonic bath Hamiltonian,
841 > where
842 > \begin{equation}
843 > hZ = hX + hY + \frac{{h^2 }}{2}[X,Y] + \frac{{h^3 }}{2}\left(
844 > {[X,[X,Y]] + [Y,[Y,X]]} \right) +  \ldots .
845 > \end{equation}
846 > Here, $[X,Y]$ is the commutators of operator $X$ and $Y$ given by
847   \[
848 < H_B  =\sum\limits_{\alpha  = 1}^N {\left\{ {\frac{{p_\alpha ^2
697 < }}{{2m_\alpha  }} + \frac{1}{2}m_\alpha  w_\alpha ^2 } \right\}}
848 > [X,Y] = XY - YX .
849   \]
850 < and $\Delta U$ is bilinear system-bath coupling,
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
860 > \begin{equation}
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
865 > order method. Yoshida proposed an elegant way to compose higher
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   \[
870 < \Delta U =  - \sum\limits_{\alpha  = 1}^N {g_\alpha  x_\alpha  x}
870 > \varphi _h^{(4)}  = \varphi _{\alpha h}^{(2)}  \circ \varphi _{\beta
871 > h}^{(2)}  \circ \varphi _{\alpha h}^{(2)}
872   \]
873 < Completing the square,
873 > where $ \alpha  =  - \frac{{2^{1/3} }}{{2 - 2^{1/3} }}$ and $ \beta
874 > = \frac{{2^{1/3} }}{{2 - 2^{1/3} }}$. Moreover, a symmetric
875 > integrator $ \varphi _h^{(2n + 2)}$ can be composed by
876 > \begin{equation}
877 > \varphi _h^{(2n + 2)}  = \varphi _{\alpha h}^{(2n)}  \circ \varphi
878 > _{\beta h}^{(2n)}  \circ \varphi _{\alpha h}^{(2n)}
879 > \end{equation}
880 > , if the weights are chosen as
881   \[
882 < H_B  + \Delta U = \sum\limits_{\alpha  = 1}^N {\left\{
883 < {\frac{{p_\alpha ^2 }}{{2m_\alpha  }} + \frac{1}{2}m_\alpha
707 < w_\alpha ^2 \left( {x_\alpha   - \frac{{g_\alpha  }}{{m_\alpha
708 < w_\alpha ^2 }}x} \right)^2 } \right\}}  - \sum\limits_{\alpha  =
709 < 1}^N {\frac{{g_\alpha ^2 }}{{2m_\alpha  w_\alpha ^2 }}} x^2
882 > \alpha  =  - \frac{{2^{1/(2n + 1)} }}{{2 - 2^{1/(2n + 1)} }},\beta =
883 > \frac{{2^{1/(2n + 1)} }}{{2 - 2^{1/(2n + 1)} }} .
884   \]
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}
885  
886 < \subsection{\label{introSection:laplaceTransform}The Laplace Transform}
886 > \section{\label{introSection:molecularDynamics}Molecular Dynamics}
887  
888 < \[
889 < L(x) = \int_0^\infty  {x(t)e^{ - pt} dt}
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 + 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{introSec:production} will
918 + discusses issues in production run. Sec.~\ref{introSection:Analysis}
919 + provides the theoretical tools for trajectory analysis.
920 +
921 + \subsection{\label{introSec:initialSystemSettings}Initialization}
922 +
923 + \subsubsection{Preliminary preparation}
924 +
925 + When selecting the starting structure of a molecule for molecular
926 + simulation, one may retrieve its Cartesian coordinates from public
927 + databases, such as RCSB Protein Data Bank \textit{etc}. Although
928 + thousands of crystal structures of molecules are discovered every
929 + year, many more remain unknown due to the difficulties of
930 + purification and crystallization. Even for the molecule with known
931 + structure, some important information is missing. For example, the
932 + missing hydrogen atom which acts as donor in hydrogen bonding must
933 + be added. Moreover, in order to include electrostatic interaction,
934 + one may need to specify the partial charges for individual atoms.
935 + Under some circumstances, we may even need to prepare the system in
936 + a special setup. For instance, when studying transport phenomenon in
937 + membrane system, we may prepare the lipids in bilayer structure
938 + instead of placing lipids randomly in solvent, since we are not
939 + interested in self-aggregation and it takes a long time to happen.
940 +
941 + \subsubsection{Minimization}
942 +
943 + It is quite possible that some of molecules in the system from
944 + preliminary preparation may be overlapped with each other. This
945 + close proximity leads to high potential energy which consequently
946 + jeopardizes any molecular dynamics simulations. To remove these
947 + steric overlaps, one typically performs energy minimization to find
948 + a more reasonable conformation. Several energy minimization methods
949 + have been developed to exploit the energy surface and to locate the
950 + local minimum. While converging slowly near the minimum, steepest
951 + descent method is extremely robust when systems are far from
952 + harmonic. Thus, it is often used to refine structure from
953 + crystallographic data. Relied on the gradient or hessian, advanced
954 + methods like conjugate gradient and Newton-Raphson converge rapidly
955 + to a local minimum, while become unstable if the energy surface is
956 + far from quadratic. Another factor must be taken into account, when
957 + choosing energy minimization method, is the size of the system.
958 + Steepest descent and conjugate gradient can deal with models of any
959 + size. Because of the limit of computation power to calculate hessian
960 + matrix and insufficient storage capacity to store them, most
961 + Newton-Raphson methods can not be used with very large models.
962 +
963 + \subsubsection{Heating}
964 +
965 + Typically, Heating is performed by assigning random velocities
966 + according to a Gaussian distribution for a temperature. Beginning at
967 + a lower temperature and gradually increasing the temperature by
968 + assigning greater random velocities, we end up with setting the
969 + temperature of the system to a final temperature at which the
970 + simulation will be conducted. In heating phase, we should also keep
971 + the system from drifting or rotating as a whole. Equivalently, the
972 + net linear momentum and angular momentum of the system should be
973 + shifted to zero.
974 +
975 + \subsubsection{Equilibration}
976 +
977 + The purpose of equilibration is to allow the system to evolve
978 + spontaneously for a period of time and reach equilibrium. The
979 + procedure is continued until various statistical properties, such as
980 + temperature, pressure, energy, volume and other structural
981 + properties \textit{etc}, become independent of time. Strictly
982 + speaking, minimization and heating are not necessary, provided the
983 + equilibration process is long enough. However, these steps can serve
984 + as a means to arrive at an equilibrated structure in an effective
985 + way.
986 +
987 + \subsection{\label{introSection:production}Production}
988 +
989 + Production run is the most important step of the simulation, in
990 + which the equilibrated structure is used as a starting point and the
991 + motions of the molecules are collected for later analysis. In order
992 + to capture the macroscopic properties of the system, the molecular
993 + dynamics simulation must be performed in correct and efficient way.
994 +
995 + The most expensive part of a molecular dynamics simulation is the
996 + calculation of non-bonded forces, such as van der Waals force and
997 + Coulombic forces \textit{etc}. For a system of $N$ particles, the
998 + complexity of the algorithm for pair-wise interactions is $O(N^2 )$,
999 + which making large simulations prohibitive in the absence of any
1000 + computation saving techniques.
1001 +
1002 + A natural approach to avoid system size issue is to represent the
1003 + bulk behavior by a finite number of the particles. However, this
1004 + approach will suffer from the surface effect. To offset this,
1005 + \textit{Periodic boundary condition} (see Fig.~\ref{introFig:pbc})
1006 + is developed to simulate bulk properties with a relatively small
1007 + number of particles. In this method, the simulation box is
1008 + replicated throughout space to form an infinite lattice. During the
1009 + simulation, when a particle moves in the primary cell, its image in
1010 + other cells move in exactly the same direction with exactly the same
1011 + orientation. Thus, as a particle leaves the primary cell, one of its
1012 + images will enter through the opposite face.
1013 + \begin{figure}
1014 + \centering
1015 + \includegraphics[width=\linewidth]{pbc.eps}
1016 + \caption[An illustration of periodic boundary conditions]{A 2-D
1017 + illustration of periodic boundary conditions. As one particle leaves
1018 + the left of the simulation box, an image of it enters the right.}
1019 + \label{introFig:pbc}
1020 + \end{figure}
1021 +
1022 + %cutoff and minimum image convention
1023 + Another important technique to improve the efficiency of force
1024 + evaluation is to apply cutoff where particles farther than a
1025 + predetermined distance, are not included in the calculation
1026 + \cite{Frenkel1996}. The use of a cutoff radius will cause a
1027 + discontinuity in the potential energy curve. Fortunately, one can
1028 + shift the potential to ensure the potential curve go smoothly to
1029 + zero at the cutoff radius. Cutoff strategy works pretty well for
1030 + Lennard-Jones interaction because of its short range nature.
