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 2706 by tim, Thu Apr 13 04:47:47 2006 UTC vs.
Revision 2779 by tim, Fri May 26 18:25:41 2006 UTC

# Line 570 | 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 $x$.
573 The free rigid body is an example of Poisson system (actually a
574 Lie-Poisson system) with Hamiltonian function of angular kinetic
575 energy.
576 \begin{equation}
577 J(\pi ) = \left( {\begin{array}{*{20}c}
578   0 & {\pi _3 } & { - \pi _2 }  \\
579   { - \pi _3 } & 0 & {\pi _1 }  \\
580   {\pi _2 } & { - \pi _1 } & 0  \\
581 \end{array}} \right)
582 \end{equation}
583
584 \begin{equation}
585 H = \frac{1}{2}\left( {\frac{{\pi _1^2 }}{{I_1 }} + \frac{{\pi _2^2
586 }}{{I_2 }} + \frac{{\pi _3^2 }}{{I_3 }}} \right)
587 \end{equation}
573  
574   \subsection{\label{introSection:exactFlow}Exact Flow}
575  
# Line 837 | Line 822 | q(\Delta t)} \right]. %
822   %
823   q(\Delta t) &= q(0) + \frac{{\Delta t}}{2}\left[ {\dot q(0) + \dot
824   q(\Delta t)} \right]. %
825 < \label{introEquation:positionVerlet1}
825 > \label{introEquation:positionVerlet2}
826   \end{align}
827  
828   \subsubsection{\label{introSection:errorAnalysis}Error Analysis and Higher Order Methods}
# Line 846 | Line 831 | $\varphi_1(t)$ and $\varphi_2(t$ respectively , we hav
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
834 > $\varphi_1(t)$ and $\varphi_2(t)$ respectively , we have
835   \begin{equation}
836   \exp (hX + hY) = \exp (hZ)
837   \end{equation}
# Line 862 | Line 847 | can obtain
847   Applying Baker-Campbell-Hausdorff formula to Sprang splitting, we
848   can obtain
849   \begin{eqnarray*}
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{} +
868 < \ldots )
850 > \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 \\
851 >                                   &   & \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 + \ldots )
853   \end{eqnarray*}
854   Since \[ [X,Y] + [Y,X] = 0\] and \[ [X,X] = 0\], the dominant local
855   error of Spring splitting is proportional to $h^3$. The same
# Line 874 | Line 858 | Careful choice of coefficient $a_1 ,\ldot , b_m$ will
858   \varphi _{b_m h}^2  \circ \varphi _{a_m h}^1  \circ \varphi _{b_{m -
859   1} h}^2  \circ  \ldots  \circ \varphi _{a_1 h}^1 .
860   \end{equation}
861 < Careful choice of coefficient $a_1 ,\ldot , b_m$ will lead to higher
861 > Careful choice of coefficient $a_1 \ldot b_m$ will lead to higher
862   order method. Yoshida proposed an elegant way to compose higher
863   order methods based on symmetric splitting. Given a symmetric second
864   order base method $ \varphi _h^{(2)} $, a fourth-order symmetric
# Line 898 | Line 882 | As a special discipline of molecular modeling, Molecul
882  
883   \section{\label{introSection:molecularDynamics}Molecular Dynamics}
884  
885 < As a special discipline of molecular modeling, Molecular dynamics
886 < has proven to be a powerful tool for studying the functions of
887 < biological systems, providing structural, thermodynamic and
888 < dynamical information.
889 <
890 < \subsection{\label{introSec:mdInit}Initialization}
891 <
892 < \subsection{\label{introSec:forceEvaluation}Force Evaluation}
885 > As one of the principal tools of molecular modeling, Molecular
886 > dynamics has proven to be a powerful tool for studying the functions
887 > of biological systems, providing structural, thermodynamic and
888 > dynamical information. The basic idea of molecular dynamics is that
889 > macroscopic properties are related to microscopic behavior and
890 > microscopic behavior can be calculated from the trajectories in
891 > simulations. For instance, instantaneous temperature of an
892 > Hamiltonian system of $N$ particle can be measured by
893 > \[
894 > T = \sum\limits_{i = 1}^N {\frac{{m_i v_i^2 }}{{fk_B }}}
895 > \]
896 > where $m_i$ and $v_i$ are the mass and velocity of $i$th particle
897 > respectively, $f$ is the number of degrees of freedom, and $k_B$ is
898 > the boltzman constant.