1031 + However, simply truncating the electrostatic interaction with the
1032 + use of cutoff has been shown to lead to severe artifacts in
1033 + simulations. Ewald summation, in which the slowly conditionally
1034 + convergent Coulomb potential is transformed into direct and
1035 + reciprocal sums with rapid and absolute convergence, has proved to
1036 + minimize the periodicity artifacts in liquid simulations. Taking the
1037 + advantages of the fast Fourier transform (FFT) for calculating
1038 + discrete Fourier transforms, the particle mesh-based
1039 + methods\cite{Hockney1981,Shimada1993, Luty1994} are accelerated from
1040 + $O(N^{3/2})$ to $O(N logN)$. An alternative approach is \emph{fast
1041 + multipole method}\cite{Greengard1987, Greengard1994}, which treats
1042 + Coulombic interaction exactly at short range, and approximate the
1043 + potential at long range through multipolar expansion. In spite of
1044 + their wide acceptances at the molecular simulation community, these
1045 + two methods are hard to be implemented correctly and efficiently.
1046 + Instead, we use a damped and charge-neutralized Coulomb potential
1047 + method developed by Wolf and his coworkers\cite{Wolf1999}. The
1048 + shifted Coulomb potential for particle $i$ and particle $j$ at
1049 + distance $r_{rj}$ is given by:
1050 + \begin{equation}
1051 + V(r_{ij})= \frac{q_i q_j \textrm{erfc}(\alpha
1052 + r_{ij})}{r_{ij}}-\lim_{r_{ij}\rightarrow
1053 + R_\textrm{c}}\left\{\frac{q_iq_j \textrm{erfc}(\alpha
1054 + r_{ij})}{r_{ij}}\right\}. \label{introEquation:shiftedCoulomb}
1055 + \end{equation}
1056 + where $\alpha$ is the convergence parameter. Due to the lack of
1057 + inherent periodicity and rapid convergence,this method is extremely
1058 + efficient and easy to implement.
1059 + \begin{figure}
1060 + \centering
1061 + \includegraphics[width=\linewidth]{shifted_coulomb.eps}
1062 + \caption[An illustration of shifted Coulomb potential]{An
1063 + illustration of shifted Coulomb potential.}
1064 + \label{introFigure:shiftedCoulomb}
1065 + \end{figure}
1066 +
1067 + %multiple time step
1068 +
1069 + \subsection{\label{introSection:Analysis} Analysis}
1070 +
1071 + Recently, advanced visualization technique are widely applied to
1072 + monitor the motions of molecules. Although the dynamics of the
1073 + system can be described qualitatively from animation, quantitative
1074 + trajectory analysis are more appreciable. According to the
1075 + principles of Statistical Mechanics,
1076 + Sec.~\ref{introSection:statisticalMechanics}, one can compute
1077 + thermodynamics properties, analyze fluctuations of structural
1078 + parameters, and investigate time-dependent processes of the molecule
1079 + from the trajectories.
1080 +
1081 + \subsubsection{\label{introSection:thermodynamicsProperties}Thermodynamics Properties}
1082 +
1083 + Thermodynamics properties, which can be expressed in terms of some
1084 + function of the coordinates and momenta of all particles in the
1085 + system, can be directly computed from molecular dynamics. The usual
1086 + way to measure the pressure is based on virial theorem of Clausius
1087 + which states that the virial is equal to $-3Nk_BT$. For a system
1088 + with forces between particles, the total virial, $W$, contains the
1089 + contribution from external pressure and interaction between the
1090 + particles:
1091   \[
1092 < L(x + y) = L(x) + L(y)
1092 > W =  - 3PV + \left\langle {\sum\limits_{i < j} {r{}_{ij} \cdot
1093 > f_{ij} } } \right\rangle
1094   \]
1095 + where $f_{ij}$ is the force between particle $i$ and $j$ at a
1096 + distance $r_{ij}$. Thus, the expression for the pressure is given
1097 + by:
1098 + \begin{equation}
1099 + P = \frac{{Nk_B T}}{V} - \frac{1}{{3V}}\left\langle {\sum\limits_{i
1100 + < j} {r{}_{ij} \cdot f_{ij} } } \right\rangle
1101 + \end{equation}
1102  
1103 + \subsubsection{\label{introSection:structuralProperties}Structural Properties}
1104 +
1105 + Structural Properties of a simple fluid can be described by a set of
1106 + distribution functions. Among these functions,\emph{pair
1107 + distribution function}, also known as \emph{radial distribution
1108 + function}, is of most fundamental importance to liquid-state theory.
1109 + Pair distribution function can be gathered by Fourier transforming
1110 + raw data from a series of neutron diffraction experiments and
1111 + integrating over the surface factor \cite{Powles1973}. The
1112 + experiment result can serve as a criterion to justify the
1113 + correctness of the theory. Moreover, various equilibrium
1114 + thermodynamic and structural properties can also be expressed in
1115 + terms of radial distribution function \cite{Allen1987}.
1116 +
1117 + A pair distribution functions $g(r)$ gives the probability that a
1118 + particle $i$ will be located at a distance $r$ from a another
1119 + particle $j$ in the system
1120   \[
1121 < L(ax) = aL(x)
1121 > g(r) = \frac{V}{{N^2 }}\left\langle {\sum\limits_i {\sum\limits_{j
1122 > \ne i} {\delta (r - r_{ij} )} } } \right\rangle.
1123   \]
1124 + Note that the delta function can be replaced by a histogram in
1125 + computer simulation. Figure
1126 + \ref{introFigure:pairDistributionFunction} shows a typical pair
1127 + distribution function for the liquid argon system. The occurrence of
1128 + several peaks in the plot of $g(r)$ suggests that it is more likely
1129 + to find particles at certain radial values than at others. This is a
1130 + result of the attractive interaction at such distances. Because of
1131 + the strong repulsive forces at short distance, the probability of
1132 + locating particles at distances less than about 2.5{\AA} from each
1133 + other is essentially zero.
1134  
1135 + %\begin{figure}
1136 + %\centering
1137 + %\includegraphics[width=\linewidth]{pdf.eps}
1138 + %\caption[Pair distribution function for the liquid argon
1139 + %]{Pair distribution function for the liquid argon}
1140 + %\label{introFigure:pairDistributionFunction}
1141 + %\end{figure}
1142 +
1143 + \subsubsection{\label{introSection:timeDependentProperties}Time-dependent
1144 + Properties}
1145 +
1146 + Time-dependent properties are usually calculated using \emph{time
1147 + correlation function}, which correlates random variables $A$ and $B$
1148 + at two different time
1149 + \begin{equation}
1150 + C_{AB} (t) = \left\langle {A(t)B(0)} \right\rangle.
1151 + \label{introEquation:timeCorrelationFunction}
1152 + \end{equation}
1153 + If $A$ and $B$ refer to same variable, this kind of correlation
1154 + function is called \emph{auto correlation function}. One example of
1155 + auto correlation function is velocity auto-correlation function
1156 + which is directly related to transport properties of molecular
1157 + liquids:
1158   \[
1159 < L(\dot x) = pL(x) - px(0)
1159 > D = \frac{1}{3}\int\limits_0^\infty  {\left\langle {v(t) \cdot v(0)}
1160 > \right\rangle } dt
1161   \]
1162 + where $D$ is diffusion constant. Unlike velocity autocorrelation
1163 + function which is averaging over time origins and over all the
1164 + atoms, dipole autocorrelation are calculated for the entire system.
1165 + The dipole autocorrelation function is given by:
1166 + \[
1167 + c_{dipole}  = \left\langle {u_{tot} (t) \cdot u_{tot} (t)}
1168 + \right\rangle
1169 + \]
1170 + Here $u_{tot}$ is the net dipole of the entire system and is given
1171 + by
1172 + \[
1173 + u_{tot} (t) = \sum\limits_i {u_i (t)}
1174 + \]
1175 + In principle, many time correlation functions can be related with
1176 + Fourier transforms of the infrared, Raman, and inelastic neutron
1177 + scattering spectra of molecular liquids. In practice, one can
1178 + extract the IR spectrum from the intensity of dipole fluctuation at
1179 + each frequency using the following relationship:
1180 + \[
1181 + \hat c_{dipole} (v) = \int_{ - \infty }^\infty  {c_{dipole} (t)e^{ -
1182 + i2\pi vt} dt}
1183 + \]
1184  
1185 + \section{\label{introSection:rigidBody}Dynamics of Rigid Bodies}
1186 +
1187 + Rigid bodies are frequently involved in the modeling of different
1188 + areas, from engineering, physics, to chemistry. For example,
1189 + missiles and vehicle are usually modeled by rigid bodies.  The
1190 + movement of the objects in 3D gaming engine or other physics
1191 + simulator is governed by the rigid body dynamics. In molecular
1192 + simulation, rigid body is used to simplify the model in
1193 + protein-protein docking study\cite{Gray2003}.
1194 +
1195 + It is very important to develop stable and efficient methods to
1196 + integrate the equations of motion of orientational degrees of
1197 + freedom. Euler angles are the nature choice to describe the
1198 + rotational degrees of freedom. However, due to its singularity, the
1199 + numerical integration of corresponding equations of motion is very
1200 + inefficient and inaccurate. Although an alternative integrator using
1201 + different sets of Euler angles can overcome this
1202 + difficulty\cite{Barojas1973}, the computational penalty and the lost
1203 + of angular momentum conservation still remain. A singularity free
1204 + representation utilizing quaternions was developed by Evans in
1205 + 1977\cite{Evans1977}. Unfortunately, this approach suffer from the
1206 + nonseparable Hamiltonian resulted from quaternion representation,
1207 + which prevents the symplectic algorithm to be utilized. Another
1208 + different approach is to apply holonomic constraints to the atoms
1209 + belonging to the rigid body. Each atom moves independently under the
1210 + normal forces deriving from potential energy and constraint forces
1211 + which are used to guarantee the rigidness. However, due to their
1212 + iterative nature, SHAKE and Rattle algorithm converge very slowly
1213 + when the number of constraint increases\cite{Ryckaert1977,
1214 + Andersen1983}.