899  
900 < \subsection{\label{introSection:mdIntegration} Integration of the Equations of Motion}
900 > A typical molecular dynamics run consists of three essential steps:
901 > \begin{enumerate}
902 >  \item Initialization
903 >    \begin{enumerate}
904 >    \item Preliminary preparation
905 >    \item Minimization
906 >    \item Heating
907 >    \item Equilibration
908 >    \end{enumerate}
909 >  \item Production
910 >  \item Analysis
911 > \end{enumerate}
912 > These three individual steps will be covered in the following
913 > sections. Sec.~\ref{introSec:initialSystemSettings} deals with the
914 > initialization of a simulation. Sec.~\ref{introSec:production} will
915 > discusses issues in production run. Sec.~\ref{introSection:Analysis}
916 > provides the theoretical tools for trajectory analysis.
917 >
918 > \subsection{\label{introSec:initialSystemSettings}Initialization}
919 >
920 > \subsubsection{Preliminary preparation}
921 >
922 > When selecting the starting structure of a molecule for molecular
923 > simulation, one may retrieve its Cartesian coordinates from public
924 > databases, such as RCSB Protein Data Bank \textit{etc}. Although
925 > thousands of crystal structures of molecules are discovered every
926 > year, many more remain unknown due to the difficulties of
927 > purification and crystallization. Even for the molecule with known
928 > structure, some important information is missing. For example, the
929 > missing hydrogen atom which acts as donor in hydrogen bonding must
930 > be added. Moreover, in order to include electrostatic interaction,
931 > one may need to specify the partial charges for individual atoms.
932 > Under some circumstances, we may even need to prepare the system in
933 > a special setup. For instance, when studying transport phenomenon in
934 > membrane system, we may prepare the lipids in bilayer structure
935 > instead of placing lipids randomly in solvent, since we are not
936 > interested in self-aggregation and it takes a long time to happen.
937 >
938 > \subsubsection{Minimization}
939 >
940 > It is quite possible that some of molecules in the system from
941 > preliminary preparation may be overlapped with each other. This
942 > close proximity leads to high potential energy which consequently
943 > jeopardizes any molecular dynamics simulations. To remove these
944 > steric overlaps, one typically performs energy minimization to find
945 > a more reasonable conformation. Several energy minimization methods
946 > have been developed to exploit the energy surface and to locate the
947 > local minimum. While converging slowly near the minimum, steepest
948 > descent method is extremely robust when systems are far from
949 > harmonic. Thus, it is often used to refine structure from
950 > crystallographic data. Relied on the gradient or hessian, advanced
951 > methods like conjugate gradient and Newton-Raphson converge rapidly
952 > to a local minimum, while become unstable if the energy surface is
953 > far from quadratic. Another factor must be taken into account, when
954 > choosing energy minimization method, is the size of the system.
955 > Steepest descent and conjugate gradient can deal with models of any
956 > size. Because of the limit of computation power to calculate hessian
957 > matrix and insufficient storage capacity to store them, most
958 > Newton-Raphson methods can not be used with very large models.
959 >
960 > \subsubsection{Heating}
961 >
962 > Typically, Heating is performed by assigning random velocities
963 > according to a Gaussian distribution for a temperature. Beginning at
964 > a lower temperature and gradually increasing the temperature by
965 > assigning greater random velocities, we end up with setting the
966 > temperature of the system to a final temperature at which the
967 > simulation will be conducted. In heating phase, we should also keep
968 > the system from drifting or rotating as a whole. Equivalently, the
969 > net linear momentum and angular momentum of the system should be
970 > shifted to zero.