1215 +
1216 + The break through in geometric literature suggests that, in order to
1217 + develop a long-term integration scheme, one should preserve the
1218 + symplectic structure of the flow. Introducing conjugate momentum to
1219 + rotation matrix $Q$ and re-formulating Hamiltonian's equation, a
1220 + symplectic integrator, RSHAKE\cite{Kol1997}, was proposed to evolve
1221 + the Hamiltonian system in a constraint manifold by iteratively
1222 + satisfying the orthogonality constraint $Q_T Q = 1$. An alternative
1223 + method using quaternion representation was developed by
1224 + Omelyan\cite{Omelyan1998}. However, both of these methods are
1225 + iterative and inefficient. In this section, we will present a
1226 + symplectic Lie-Poisson integrator for rigid body developed by
1227 + Dullweber and his coworkers\cite{Dullweber1997} in depth.
1228 +
1229 + \subsection{\label{introSection:constrainedHamiltonianRB}Constrained Hamiltonian for Rigid Body}
1230 + The motion of the rigid body is Hamiltonian with the Hamiltonian
1231 + function
1232 + \begin{equation}
1233 + H = \frac{1}{2}(p^T m^{ - 1} p) + \frac{1}{2}tr(PJ^{ - 1} P) +
1234 + V(q,Q) + \frac{1}{2}tr[(QQ^T  - 1)\Lambda ].
1235 + \label{introEquation:RBHamiltonian}
1236 + \end{equation}
1237 + Here, $q$ and $Q$  are the position and rotation matrix for the
1238 + rigid-body, $p$ and $P$  are conjugate momenta to $q$  and $Q$ , and
1239 + $J$, a diagonal matrix, is defined by
1240   \[
1241 < L(\ddot x) = p^2 L(x) - px(0) - \dot x(0)
1241 > I_{ii}^{ - 1}  = \frac{1}{2}\sum\limits_{i \ne j} {J_{jj}^{ - 1} }
1242   \]
1243 + where $I_{ii}$ is the diagonal element of the inertia tensor. This
1244 + constrained Hamiltonian equation subjects to a holonomic constraint,
1245 + \begin{equation}
1246 + Q^T Q = 1, \label{introEquation:orthogonalConstraint}
1247 + \end{equation}
1248 + which is used to ensure rotation matrix's orthogonality.
1249 + Differentiating \ref{introEquation:orthogonalConstraint} and using
1250 + Equation \ref{introEquation:RBMotionMomentum}, one may obtain,
1251 + \begin{equation}
1252 + Q^T PJ^{ - 1}  + J^{ - 1} P^T Q = 0 . \\
1253 + \label{introEquation:RBFirstOrderConstraint}
1254 + \end{equation}
1255  
1256 + Using Equation (\ref{introEquation:motionHamiltonianCoordinate},
1257 + \ref{introEquation:motionHamiltonianMomentum}), one can write down
1258 + the equations of motion,
1259   \[
1260 < L\left( {\int_0^t {g(t - \tau )h(\tau )d\tau } } \right) = G(p)H(p)
1260 > \begin{array}{c}
1261 > \frac{{dq}}{{dt}} = \frac{p}{m} \label{introEquation:RBMotionPosition}\\
1262 > \frac{{dp}}{{dt}} =  - \nabla _q V(q,Q) \label{introEquation:RBMotionMomentum}\\
1263 > \frac{{dQ}}{{dt}} = PJ^{ - 1}  \label{introEquation:RBMotionRotation}\\
1264 > \frac{{dP}}{{dt}} =  - \nabla _Q V(q,Q) - 2Q\Lambda . \label{introEquation:RBMotionP}\\
1265 > \end{array}
1266   \]
1267  
1268 < Some relatively important transformation,
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   \[
1275 < L(\cos at) = \frac{p}{{p^2  + a^2 }}
1275 > M = \left\{ {(Q,P):Q^T Q = 1,Q^T PJ^{ - 1}  + J^{ - 1} P^T Q = 0}
1276 > \right\}.
1277   \]
1278  
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   \[
1284 < L(\sin at) = \frac{a}{{p^2  + a^2 }}
1284 > \tilde Q = Q,\tilde P = \frac{1}{2}\left( {P - QP^T Q} \right),
1285   \]
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   \[
1301 < L(1) = \frac{1}{p}
1301 > V(q,Q) = V(Q X_0 + q).
1302 > \]
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 > \[
1325 > \begin{array}{l}
1326 > \Pi J^{ - 1}  + J^{ - 1} \Pi ^t  = 0 \\
1327 > Q^T Q = 1 \\
1328 > \end{array}
1329   \]
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   \[
1345 < p^2 L(x_\alpha  ) - px_\alpha  (0) - \dot x_\alpha  (0) =  - \omega
783 < _\alpha ^2 L(x_\alpha  ) + \frac{{g_\alpha  }}{{\omega _\alpha
784 < }}L(x)
1345 > \hat vu = v \times u
1346   \]
1347 +
1348 + Using \ref{introEqaution:RBMotionPI}, one can construct a skew
1349 + matrix,
1350 + \begin{equation}
1351 + (\mathop \Pi \limits^ \bullet   - \mathop \Pi \limits^ \bullet  ^T
1352 + ){\rm{ }} = {\rm{ }}(\Pi  - \Pi ^T ){\rm{ }}(J^{ - 1} \Pi  + \Pi J^{
1353 + - 1} ) + \sum\limits_i {[Q^T F_i (r,Q)X_i^T  - X_i F_i (r,Q)^T Q]} -
1354 + (\Lambda  - \Lambda ^T ) . \label{introEquation:skewMatrixPI}
1355 + \end{equation}
1356 + Since $\Lambda$ is symmetric, the last term of Equation
1357 + \ref{introEquation:skewMatrixPI} is zero, which implies the Lagrange
1358 + multiplier $\Lambda$ is absent from the equations of motion. This
1359 + unique property eliminate the requirement of iterations which can
1360 + not be avoided in other methods\cite{Kol1997, Omelyan1998}.
1361 +
1362 + Applying hat-map isomorphism, we obtain the equation of motion for
1363 + angular momentum on body frame
1364 + \begin{equation}
1365 + \dot \pi  = \pi  \times I^{ - 1} \pi  + \sum\limits_i {\left( {Q^T
1366 + F_i (r,Q)} \right) \times X_i }.