971 >
972 > \subsubsection{Equilibration}
973 >
974 > The purpose of equilibration is to allow the system to evolve
975 > spontaneously for a period of time and reach equilibrium. The
976 > procedure is continued until various statistical properties, such as
977 > temperature, pressure, energy, volume and other structural
978 > properties \textit{etc}, become independent of time. Strictly
979 > speaking, minimization and heating are not necessary, provided the
980 > equilibration process is long enough. However, these steps can serve
981 > as a means to arrive at an equilibrated structure in an effective
982 > way.
983 >
984 > \subsection{\label{introSection:production}Production}
985 >
986 > Production run is the most important steps of the simulation, in
987 > which the equilibrated structure is used as a starting point and the
988 > motions of the molecules are collected for later analysis. In order
989 > to capture the macroscopic properties of the system, the molecular
990 > dynamics simulation must be performed in correct and efficient way.
991 >
992 > The most expensive part of a molecular dynamics simulation is the
993 > calculation of non-bonded forces, such as van der Waals force and
994 > Coulombic forces \textit{etc}. For a system of $N$ particles, the
995 > complexity of the algorithm for pair-wise interactions is $O(N^2 )$,
996 > which making large simulations prohibitive in the absence of any
997 > computation saving techniques.
998 >
999 > A natural approach to avoid system size issue is to represent the
1000 > bulk behavior by a finite number of the particles. However, this
1001 > approach will suffer from the surface effect. To offset this,
1002 > \textit{Periodic boundary condition} is developed to simulate bulk
1003 > properties with a relatively small number of particles. In this
1004 > method, the simulation box is replicated throughout space to form an
1005 > infinite lattice. During the simulation, when a particle moves in
1006 > the primary cell, its image in other cells move in exactly the same
1007 > direction with exactly the same orientation. Thus, as a particle
1008 > leaves the primary cell, one of its images will enter through the
1009 > opposite face.
1010 > %\begin{figure}
1011 > %\centering
1012 > %\includegraphics[width=\linewidth]{pbcFig.eps}
1013 > %\caption[An illustration of periodic boundary conditions]{A 2-D
1014 > %illustration of periodic boundary conditions. As one particle leaves
1015 > %the right of the simulation box, an image of it enters the left.}
1016 > %\label{introFig:pbc}
1017 > %\end{figure}
1018 >
1019 > %cutoff and minimum image convention
1020 > Another important technique to improve the efficiency of force
1021 > evaluation is to apply cutoff where particles farther than a
1022 > predetermined distance, are not included in the calculation
1023 > \cite{Frenkel1996}. The use of a cutoff radius will cause a
1024 > discontinuity in the potential energy curve. Fortunately, one can
1025 > shift the potential to ensure the potential curve go smoothly to
1026 > zero at the cutoff radius. Cutoff strategy works pretty well for
1027 > Lennard-Jones interaction because of its short range nature.
1028 > However, simply truncating the electrostatic interaction with the
1029 > use of cutoff has been shown to lead to severe artifacts in
1030 > simulations. Ewald summation, in which the slowly conditionally
1031 > convergent Coulomb potential is transformed into direct and
1032 > reciprocal sums with rapid and absolute convergence, has proved to
1033 > minimize the periodicity artifacts in liquid simulations. Taking the
1034 > advantages of the fast Fourier transform (FFT) for calculating
1035 > discrete Fourier transforms, the particle mesh-based methods are
1036 > accelerated from $O(N^{3/2})$ to $O(N logN)$. An alternative
1037 > approach is \emph{fast multipole method}, which treats Coulombic
1038 > interaction exactly at short range, and approximate the potential at
1039 > long range through multipolar expansion. In spite of their wide
1040 > acceptances at the molecular simulation community, these two methods
1041 > are hard to be implemented correctly and efficiently. Instead, we
1042 > use a damped and charge-neutralized Coulomb potential method
1043 > developed by Wolf and his coworkers. The shifted Coulomb potential
1044 > for particle $i$ and particle $j$ at distance $r_{rj}$ is given by:
1045 > \begin{equation}
1046 > V(r_{ij})= \frac{q_i q_j \textrm{erfc}(\alpha
1047 > r_{ij})}{r_{ij}}-\lim_{r_{ij}\rightarrow
1048 > R_\textrm{c}}\left\{\frac{q_iq_j \textrm{erfc}(\alpha
1049 > r_{ij})}{r_{ij}}\right\}. \label{introEquation:shiftedCoulomb}
1050 > \end{equation}
1051 > where $\alpha$ is the convergence parameter. Due to the lack of
1052 > inherent periodicity and rapid convergence,this method is extremely
1053 > efficient and easy to implement.