1367 + \label{introEquation:bodyAngularMotion}
1368 + \end{equation}
1369 + In the same manner, the equation of motion for rotation matrix is
1370 + given by
1371   \[
1372 < L(x_\alpha  ) = \frac{{\frac{{g_\alpha  }}{{\omega _\alpha  }}L(x) +
788 < px_\alpha  (0) + \dot x_\alpha  (0)}}{{p^2  + \omega _\alpha ^2 }}
1372 > \dot Q = Qskew(I^{ - 1} \pi )
1373   \]
790 Then, the system coordinates,
791 \begin{align}
792 mL(\ddot x) &=  - \frac{1}{p}\frac{{\partial W(x)}}{{\partial x}} -
793 \sum\limits_{\alpha  = 1}^N {\left\{ {\frac{{\frac{{g_\alpha
794 }}{{\omega _\alpha  }}L(x) + px_\alpha  (0) + \dot x_\alpha
795 (0)}}{{p^2  + \omega _\alpha ^2 }} - \frac{{g_\alpha ^2 }}{{m_\alpha
796 }}\omega _\alpha ^2 L(x)} \right\}}
797 %
798 &= - \frac{1}{p}\frac{{\partial W(x)}}{{\partial x}} -
799 \sum\limits_{\alpha  = 1}^N {\left\{ { - \frac{{g_\alpha ^2 }}{{m_\alpha  \omega _\alpha ^2 }}\frac{p}{{p^2  + \omega _\alpha ^2 }}pL(x)
800 - \frac{p}{{p^2  + \omega _\alpha ^2 }}g_\alpha  x_\alpha  (0)
801 - \frac{1}{{p^2  + \omega _\alpha ^2 }}g_\alpha  \dot x_\alpha  (0)} \right\}}
802 \end{align}
803 Then, the inverse transform,
1374  
1375 < \begin{align}
1376 < m\ddot x &=  - \frac{{\partial W(x)}}{{\partial x}} -
807 < \sum\limits_{\alpha  = 1}^N {\left\{ {\left( { - \frac{{g_\alpha ^2
808 < }}{{m_\alpha  \omega _\alpha ^2 }}} \right)\int_0^t {\cos (\omega
809 < _\alpha  t)\dot x(t - \tau )d\tau  - \left[ {g_\alpha  x_\alpha  (0)
810 < - \frac{{g_\alpha  }}{{m_\alpha  \omega _\alpha  }}} \right]\cos
811 < (\omega _\alpha  t) - \frac{{g_\alpha  \dot x_\alpha  (0)}}{{\omega
812 < _\alpha  }}\sin (\omega _\alpha  t)} } \right\}}
813 < %
814 < &= - \frac{{\partial W(x)}}{{\partial x}} - \int_0^t
815 < {\sum\limits_{\alpha  = 1}^N {\left( { - \frac{{g_\alpha ^2
816 < }}{{m_\alpha  \omega _\alpha ^2 }}} \right)\cos (\omega _\alpha
817 < t)\dot x(t - \tau )d} \tau }  + \sum\limits_{\alpha  = 1}^N {\left\{
818 < {\left[ {g_\alpha  x_\alpha  (0) - \frac{{g_\alpha  }}{{m_\alpha
819 < \omega _\alpha  }}} \right]\cos (\omega _\alpha  t) +
820 < \frac{{g_\alpha  \dot x_\alpha  (0)}}{{\omega _\alpha  }}\sin
821 < (\omega _\alpha  t)} \right\}}
822 < \end{align}
1375 > \subsection{\label{introSection:SymplecticFreeRB}Symplectic
1376 > Lie-Poisson Integrator for Free Rigid Body}
1377  
1378 + If there is not external forces exerted on the rigid body, the only
1379 + contribution to the rotational is from the kinetic potential (the
1380 + first term of \ref{ introEquation:bodyAngularMotion}). The free
1381 + rigid body is an example of Lie-Poisson system with Hamiltonian
1382 + function
1383   \begin{equation}
1384 < m\ddot x =  - \frac{{\partial W}}{{\partial x}} - \int_0^t {\xi
1385 < (t)\dot x(t - \tau )d\tau }  + R(t)
827 < \label{introEuqation:GeneralizedLangevinDynamics}
1384 > T^r (\pi ) = T_1 ^r (\pi _1 ) + T_2^r (\pi _2 ) + T_3^r (\pi _3 )
1385 > \label{introEquation:rotationalKineticRB}
1386   \end{equation}
1387 < %where $ {\xi (t)}$ is friction kernel, $R(t)$ is random force and
1388 < %$W$ is the potential of mean force. $W(x) =  - kT\ln p(x)$
1387 > where $T_i^r (\pi _i ) = \frac{{\pi _i ^2 }}{{2I_i }}$ and
1388 > Lie-Poisson structure matrix,
1389 > \begin{equation}
1390 > J(\pi ) = \left( {\begin{array}{*{20}c}
1391 >   0 & {\pi _3 } & { - \pi _2 }  \\
1392 >   { - \pi _3 } & 0 & {\pi _1 }  \\
1393 >   {\pi _2 } & { - \pi _1 } & 0  \\
1394 > \end{array}} \right)
1395 > \end{equation}
1396 > Thus, the dynamics of free rigid body is governed by
1397 > \begin{equation}
1398 > \frac{d}{{dt}}\pi  = J(\pi )\nabla _\pi  T^r (\pi )
1399 > \end{equation}
1400 >
1401 > One may notice that each $T_i^r$ in Equation
1402 > \ref{introEquation:rotationalKineticRB} can be solved exactly. For
1403 > instance, the equations of motion due to $T_1^r$ are given by
1404 > \begin{equation}
1405 > \frac{d}{{dt}}\pi  = R_1 \pi ,\frac{d}{{dt}}Q = QR_1
1406 > \label{introEqaution:RBMotionSingleTerm}
1407 > \end{equation}
1408 > where
1409 > \[ R_1  = \left( {\begin{array}{*{20}c}
1410 >   0 & 0 & 0  \\
1411 >   0 & 0 & {\pi _1 }  \\
1412 >   0 & { - \pi _1 } & 0  \\
1413 > \end{array}} \right).
1414 > \]
1415 > The solutions of Equation \ref{introEqaution:RBMotionSingleTerm} is
1416   \[
1417 < \xi (t) = \sum\limits_{\alpha  = 1}^N {\left( { - \frac{{g_\alpha ^2
1418 < }}{{m_\alpha  \omega _\alpha ^2 }}} \right)\cos (\omega _\alpha  t)}
1417 > \pi (\Delta t) = e^{\Delta tR_1 } \pi (0),Q(\Delta t) =
1418 > Q(0)e^{\Delta tR_1 }
1419   \]
1420 < For an infinite harmonic bath, we can use the spectral density and
1421 < an integral over frequencies.
1420 > with
1421 > \[
1422 > e^{\Delta tR_1 }  = \left( {\begin{array}{*{20}c}
1423 >   0 & 0 & 0  \\
1424 >   0 & {\cos \theta _1 } & {\sin \theta _1 }  \\
1425 >   0 & { - \sin \theta _1 } & {\cos \theta _1 }  \\
1426 > \end{array}} \right),\theta _1  = \frac{{\pi _1 }}{{I_1 }}\Delta t.
1427 > \]
1428 > To reduce the cost of computing expensive functions in $e^{\Delta
1429 > tR_1 }$, we can use Cayley transformation,
1430 > \[
1431 > e^{\Delta tR_1 }  \approx (1 - \Delta tR_1 )^{ - 1} (1 + \Delta tR_1
1432 > )
1433 > \]
1434 > The flow maps for $T_2^r$ and $T_3^r$ can be found in the same
1435 > manner.
1436  
1437 + In order to construct a second-order symplectic method, we split the
1438 + angular kinetic Hamiltonian function can into five terms
1439   \[
1440 < R(t) = \sum\limits_{\alpha  = 1}^N {\left( {g_\alpha  x_\alpha  (0)
1441 < - \frac{{g_\alpha ^2 }}{{m_\alpha  \omega _\alpha ^2 }}x(0)}
1442 < \right)\cos (\omega _\alpha  t)}  + \frac{{\dot x_\alpha
1443 < (0)}}{{\omega _\alpha  }}\sin (\omega _\alpha  t)
1440 > T^r (\pi ) = \frac{1}{2}T_1 ^r (\pi _1 ) + \frac{1}{2}T_2^r (\pi _2
1441 > ) + T_3^r (\pi _3 ) + \frac{1}{2}T_2^r (\pi _2 ) + \frac{1}{2}T_1 ^r
1442 > (\pi _1 )
1443 > \].
1444 > Concatenating flows corresponding to these five terms, we can obtain
1445 > an symplectic integrator,
1446 > \[
1447 > \varphi _{\Delta t,T^r }  = \varphi _{\Delta t/2,\pi _1 }  \circ
1448 > \varphi _{\Delta t/2,\pi _2 }  \circ \varphi _{\Delta t,\pi _3 }
1449 > \circ \varphi _{\Delta t/2,\pi _2 }  \circ \varphi _{\Delta t/2,\pi
1450 > _1 }.
1451   \]
844 The random forces depend only on initial conditions.
1452  
1453 < \subsubsection{\label{introSection:secondFluctuationDissipation}The Second Fluctuation Dissipation Theorem}
1454 < So we can define a new set of coordinates,
1453 > The non-canonical Lie-Poisson bracket ${F, G}$ of two function
1454 > $F(\pi )$ and $G(\pi )$ is defined by
1455   \[
1456 < q_\alpha  (t) = x_\alpha  (t) - \frac{1}{{m_\alpha  \omega _\alpha
1457 < ^2 }}x(0)
1456 > \{ F,G\} (\pi ) = [\nabla _\pi  F(\pi )]^T J(\pi )\nabla _\pi  G(\pi
1457 > )
1458   \]
1459 < This makes
1459 > If the Poisson bracket of a function $F$ with an arbitrary smooth
1460 > function $G$ is zero, $F$ is a \emph{Casimir}, which is the
1461 > conserved quantity in Poisson system. We can easily verify that the
1462 > norm of the angular momentum, $\parallel \pi
1463 > \parallel$, is a \emph{Casimir}. Let$ F(\pi ) = S(\frac{{\parallel
1464 > \pi \parallel ^2 }}{2})$ for an arbitrary function $ S:R \to R$ ,
1465 > then by the chain rule
1466   \[
1467 < R(t) = \sum\limits_{\alpha  = 1}^N {g_\alpha  q_\alpha  (t)}
1467 > \nabla _\pi  F(\pi ) = S'(\frac{{\parallel \pi \parallel ^2
1468 > }}{2})\pi
1469   \]
1470 < And since the $q$ coordinates are harmonic oscillators,
1470 > Thus $ [\nabla _\pi  F(\pi )]^T J(\pi ) =  - S'(\frac{{\parallel \pi
1471 > \parallel ^2 }}{2})\pi  \times \pi  = 0 $. This explicit
1472 > Lie-Poisson integrator is found to be extremely efficient and stable
1473 > which can be explained by the fact the small angle approximation is
1474 > used and the norm of the angular momentum is conserved.