1054 > %\begin{figure}
1055 > %\centering
1056 > %\includegraphics[width=\linewidth]{pbcFig.eps}
1057 > %\caption[An illustration of shifted Coulomb potential]{An illustration of shifted Coulomb potential.}
1058 > %\label{introFigure:shiftedCoulomb}
1059 > %\end{figure}
1060 >
1061 > %multiple time step
1062 >
1063 > \subsection{\label{introSection:Analysis} Analysis}
1064 >
1065 > Recently, advanced visualization technique are widely applied to
1066 > monitor the motions of molecules. Although the dynamics of the
1067 > system can be described qualitatively from animation, quantitative
1068 > trajectory analysis are more appreciable. According to the
1069 > principles of Statistical Mechanics,
1070 > Sec.~\ref{introSection:statisticalMechanics}, one can compute
1071 > thermodynamics properties, analyze fluctuations of structural
1072 > parameters, and investigate time-dependent processes of the molecule
1073 > from the trajectories.
1074 >
1075 > \subsubsection{\label{introSection:thermodynamicsProperties}Thermodynamics Properties}
1076 >
1077 > Thermodynamics properties, which can be expressed in terms of some
1078 > function of the coordinates and momenta of all particles in the
1079 > system, can be directly computed from molecular dynamics. The usual
1080 > way to measure the pressure is based on virial theorem of Clausius
1081 > which states that the virial is equal to $-3Nk_BT$. For a system
1082 > with forces between particles, the total virial, $W$, contains the
1083 > contribution from external pressure and interaction between the
1084 > particles:
1085 > \[
1086 > W =  - 3PV + \left\langle {\sum\limits_{i < j} {r{}_{ij} \cdot
1087 > f_{ij} } } \right\rangle
1088 > \]
1089 > where $f_{ij}$ is the force between particle $i$ and $j$ at a
1090 > distance $r_{ij}$. Thus, the expression for the pressure is given
1091 > by:
1092 > \begin{equation}
1093 > P = \frac{{Nk_B T}}{V} - \frac{1}{{3V}}\left\langle {\sum\limits_{i
1094 > < j} {r{}_{ij} \cdot f_{ij} } } \right\rangle
1095 > \end{equation}
1096 >
1097 > \subsubsection{\label{introSection:structuralProperties}Structural Properties}
1098 >
1099 > Structural Properties of a simple fluid can be described by a set of
1100 > distribution functions. Among these functions,\emph{pair
1101 > distribution function}, also known as \emph{radial distribution
1102 > function}, is of most fundamental importance to liquid-state theory.
1103 > Pair distribution function can be gathered by Fourier transforming
1104 > raw data from a series of neutron diffraction experiments and
1105 > integrating over the surface factor \cite{Powles73}. The experiment
1106 > result can serve as a criterion to justify the correctness of the
1107 > theory. Moreover, various equilibrium thermodynamic and structural
1108 > properties can also be expressed in terms of radial distribution
1109 > function \cite{allen87:csl}.
1110 >
1111 > A pair distribution functions $g(r)$ gives the probability that a
1112 > particle $i$ will be located at a distance $r$ from a another
1113 > particle $j$ in the system
1114 > \[
1115 > g(r) = \frac{V}{{N^2 }}\left\langle {\sum\limits_i {\sum\limits_{j
1116 > \ne i} {\delta (r - r_{ij} )} } } \right\rangle.
1117 > \]
1118 > Note that the delta function can be replaced by a histogram in
1119 > computer simulation. Figure
1120 > \ref{introFigure:pairDistributionFunction} shows a typical pair
1121 > distribution function for the liquid argon system. The occurrence of
1122 > several peaks in the plot of $g(r)$ suggests that it is more likely
1123 > to find particles at certain radial values than at others. This is a
1124 > result of the attractive interaction at such distances. Because of
1125 > the strong repulsive forces at short distance, the probability of
1126 > locating particles at distances less than about 2.5{\AA} from each
1127 > other is essentially zero.
1128 >
1129 > %\begin{figure}
1130 > %\centering
1131 > %\includegraphics[width=\linewidth]{pdf.eps}
1132 > %\caption[Pair distribution function for the liquid argon
1133 > %]{Pair distribution function for the liquid argon}
1134 > %\label{introFigure:pairDistributionFunction}
1135 > %\end{figure}
1136 >
1137 > \subsubsection{\label{introSection:timeDependentProperties}Time-dependent
1138 > Properties}
1139 >
1140 > Time-dependent properties are usually calculated using \emph{time
1141 > correlation function}, which correlates random variables $A$ and $B$
1142 > at two different time
1143 > \begin{equation}
1144 > C_{AB} (t) = \left\langle {A(t)B(0)} \right\rangle.
1145 > \label{introEquation:timeCorrelationFunction}
1146 > \end{equation}
1147 > If $A$ and $B$ refer to same variable, this kind of correlation
1148 > function is called \emph{auto correlation function}. One example of
1149 > auto correlation function is velocity auto-correlation function
1150 > which is directly related to transport properties of molecular
1151 > liquids:
1152 > \[
1153 > D = \frac{1}{3}\int\limits_0^\infty  {\left\langle {v(t) \cdot v(0)}
1154 > \right\rangle } dt
1155 > \]
1156 > where $D$ is diffusion constant. Unlike velocity autocorrelation
1157 > function which is averaging over time origins and over all the
1158 > atoms, dipole autocorrelation are calculated for the entire system.
1159 > The dipole autocorrelation function is given by:
1160 > \[
1161 > c_{dipole}  = \left\langle {u_{tot} (t) \cdot u_{tot} (t)}
1162 > \right\rangle
1163 > \]
1164 > Here $u_{tot}$ is the net dipole of the entire system and is given
1165 > by
1166 > \[
1167 > u_{tot} (t) = \sum\limits_i {u_i (t)}
1168 > \]
1169 > In principle, many time correlation functions can be related with
1170 > Fourier transforms of the infrared, Raman, and inelastic neutron
1171 > scattering spectra of molecular liquids. In practice, one can
1172 > extract the IR spectrum from the intensity of dipole fluctuation at
1173 > each frequency using the following relationship:
1174 > \[
1175 > \hat c_{dipole} (v) = \int_{ - \infty }^\infty  {c_{dipole} (t)e^{ -
1176 > i2\pi vt} dt}
1177 > \]
1178  
1179   \section{\label{introSection:rigidBody}Dynamics of Rigid Bodies}
1180  
# Line 942 | Line 1209 | rotation matrix $A$ and re-formulating Hamiltonian's e
1209   The break through in geometric literature suggests that, in order to
1210   develop a long-term integration scheme, one should preserve the
1211   symplectic structure of the flow. Introducing conjugate momentum to
1212 < rotation matrix $A$ and re-formulating Hamiltonian's equation, a
1212 > rotation matrix $Q$ and re-formulating Hamiltonian's equation, a
1213   symplectic integrator, RSHAKE, was proposed to evolve the
1214   Hamiltonian system in a constraint manifold by iteratively
1215 < satisfying the orthogonality constraint $A_t A = 1$. An alternative
1215 > satisfying the orthogonality constraint $Q_T Q = 1$. An alternative
1216   method using quaternion representation was developed by Omelyan.