1475 >
1476 > \subsection{\label{introSection:RBHamiltonianSplitting} Hamiltonian
1477 > Splitting for Rigid Body}
1478 >
1479 > The Hamiltonian of rigid body can be separated in terms of kinetic
1480 > energy and potential energy,
1481   \[
1482 < \begin{array}{l}
859 < \left\langle {q_\alpha  (t)q_\alpha  (0)} \right\rangle  = \left\langle {q_\alpha ^2 (0)} \right\rangle \cos (\omega _\alpha  t) \\
860 < \left\langle {q_\alpha  (t)q_\beta  (0)} \right\rangle  = \delta _{\alpha \beta } \left\langle {q_\alpha  (t)q_\alpha  (0)} \right\rangle  \\
861 < \end{array}
1482 > H = T(p,\pi ) + V(q,Q)
1483   \]
1484 + The equations of motion corresponding to potential energy and
1485 + kinetic energy are listed in the below table,
1486 + \begin{table}
1487 + \caption{Equations of motion due to Potential and Kinetic Energies}
1488 + \begin{center}
1489 + \begin{tabular}{|l|l|}
1490 +  \hline
1491 +  % after \\: \hline or \cline{col1-col2} \cline{col3-col4} ...
1492 +  Potential & Kinetic \\
1493 +  $\frac{{dq}}{{dt}} = \frac{p}{m}$ & $\frac{d}{{dt}}q = p$ \\
1494 +  $\frac{d}{{dt}}p =  - \frac{{\partial V}}{{\partial q}}$ & $ \frac{d}{{dt}}p = 0$ \\
1495 +  $\frac{d}{{dt}}Q = 0$ & $ \frac{d}{{dt}}Q = Qskew(I^{ - 1} j)$ \\
1496 +  $ \frac{d}{{dt}}\pi  = \sum\limits_i {\left( {Q^T F_i (r,Q)} \right) \times X_i }$ & $\frac{d}{{dt}}\pi  = \pi  \times I^{ - 1} \pi$\\
1497 +  \hline
1498 + \end{tabular}
1499 + \end{center}
1500 + \end{table}
1501 + A second-order symplectic method is now obtained by the
1502 + composition of the flow maps,
1503 + \[
1504 + \varphi _{\Delta t}  = \varphi _{\Delta t/2,V}  \circ \varphi
1505 + _{\Delta t,T}  \circ \varphi _{\Delta t/2,V}.
1506 + \]
1507 + Moreover, $\varphi _{\Delta t/2,V}$ can be divided into two
1508 + sub-flows which corresponding to force and torque respectively,
1509 + \[
1510 + \varphi _{\Delta t/2,V}  = \varphi _{\Delta t/2,F}  \circ \varphi
1511 + _{\Delta t/2,\tau }.
1512 + \]
1513 + Since the associated operators of $\varphi _{\Delta t/2,F} $ and
1514 + $\circ \varphi _{\Delta t/2,\tau }$ are commuted, the composition
1515 + order inside $\varphi _{\Delta t/2,V}$ does not matter.
1516  
1517 < \begin{align}
1518 < \left\langle {R(t)R(0)} \right\rangle  &= \sum\limits_\alpha
1519 < {\sum\limits_\beta  {g_\alpha  g_\beta  \left\langle {q_\alpha
1520 < (t)q_\beta  (0)} \right\rangle } }
1521 < %
1522 < &= \sum\limits_\alpha  {g_\alpha ^2 \left\langle {q_\alpha ^2 (0)}
1523 < \right\rangle \cos (\omega _\alpha  t)}
1524 < %
1525 < &= kT\xi (t)
1526 < \end{align}
1517 > Furthermore, kinetic potential can be separated to translational
1518 > kinetic term, $T^t (p)$, and rotational kinetic term, $T^r (\pi )$,
1519 > \begin{equation}
1520 > T(p,\pi ) =T^t (p) + T^r (\pi ).
1521 > \end{equation}
1522 > where $ T^t (p) = \frac{1}{2}p^T m^{ - 1} p $ and $T^r (\pi )$ is
1523 > defined by \ref{introEquation:rotationalKineticRB}. Therefore, the
1524 > corresponding flow maps are given by
1525 > \[
1526 > \varphi _{\Delta t,T}  = \varphi _{\Delta t,T^t }  \circ \varphi
1527 > _{\Delta t,T^r }.
1528 > \]
1529 > Finally, we obtain the overall symplectic flow maps for free moving
1530 > rigid body
1531 > \begin{equation}
1532 > \begin{array}{c}
1533 > \varphi _{\Delta t}  = \varphi _{\Delta t/2,F}  \circ \varphi _{\Delta t/2,\tau }  \\
1534 >  \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 }  \\
1535 >  \circ \varphi _{\Delta t/2,\tau }  \circ \varphi _{\Delta t/2,F}  .\\
1536 > \end{array}
1537 > \label{introEquation:overallRBFlowMaps}
1538 > \end{equation}
1539  
1540 + \section{\label{introSection:langevinDynamics}Langevin Dynamics}
1541 + As an alternative to newtonian dynamics, Langevin dynamics, which
1542 + mimics a simple heat bath with stochastic and dissipative forces,
1543 + has been applied in a variety of studies. This section will review
1544 + the theory of Langevin dynamics simulation. A brief derivation of
1545 + generalized Langevin equation will be given first. Follow that, we
1546 + will discuss the physical meaning of the terms appearing in the
1547 + equation as well as the calculation of friction tensor from
1548 + hydrodynamics theory.
1549 +
1550 + \subsection{\label{introSection:generalizedLangevinDynamics}Derivation of Generalized Langevin Equation}
1551 +
1552 + Harmonic bath model, in which an effective set of harmonic
1553 + oscillators are used to mimic the effect of a linearly responding
1554 + environment, has been widely used in quantum chemistry and
1555 + statistical mechanics. One of the successful applications of
1556 + Harmonic bath model is the derivation of Deriving Generalized
1557 + Langevin Dynamics. Lets consider a system, in which the degree of
1558 + freedom $x$ is assumed to couple to the bath linearly, giving a
1559 + Hamiltonian of the form
1560   \begin{equation}
1561 < \xi (t) = \left\langle {R(t)R(0)} \right\rangle
1562 < \label{introEquation:secondFluctuationDissipation}
1561 > H = \frac{{p^2 }}{{2m}} + U(x) + H_B  + \Delta U(x,x_1 , \ldots x_N)
1562 > \label{introEquation:bathGLE}.
1563   \end{equation}
1564 + Here $p$ is a momentum conjugate to $q$, $m$ is the mass associated
1565 + with this degree of freedom, $H_B$ is harmonic bath Hamiltonian,
1566 + \[
1567 + H_B  = \sum\limits_{\alpha  = 1}^N {\left\{ {\frac{{p_\alpha ^2
1568 + }}{{2m_\alpha  }} + \frac{1}{2}m_\alpha  \omega _\alpha ^2 }
1569 + \right\}}
1570 + \]
1571 + where the index $\alpha$ runs over all the bath degrees of freedom,
1572 + $\omega _\alpha$ are the harmonic bath frequencies, $m_\alpha$ are
1573 + the harmonic bath masses, and $\Delta U$ is bilinear system-bath
1574 + coupling,
1575 + \[
1576 + \Delta U =  - \sum\limits_{\alpha  = 1}^N {g_\alpha  x_\alpha  x}
1577 + \]
1578 + where $g_\alpha$ are the coupling constants between the bath and the
1579 + coordinate $x$. Introducing
1580 + \[
1581 + W(x) = U(x) - \sum\limits_{\alpha  = 1}^N {\frac{{g_\alpha ^2
1582 + }}{{2m_\alpha  w_\alpha ^2 }}} x^2
1583 + \] and combining the last two terms in Equation
1584 + \ref{introEquation:bathGLE}, we may rewrite the Harmonic bath
1585 + Hamiltonian as
1586 + \[
1587 + H = \frac{{p^2 }}{{2m}} + W(x) + \sum\limits_{\alpha  = 1}^N
1588 + {\left\{ {\frac{{p_\alpha ^2 }}{{2m_\alpha  }} + \frac{1}{2}m_\alpha
1589 + w_\alpha ^2 \left( {x_\alpha   - \frac{{g_\alpha  }}{{m_\alpha
1590 + w_\alpha ^2 }}x} \right)^2 } \right\}}
1591 + \]
1592 + Since the first two terms of the new Hamiltonian depend only on the
1593 + system coordinates, we can get the equations of motion for
1594 + Generalized Langevin Dynamics by Hamilton's equations
1595 + \ref{introEquation:motionHamiltonianCoordinate,
1596 + introEquation:motionHamiltonianMomentum},
1597 + \begin{equation}
1598 + m\ddot x =  - \frac{{\partial W(x)}}{{\partial x}} -
1599 + \sum\limits_{\alpha  = 1}^N {g_\alpha  \left( {x_\alpha   -
1600 + \frac{{g_\alpha  }}{{m_\alpha  w_\alpha ^2 }}x} \right)},
1601 + \label{introEquation:coorMotionGLE}
1602 + \end{equation}
1603 + and
1604 + \begin{equation}
1605 + m\ddot x_\alpha   =  - m_\alpha  w_\alpha ^2 \left( {x_\alpha   -
1606 + \frac{{g_\alpha  }}{{m_\alpha  w_\alpha ^2 }}x} \right).
1607 + \label{introEquation:bathMotionGLE}
1608 + \end{equation}
1609  
1610 < \section{\label{introSection:hydroynamics}Hydrodynamics}
1610 > In order to derive an equation for $x$, the dynamics of the bath
1611 > variables $x_\alpha$ must be solved exactly first. As an integral
1612 > transform which is particularly useful in solving linear ordinary
1613 > differential equations, Laplace transform is the appropriate tool to
1614 > solve this problem. The basic idea is to transform the difficult
1615 > differential equations into simple algebra problems which can be
1616 > solved easily. Then applying inverse Laplace transform, also known
1617 > as the Bromwich integral, we can retrieve the solutions of the
1618 > original problems.
1619  
1620 < \subsection{\label{introSection:frictionTensor} Friction Tensor}
1621 < \subsection{\label{introSection:analyticalApproach}Analytical
1622 < Approach}
1620 > Let $f(t)$ be a function defined on $ [0,\infty ) $. The Laplace
1621 > transform of f(t) is a new function defined as
1622 > \[
1623 > L(f(t)) \equiv F(p) = \int_0^\infty  {f(t)e^{ - pt} dt}
1624 > \]
1625 > where  $p$ is real and  $L$ is called the Laplace Transform
1626 > Operator. Below are some important properties of Laplace transform
1627  
1628 < \subsection{\label{introSection:approximationApproach}Approximation
1629 < Approach}
1628 > \begin{eqnarray*}
1629 > L(x + y)  & = & L(x) + L(y) \\
1630 > L(ax)     & = & aL(x) \\
1631 > L(\dot x) & = & pL(x) - px(0) \\
1632 > L(\ddot x)& = & p^2 L(x) - px(0) - \dot x(0) \\
1633 > L\left( {\int_0^t {g(t - \tau )h(\tau )d\tau } } \right)& = & G(p)H(p) \\
1634 > \end{eqnarray*}
1635  
1636 < \subsection{\label{introSection:centersRigidBody}Centers of Rigid
1637 < Body}
1636 >
1637 > Applying Laplace transform to the bath coordinates, we obtain
1638 > \begin{eqnarray*}
1639 > 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) \\
1640 > L(x_\alpha  ) & = & \frac{{\frac{{g_\alpha  }}{{\omega _\alpha  }}L(x) + px_\alpha  (0) + \dot x_\alpha  (0)}}{{p^2  + \omega _\alpha ^2 }} \\
1641 > \end{eqnarray*}
1642 >
1643 > By the same way, the system coordinates become
1644 > \begin{eqnarray*}
1645 > mL(\ddot x) & = & - \frac{1}{p}\frac{{\partial W(x)}}{{\partial x}} \\
1646 >  & & \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\}}  \\
1647 > \end{eqnarray*}
1648 >
1649 > With the help of some relatively important inverse Laplace
1650 > transformations:
1651 > \[
1652 > \begin{array}{c}
1653 > L(\cos at) = \frac{p}{{p^2  + a^2 }} \\
1654 > L(\sin at) = \frac{a}{{p^2  + a^2 }} \\
1655 > L(1) = \frac{1}{p} \\
1656 > \end{array}
1657 > \]
1658 > , we obtain
1659 > \begin{eqnarray*}
1660 > m\ddot x & = & - \frac{{\partial W(x)}}{{\partial x}} -
1661 > \sum\limits_{\alpha  = 1}^N {\left\{ {\left( { - \frac{{g_\alpha ^2
1662 > }}{{m_\alpha  \omega _\alpha ^2 }}} \right)\int_0^t {\cos (\omega
1663 > _\alpha  t)\dot x(t - \tau )d\tau  \\
1664 > & &\mbox{} - \left[ {g_\alpha  x_\alpha (0) - \frac{{g_\alpha
1665 > }}{{m_\alpha \omega _\alpha  }}} \right]\cos (\omega _\alpha  t) -
1666 > \frac{{g_\alpha  \dot x_\alpha  (0)}}{{\omega
1667 > _\alpha  }}\sin (\omega _\alpha  t)} } \right\}} \\
1668 > %
1669 > & = & - \frac{{\partial W(x)}}{{\partial x}} - \int_0^t
1670 > {\sum\limits_{\alpha  = 1}^N {\left( { - \frac{{g_\alpha ^2
1671 > }}{{m_\alpha  \omega _\alpha ^2 }}} \right)\cos (\omega _\alpha
1672 > t)\dot x(t - \tau )d} \tau }  + \sum\limits_{\alpha  = 1}^N {\left\{
1673 > {\left[ {g_\alpha  x_\alpha  (0) \\
1674 > & & \mbox{} - \frac{{g_\alpha  }}{{m_\alpha \omega _\alpha  }}}
1675 > \right]\cos (\omega _\alpha  t) + \frac{{g_\alpha  \dot x_\alpha
1676 > (0)}}{{\omega _\alpha  }}\sin (\omega _\alpha  t)} \right\}}
1677 > \end{eqnarray*}
1678 > Introducing a \emph{dynamic friction kernel}
1679 > \begin{equation}
1680 > \xi (t) = \sum\limits_{\alpha  = 1}^N {\left( { - \frac{{g_\alpha ^2
1681 > }}{{m_\alpha  \omega _\alpha ^2 }}} \right)\cos (\omega _\alpha  t)}
1682 > \label{introEquation:dynamicFrictionKernelDefinition}
1683 > \end{equation}
1684 > and \emph{a random force}
1685 > \begin{equation}
1686 > R(t) = \sum\limits_{\alpha  = 1}^N {\left( {g_\alpha  x_\alpha  (0)
1687 > - \frac{{g_\alpha ^2 }}{{m_\alpha  \omega _\alpha ^2 }}x(0)}
1688 > \right)\cos (\omega _\alpha  t)}  + \frac{{\dot x_\alpha
1689 > (0)}}{{\omega _\alpha  }}\sin (\omega _\alpha  t),
1690 > \label{introEquation:randomForceDefinition}
1691 > \end{equation}
1692 > the equation of motion can be rewritten as
1693 > \begin{equation}
1694 > m\ddot x =  - \frac{{\partial W}}{{\partial x}} - \int_0^t {\xi
1695 > (t)\dot x(t - \tau )d\tau }  + R(t)
1696 > \label{introEuqation:GeneralizedLangevinDynamics}
1697 > \end{equation}
1698 > which is known as the \emph{generalized Langevin equation}.
1699 >
1700 > \subsubsection{\label{introSection:randomForceDynamicFrictionKernel}Random Force and Dynamic Friction Kernel}
1701 >
1702 > One may notice that $R(t)$ depends only on initial conditions, which
1703 > implies it is completely deterministic within the context of a
1704 > harmonic bath. However, it is easy to verify that $R(t)$ is totally
1705 > uncorrelated to $x$ and $\dot x$,
1706 > \[
1707 > \begin{array}{l}
1708 > \left\langle {x(t)R(t)} \right\rangle  = 0, \\
1709 > \left\langle {\dot x(t)R(t)} \right\rangle  = 0. \\
1710 > \end{array}
1711 > \]
1712 > This property is what we expect from a truly random process. As long
1713 > as the model, which is gaussian distribution in general, chosen for
1714 > $R(t)$ is a truly random process, the stochastic nature of the GLE
1715 > still remains.
1716 >
1717 > %dynamic friction kernel
1718 > The convolution integral
1719 > \[
1720 > \int_0^t {\xi (t)\dot x(t - \tau )d\tau }
1721 > \]
1722 > depends on the entire history of the evolution of $x$, which implies
1723 > that the bath retains memory of previous motions. In other words,
1724 > the bath requires a finite time to respond to change in the motion
1725 > of the system. For a sluggish bath which responds slowly to changes
1726 > in the system coordinate, we may regard $\xi(t)$ as a constant
1727 > $\xi(t) = \Xi_0$. Hence, the convolution integral becomes
1728 > \[
1729 > \int_0^t {\xi (t)\dot x(t - \tau )d\tau }  = \xi _0 (x(t) - x(0))
1730 > \]
1731 > and Equation \ref{introEuqation:GeneralizedLangevinDynamics} becomes
1732 > \[
1733 > m\ddot x =  - \frac{\partial }{{\partial x}}\left( {W(x) +
1734 > \frac{1}{2}\xi _0 (x - x_0 )^2 } \right) + R(t),
1735 > \]
1736 > which can be used to describe dynamic caging effect. The other
1737 > extreme is the bath that responds infinitely quickly to motions in
1738 > the system. Thus, $\xi (t)$ can be taken as a $delta$ function in
1739 > time:
1740 > \[
1741 > \xi (t) = 2\xi _0 \delta (t)
1742 > \]
1743 > Hence, the convolution integral becomes
1744 > \[
1745 > \int_0^t {\xi (t)\dot x(t - \tau )d\tau }  = 2\xi _0 \int_0^t
1746 > {\delta (t)\dot x(t - \tau )d\tau }  = \xi _0 \dot x(t),
1747 > \]
1748 > and Equation \ref{introEuqation:GeneralizedLangevinDynamics} becomes
1749 > \begin{equation}
1750 > m\ddot x =  - \frac{{\partial W(x)}}{{\partial x}} - \xi _0 \dot
1751 > x(t) + R(t) \label{introEquation:LangevinEquation}
1752 > \end{equation}
1753 > which is known as the Langevin equation. The static friction
1754 > coefficient $\xi _0$ can either be calculated from spectral density
1755 > or be determined by Stokes' law for regular shaped particles.A
1756 > briefly review on calculating friction tensor for arbitrary shaped
1757 > particles is given in Sec.~\ref{introSection:frictionTensor}.