1217   However, both of these methods are iterative and inefficient. In
1218   this section, we will present a symplectic Lie-Poisson integrator
1219 < for rigid body developed by Dullweber and his coworkers\cite{}.
1219 > for rigid body developed by Dullweber and his
1220 > coworkers\cite{Dullweber1997} in depth.
1221  
954 \subsection{\label{introSection:lieAlgebra}Lie Algebra}
955
1222   \subsection{\label{introSection:constrainedHamiltonianRB}Constrained Hamiltonian for Rigid Body}
1223 <
1223 > The motion of the rigid body is Hamiltonian with the Hamiltonian
1224 > function
1225   \begin{equation}
1226   H = \frac{1}{2}(p^T m^{ - 1} p) + \frac{1}{2}tr(PJ^{ - 1} P) +
1227   V(q,Q) + \frac{1}{2}tr[(QQ^T  - 1)\Lambda ].
# Line 969 | Line 1236 | Q^T Q = 1$, \label{introEquation:orthogonalConstraint}
1236   where $I_{ii}$ is the diagonal element of the inertia tensor. This
1237   constrained Hamiltonian equation subjects to a holonomic constraint,
1238   \begin{equation}
1239 < Q^T Q = 1$, \label{introEquation:orthogonalConstraint}
1239 > Q^T Q = 1, \label{introEquation:orthogonalConstraint}
1240   \end{equation}
1241   which is used to ensure rotation matrix's orthogonality.
1242   Differentiating \ref{introEquation:orthogonalConstraint} and using
1243   Equation \ref{introEquation:RBMotionMomentum}, one may obtain,
1244   \begin{equation}
1245 < Q^t PJ^{ - 1}  + J^{ - 1} P^t Q = 0 . \\
1245 > Q^T PJ^{ - 1}  + J^{ - 1} P^T Q = 0 . \\
1246   \label{introEquation:RBFirstOrderConstraint}
1247   \end{equation}
1248  
# Line 987 | Line 1254 | the equations of motion,
1254   \frac{{dq}}{{dt}} = \frac{p}{m} \label{introEquation:RBMotionPosition}\\
1255   \frac{{dp}}{{dt}} =  - \nabla _q V(q,Q) \label{introEquation:RBMotionMomentum}\\
1256   \frac{{dQ}}{{dt}} = PJ^{ - 1}  \label{introEquation:RBMotionRotation}\\
1257 < \frac{{dP}}{{dt}} =  - \nabla _q V(q,Q) - 2Q\Lambda . \label{introEquation:RBMotionP}\\
1257 > \frac{{dP}}{{dt}} =  - \nabla _Q V(q,Q) - 2Q\Lambda . \label{introEquation:RBMotionP}\\
1258   \end{array}
1259   \]
1260  
1261 <
1261 > In general, there are two ways to satisfy the holonomic constraints.
1262 > We can use constraint force provided by lagrange multiplier on the
1263 > normal manifold to keep the motion on constraint space. Or we can
1264 > simply evolve the system in constraint manifold. These two methods
1265 > are proved to be equivalent. The holonomic constraint and equations
1266 > of motions define a constraint manifold for rigid body
1267   \[
1268 < M = \left\{ {(Q,P):Q^T Q = 1,Q^t PJ^{ - 1}  + J^{ - 1} P^t Q = 0}
1269 < \right\} .
1268 > M = \left\{ {(Q,P):Q^T Q = 1,Q^T PJ^{ - 1}  + J^{ - 1} P^T Q = 0}
1269 > \right\}.