1758 >
1759 > \subsubsection{\label{introSection:secondFluctuationDissipation}The Second Fluctuation Dissipation Theorem}
1760 >
1761 > Defining a new set of coordinates,
1762 > \[
1763 > q_\alpha  (t) = x_\alpha  (t) - \frac{1}{{m_\alpha  \omega _\alpha
1764 > ^2 }}x(0)
1765 > \],
1766 > we can rewrite $R(T)$ as
1767 > \[
1768 > R(t) = \sum\limits_{\alpha  = 1}^N {g_\alpha  q_\alpha  (t)}.
1769 > \]
1770 > And since the $q$ coordinates are harmonic oscillators,
1771 >
1772 > \begin{eqnarray*}
1773 > \left\langle {q_\alpha ^2 } \right\rangle  & = & \frac{{kT}}{{m_\alpha  \omega _\alpha ^2 }} \\
1774 > \left\langle {q_\alpha  (t)q_\alpha  (0)} \right\rangle & = & \left\langle {q_\alpha ^2 (0)} \right\rangle \cos (\omega _\alpha  t) \\
1775 > \left\langle {q_\alpha  (t)q_\beta  (0)} \right\rangle & = &\delta _{\alpha \beta } \left\langle {q_\alpha  (t)q_\alpha  (0)} \right\rangle  \\
1776 > \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 } }  \\
1777 >  & = &\sum\limits_\alpha  {g_\alpha ^2 \left\langle {q_\alpha ^2 (0)} \right\rangle \cos (\omega _\alpha  t)}  \\
1778 >  & = &kT\xi (t) \\
1779 > \end{eqnarray*}
1780 >
1781 > Thus, we recover the \emph{second fluctuation dissipation theorem}
1782 > \begin{equation}
1783 > \xi (t) = \left\langle {R(t)R(0)} \right\rangle
1784 > \label{introEquation:secondFluctuationDissipation}.
1785 > \end{equation}
1786 > In effect, it acts as a constraint on the possible ways in which one
1787 > can model the random force and friction kernel.
1788 >
1789 > \subsection{\label{introSection:frictionTensor} Friction Tensor}
1790 > Theoretically, the friction kernel can be determined using velocity
1791 > autocorrelation function. However, this approach become impractical
1792 > when the system become more and more complicate. Instead, various
1793 > approaches based on hydrodynamics have been developed to calculate
1794 > the friction coefficients. The friction effect is isotropic in
1795 > Equation, $\zeta$ can be taken as a scalar. In general, friction
1796 > tensor $\Xi$ is a $6\times 6$ matrix given by
1797 > \[
1798 > \Xi  = \left( {\begin{array}{*{20}c}
1799 >   {\Xi _{}^{tt} } & {\Xi _{}^{rt} }  \\
1800 >   {\Xi _{}^{tr} } & {\Xi _{}^{rr} }  \\
1801 > \end{array}} \right).
1802 > \]
1803 > Here, $ {\Xi^{tt} }$ and $ {\Xi^{rr} }$ are translational friction
1804 > tensor and rotational resistance (friction) tensor respectively,
1805 > while ${\Xi^{tr} }$ is translation-rotation coupling tensor and $
1806 > {\Xi^{rt} }$ is rotation-translation coupling tensor. When a
1807 > particle moves in a fluid, it may experience friction force or
1808 > torque along the opposite direction of the velocity or angular
1809 > velocity,
1810 > \[
1811 > \left( \begin{array}{l}
1812 > F_R  \\
1813 > \tau _R  \\
1814 > \end{array} \right) =  - \left( {\begin{array}{*{20}c}
1815 >   {\Xi ^{tt} } & {\Xi ^{rt} }  \\
1816 >   {\Xi ^{tr} } & {\Xi ^{rr} }  \\
1817 > \end{array}} \right)\left( \begin{array}{l}
1818 > v \\
1819 > w \\
1820 > \end{array} \right)
1821 > \]
1822 > where $F_r$ is the friction force and $\tau _R$ is the friction
1823 > toque.
1824 >
1825 > \subsubsection{\label{introSection:resistanceTensorRegular}The Resistance Tensor for Regular Shape}
1826 >
1827 > For a spherical particle, the translational and rotational friction
1828 > constant can be calculated from Stoke's law,
1829 > \[
1830 > \Xi ^{tt}  = \left( {\begin{array}{*{20}c}
1831 >   {6\pi \eta R} & 0 & 0  \\
1832 >   0 & {6\pi \eta R} & 0  \\
1833 >   0 & 0 & {6\pi \eta R}  \\
1834 > \end{array}} \right)
1835 > \]
1836 > and
1837 > \[
1838 > \Xi ^{rr}  = \left( {\begin{array}{*{20}c}
1839 >   {8\pi \eta R^3 } & 0 & 0  \\
1840 >   0 & {8\pi \eta R^3 } & 0  \\
1841 >   0 & 0 & {8\pi \eta R^3 }  \\
1842 > \end{array}} \right)
1843 > \]
1844 > where $\eta$ is the viscosity of the solvent and $R$ is the
1845 > hydrodynamics radius.
1846 >
1847 > Other non-spherical shape, such as cylinder and ellipsoid
1848 > \textit{etc}, are widely used as reference for developing new
1849 > hydrodynamics theory, because their properties can be calculated
1850 > exactly. In 1936, Perrin extended Stokes's law to general ellipsoid,
1851 > also called a triaxial ellipsoid, which is given in Cartesian
1852 > coordinates by\cite{Perrin1934, Perrin1936}
1853 > \[
1854 > \frac{{x^2 }}{{a^2 }} + \frac{{y^2 }}{{b^2 }} + \frac{{z^2 }}{{c^2
1855 > }} = 1
1856 > \]
1857 > where the semi-axes are of lengths $a$, $b$, and $c$. Unfortunately,
1858 > due to the complexity of the elliptic integral, only the ellipsoid
1859 > with the restriction of two axes having to be equal, \textit{i.e.}
1860 > prolate($ a \ge b = c$) and oblate ($ a < b = c $), can be solved
1861 > exactly. Introducing an elliptic integral parameter $S$ for prolate,
1862 > \[
1863 > S = \frac{2}{{\sqrt {a^2  - b^2 } }}\ln \frac{{a + \sqrt {a^2  - b^2
1864 > } }}{b},
1865 > \]
1866 > and oblate,
1867 > \[
1868 > S = \frac{2}{{\sqrt {b^2  - a^2 } }}arctg\frac{{\sqrt {b^2  - a^2 }
1869 > }}{a}
1870 > \],
1871 > one can write down the translational and rotational resistance
1872 > tensors
1873 > \[
1874 > \begin{array}{l}
1875 > \Xi _a^{tt}  = 16\pi \eta \frac{{a^2  - b^2 }}{{(2a^2  - b^2 )S - 2a}} \\
1876 > \Xi _b^{tt}  = \Xi _c^{tt}  = 32\pi \eta \frac{{a^2  - b^2 }}{{(2a^2  - 3b^2 )S + 2a}} \\
1877 > \end{array},
1878 > \]
1879 > and
1880 > \[
1881 > \begin{array}{l}
1882 > \Xi _a^{rr}  = \frac{{32\pi }}{3}\eta \frac{{(a^2  - b^2 )b^2 }}{{2a - b^2 S}} \\
1883 > \Xi _b^{rr}  = \Xi _c^{rr}  = \frac{{32\pi }}{3}\eta \frac{{(a^4  - b^4 )}}{{(2a^2  - b^2 )S - 2a}} \\
1884 > \end{array}.
1885 > \]
1886 >
1887 > \subsubsection{\label{introSection:resistanceTensorRegularArbitrary}The Resistance Tensor for Arbitrary Shape}
1888 >
1889 > Unlike spherical and other regular shaped molecules, there is not
1890 > analytical solution for friction tensor of any arbitrary shaped
1891 > rigid molecules. The ellipsoid of revolution model and general
1892 > triaxial ellipsoid model have been used to approximate the
1893 > hydrodynamic properties of rigid bodies. However, since the mapping
1894 > from all possible ellipsoidal space, $r$-space, to all possible
1895 > combination of rotational diffusion coefficients, $D$-space is not
1896 > unique\cite{Wegener1979} as well as the intrinsic coupling between
1897 > translational and rotational motion of rigid body, general ellipsoid
1898 > is not always suitable for modeling arbitrarily shaped rigid
1899 > molecule. A number of studies have been devoted to determine the
1900 > friction tensor for irregularly shaped rigid bodies using more
1901 > advanced method where the molecule of interest was modeled by
1902 > combinations of spheres(beads)\cite{Carrasco1999} and the
1903 > hydrodynamics properties of the molecule can be calculated using the
1904 > hydrodynamic interaction tensor. Let us consider a rigid assembly of
1905 > $N$ beads immersed in a continuous medium. Due to hydrodynamics
1906 > interaction, the ``net'' velocity of $i$th bead, $v'_i$ is different
1907 > than its unperturbed velocity $v_i$,
1908 > \[
1909 > v'_i  = v_i  - \sum\limits_{j \ne i} {T_{ij} F_j }
1910 > \]
1911 > where $F_i$ is the frictional force, and $T_{ij}$ is the
1912 > hydrodynamic interaction tensor. The friction force of $i$th bead is
1913 > proportional to its ``net'' velocity
1914 > \begin{equation}
1915 > F_i  = \zeta _i v_i  - \zeta _i \sum\limits_{j \ne i} {T_{ij} F_j }.