1270   \]
1271  
1272 < \subsection{\label{introSection:DLMMotionEquation}The Euler Equations of Rigid Body Motion}
1273 <
1274 < \subsection{\label{introSection:symplecticDiscretizationRB}Symplectic Discretization of Euler Equations}
1275 <
1004 <
1005 < \section{\label{introSection:langevinDynamics}Langevin Dynamics}
1006 <
1007 < \subsection{\label{introSection:LDIntroduction}Introduction and application of Langevin Dynamics}
1008 <
1009 < \subsection{\label{introSection:generalizedLangevinDynamics}Generalized Langevin Dynamics}
1010 <
1011 < \begin{equation}
1012 < H = \frac{{p^2 }}{{2m}} + U(x) + H_B  + \Delta U(x,x_1 , \ldots x_N)
1013 < \label{introEquation:bathGLE}
1014 < \end{equation}
1015 < where $H_B$ is harmonic bath Hamiltonian,
1272 > Unfortunately, this constraint manifold is not the cotangent bundle
1273 > $T_{\star}SO(3)$. However, it turns out that under symplectic
1274 > transformation, the cotangent space and the phase space are
1275 > diffeomorphic. Introducing
1276   \[
1277 < H_B  =\sum\limits_{\alpha  = 1}^N {\left\{ {\frac{{p_\alpha ^2
1018 < }}{{2m_\alpha  }} + \frac{1}{2}m_\alpha  w_\alpha ^2 } \right\}}
1277 > \tilde Q = Q,\tilde P = \frac{1}{2}\left( {P - QP^T Q} \right),
1278   \]
1279 < and $\Delta U$ is bilinear system-bath coupling,
1279 > the mechanical system subject to a holonomic constraint manifold $M$
1280 > can be re-formulated as a Hamiltonian system on the cotangent space
1281   \[
1282 < \Delta U =  - \sum\limits_{\alpha  = 1}^N {g_\alpha  x_\alpha  x}
1282 > T^* SO(3) = \left\{ {(\tilde Q,\tilde P):\tilde Q^T \tilde Q =
1283 > 1,\tilde Q^T \tilde PJ^{ - 1}  + J^{ - 1} P^T \tilde Q = 0} \right\}
1284   \]
1285 < Completing the square,
1285 >
1286 > For a body fixed vector $X_i$ with respect to the center of mass of
1287 > the rigid body, its corresponding lab fixed vector $X_0^{lab}$  is
1288 > given as
1289 > \begin{equation}
1290 > X_i^{lab} = Q X_i + q.
1291 > \end{equation}
1292 > Therefore, potential energy $V(q,Q)$ is defined by
1293   \[
1294 < H_B  + \Delta U = \sum\limits_{\alpha  = 1}^N {\left\{
1027 < {\frac{{p_\alpha ^2 }}{{2m_\alpha  }} + \frac{1}{2}m_\alpha
1028 < w_\alpha ^2 \left( {x_\alpha   - \frac{{g_\alpha  }}{{m_\alpha
1029 < w_\alpha ^2 }}x} \right)^2 } \right\}}  - \sum\limits_{\alpha  =
1030 < 1}^N {\frac{{g_\alpha ^2 }}{{2m_\alpha  w_\alpha ^2 }}} x^2
1294 > V(q,Q) = V(Q X_0 + q).
1295   \]
1296 < and putting it back into Eq.~\ref{introEquation:bathGLE},
1296 > Hence, the force and torque are given by
1297   \[
1298 < H = \frac{{p^2 }}{{2m}} + W(x) + \sum\limits_{\alpha  = 1}^N
1035 < {\left\{ {\frac{{p_\alpha ^2 }}{{2m_\alpha  }} + \frac{1}{2}m_\alpha
1036 < w_\alpha ^2 \left( {x_\alpha   - \frac{{g_\alpha  }}{{m_\alpha
1037 < w_\alpha ^2 }}x} \right)^2 } \right\}}
1298 > \nabla _q V(q,Q) = F(q,Q) = \sum\limits_i {F_i (q,Q)},
1299   \]
1300 < where
1300 > and
1301   \[
1302 < W(x) = U(x) - \sum\limits_{\alpha  = 1}^N {\frac{{g_\alpha ^2
1042 < }}{{2m_\alpha  w_\alpha ^2 }}} x^2
1302 > \nabla _Q V(q,Q) = F(q,Q)X_i^t
1303   \]
1304 < Since the first two terms of the new Hamiltonian depend only on the
1045 < system coordinates, we can get the equations of motion for
1046 < Generalized Langevin Dynamics by Hamilton's equations
1047 < \ref{introEquation:motionHamiltonianCoordinate,
1048 < introEquation:motionHamiltonianMomentum},
1049 < \begin{align}
1050 < \dot p &=  - \frac{{\partial H}}{{\partial x}}
1051 <       &= m\ddot x
1052 <       &= - \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)}
1053 < \label{introEquation:Lp5}
1054 < \end{align}
1055 < , and
1056 < \begin{align}
1057 < \dot p_\alpha   &=  - \frac{{\partial H}}{{\partial x_\alpha  }}
1058 <                &= m\ddot x_\alpha
1059 <                &= \- m_\alpha  w_\alpha ^2 \left( {x_\alpha   - \frac{{g_\alpha}}{{m_\alpha  w_\alpha ^2 }}x} \right)
1060 < \end{align}
1304 > respectively.