1916 > \label{introEquation:tensorExpression}
1917 > \end{equation}
1918 > This equation is the basis for deriving the hydrodynamic tensor. In
1919 > 1930, Oseen and Burgers gave a simple solution to Equation
1920 > \ref{introEquation:tensorExpression}
1921 > \begin{equation}
1922 > T_{ij}  = \frac{1}{{8\pi \eta r_{ij} }}\left( {I + \frac{{R_{ij}
1923 > R_{ij}^T }}{{R_{ij}^2 }}} \right).
1924 > \label{introEquation:oseenTensor}
1925 > \end{equation}
1926 > Here $R_{ij}$ is the distance vector between bead $i$ and bead $j$.
1927 > A second order expression for element of different size was
1928 > introduced by Rotne and Prager\cite{Rotne1969} and improved by
1929 > Garc\'{i}a de la Torre and Bloomfield\cite{Torre1977},
1930 > \begin{equation}
1931 > T_{ij}  = \frac{1}{{8\pi \eta R_{ij} }}\left[ {\left( {I +
1932 > \frac{{R_{ij} R_{ij}^T }}{{R_{ij}^2 }}} \right) + R\frac{{\sigma
1933 > _i^2  + \sigma _j^2 }}{{r_{ij}^2 }}\left( {\frac{I}{3} -
1934 > \frac{{R_{ij} R_{ij}^T }}{{R_{ij}^2 }}} \right)} \right].
1935 > \label{introEquation:RPTensorNonOverlapped}
1936 > \end{equation}
1937 > Both of the Equation \ref{introEquation:oseenTensor} and Equation
1938 > \ref{introEquation:RPTensorNonOverlapped} have an assumption $R_{ij}
1939 > \ge \sigma _i  + \sigma _j$. An alternative expression for
1940 > overlapping beads with the same radius, $\sigma$, is given by
1941 > \begin{equation}
1942 > T_{ij}  = \frac{1}{{6\pi \eta R_{ij} }}\left[ {\left( {1 -
1943 > \frac{2}{{32}}\frac{{R_{ij} }}{\sigma }} \right)I +
1944 > \frac{2}{{32}}\frac{{R_{ij} R_{ij}^T }}{{R_{ij} \sigma }}} \right]
1945 > \label{introEquation:RPTensorOverlapped}
1946 > \end{equation}
1947 >
1948 > To calculate the resistance tensor at an arbitrary origin $O$, we
1949 > construct a $3N \times 3N$ matrix consisting of $N \times N$
1950 > $B_{ij}$ blocks
1951 > \begin{equation}
1952 > B = \left( {\begin{array}{*{20}c}
1953 >   {B_{11} } &  \ldots  & {B_{1N} }  \\
1954 >    \vdots  &  \ddots  &  \vdots   \\
1955 >   {B_{N1} } &  \cdots  & {B_{NN} }  \\
1956 > \end{array}} \right),
1957 > \end{equation}
1958 > where $B_{ij}$ is given by
1959 > \[
1960 > B_{ij}  = \delta _{ij} \frac{I}{{6\pi \eta R}} + (1 - \delta _{ij}
1961 > )T_{ij}
1962 > \]
1963 > where $\delta _{ij}$ is Kronecker delta function. Inverting matrix
1964 > $B$, we obtain
1965 >
1966 > \[
1967 > C = B^{ - 1}  = \left( {\begin{array}{*{20}c}
1968 >   {C_{11} } &  \ldots  & {C_{1N} }  \\
1969 >    \vdots  &  \ddots  &  \vdots   \\
1970 >   {C_{N1} } &  \cdots  & {C_{NN} }  \\
1971 > \end{array}} \right)
1972 > \]
1973 > , which can be partitioned into $N \times N$ $3 \times 3$ block
1974 > $C_{ij}$. With the help of $C_{ij}$ and skew matrix $U_i$
1975 > \[
1976 > U_i  = \left( {\begin{array}{*{20}c}
1977 >   0 & { - z_i } & {y_i }  \\
1978 >   {z_i } & 0 & { - x_i }  \\
1979 >   { - y_i } & {x_i } & 0  \\
1980 > \end{array}} \right)
1981 > \]
1982 > where $x_i$, $y_i$, $z_i$ are the components of the vector joining
1983 > bead $i$ and origin $O$. Hence, the elements of resistance tensor at
1984 > arbitrary origin $O$ can be written as
1985 > \begin{equation}
1986 > \begin{array}{l}
1987 > \Xi _{}^{tt}  = \sum\limits_i {\sum\limits_j {C_{ij} } } , \\
1988 > \Xi _{}^{tr}  = \Xi _{}^{rt}  = \sum\limits_i {\sum\limits_j {U_i C_{ij} } } , \\
1989 > \Xi _{}^{rr}  =  - \sum\limits_i {\sum\limits_j {U_i C_{ij} } } U_j  \\
1990 > \end{array}
1991 > \label{introEquation:ResistanceTensorArbitraryOrigin}
1992 > \end{equation}
1993 >
1994 > The resistance tensor depends on the origin to which they refer. The
1995 > proper location for applying friction force is the center of
1996 > resistance (reaction), at which the trace of rotational resistance
1997 > tensor, $ \Xi ^{rr}$ reaches minimum. Mathematically, the center of
1998 > resistance is defined as an unique point of the rigid body at which
1999 > the translation-rotation coupling tensor are symmetric,
2000 > \begin{equation}
2001 > \Xi^{tr}  = \left( {\Xi^{tr} } \right)^T
2002 > \label{introEquation:definitionCR}
2003 > \end{equation}
2004 > Form Equation \ref{introEquation:ResistanceTensorArbitraryOrigin},
2005 > we can easily find out that the translational resistance tensor is
2006 > origin independent, while the rotational resistance tensor and
2007 > translation-rotation coupling resistance tensor depend on the
2008 > origin. Given resistance tensor at an arbitrary origin $O$, and a
2009 > vector ,$r_{OP}(x_{OP}, y_{OP}, z_{OP})$, from $O$ to $P$, we can
2010 > obtain the resistance tensor at $P$ by
2011 > \begin{equation}
2012 > \begin{array}{l}
2013 > \Xi _P^{tt}  = \Xi _O^{tt}  \\
2014 > \Xi _P^{tr}  = \Xi _P^{rt}  = \Xi _O^{tr}  - U_{OP} \Xi _O^{tt}  \\
2015 > \Xi _P^{rr}  = \Xi _O^{rr}  - U_{OP} \Xi _O^{tt} U_{OP}  + \Xi _O^{tr} U_{OP}  - U_{OP} \Xi _O^{tr} ^{^T }  \\
2016 > \end{array}
2017 > \label{introEquation:resistanceTensorTransformation}
2018 > \end{equation}
2019 > where
2020 > \[
2021 > U_{OP}  = \left( {\begin{array}{*{20}c}
2022 >   0 & { - z_{OP} } & {y_{OP} }  \\
2023 >   {z_i } & 0 & { - x_{OP} }  \\
2024 >   { - y_{OP} } & {x_{OP} } & 0  \\
2025 > \end{array}} \right)
2026 > \]
2027 > Using Equations \ref{introEquation:definitionCR} and
2028 > \ref{introEquation:resistanceTensorTransformation}, one can locate
2029 > the position of center of resistance,
2030 > \begin{eqnarray*}
2031 > \left( \begin{array}{l}
2032 > x_{OR}  \\
2033 > y_{OR}  \\
2034 > z_{OR}  \\
2035 > \end{array} \right) & = &\left( {\begin{array}{*{20}c}
2036 >   {(\Xi _O^{rr} )_{yy}  + (\Xi _O^{rr} )_{zz} } & { - (\Xi _O^{rr} )_{xy} } & { - (\Xi _O^{rr} )_{xz} }  \\
2037 >   { - (\Xi _O^{rr} )_{xy} } & {(\Xi _O^{rr} )_{zz}  + (\Xi _O^{rr} )_{xx} } & { - (\Xi _O^{rr} )_{yz} }  \\
2038 >   { - (\Xi _O^{rr} )_{xz} } & { - (\Xi _O^{rr} )_{yz} } & {(\Xi _O^{rr} )_{xx}  + (\Xi _O^{rr} )_{yy} }  \\
2039 > \end{array}} \right)^{ - 1}  \\
2040 >  & & \left( \begin{array}{l}
2041 > (\Xi _O^{tr} )_{yz}  - (\Xi _O^{tr} )_{zy}  \\
2042 > (\Xi _O^{tr} )_{zx}  - (\Xi _O^{tr} )_{xz}  \\
2043 > (\Xi _O^{tr} )_{xy}  - (\Xi _O^{tr} )_{yx}  \\
2044 > \end{array} \right) \\
2045 > \end{eqnarray*}
2046 >
2047 >
2048 >
2049 > where $x_OR$, $y_OR$, $z_OR$ are the components of the vector
2050 > joining center of resistance $R$ and origin $O$.

Diff Legend

Removed lines
+ Added lines
< Changed lines
> Changed lines