1305  
1306 < \subsection{\label{introSection:laplaceTransform}The Laplace Transform}
1307 <
1306 > As a common choice to describe the rotation dynamics of the rigid
1307 > body, angular momentum on body frame $\Pi  = Q^t P$ is introduced to
1308 > rewrite the equations of motion,
1309 > \begin{equation}
1310 > \begin{array}{l}
1311 > \mathop \Pi \limits^ \bullet   = J^{ - 1} \Pi ^T \Pi  + Q^T \sum\limits_i {F_i (q,Q)X_i^T }  - \Lambda  \\
1312 > \mathop Q\limits^{{\rm{   }} \bullet }  = Q\Pi {\rm{ }}J^{ - 1}  \\
1313 > \end{array}
1314 > \label{introEqaution:RBMotionPI}
1315 > \end{equation}
1316 > , as well as holonomic constraints,
1317   \[
1318 < L(x) = \int_0^\infty  {x(t)e^{ - pt} dt}
1318 > \begin{array}{l}
1319 > \Pi J^{ - 1}  + J^{ - 1} \Pi ^t  = 0 \\
1320 > Q^T Q = 1 \\
1321 > \end{array}
1322   \]
1323  
1324 + For a vector $v(v_1 ,v_2 ,v_3 ) \in R^3$ and a matrix $\hat v \in
1325 + so(3)^ \star$, the hat-map isomorphism,
1326 + \begin{equation}
1327 + v(v_1 ,v_2 ,v_3 ) \Leftrightarrow \hat v = \left(
1328 + {\begin{array}{*{20}c}
1329 +   0 & { - v_3 } & {v_2 }  \\
1330 +   {v_3 } & 0 & { - v_1 }  \\
1331 +   { - v_2 } & {v_1 } & 0  \\
1332 + \end{array}} \right),
1333 + \label{introEquation:hatmapIsomorphism}
1334 + \end{equation}
1335 + will let us associate the matrix products with traditional vector
1336 + operations
1337   \[
1338 < L(x + y) = L(x) + L(y)
1338 > \hat vu = v \times u
1339   \]
1340  
1341 + Using \ref{introEqaution:RBMotionPI}, one can construct a skew
1342 + matrix,
1343 + \begin{equation}
1344 + (\mathop \Pi \limits^ \bullet   - \mathop \Pi \limits^ \bullet  ^T
1345 + ){\rm{ }} = {\rm{ }}(\Pi  - \Pi ^T ){\rm{ }}(J^{ - 1} \Pi  + \Pi J^{
1346 + - 1} ) + \sum\limits_i {[Q^T F_i (r,Q)X_i^T  - X_i F_i (r,Q)^T Q]} -
1347 + (\Lambda  - \Lambda ^T ) . \label{introEquation:skewMatrixPI}
1348 + \end{equation}
1349 + Since $\Lambda$ is symmetric, the last term of Equation
1350 + \ref{introEquation:skewMatrixPI} is zero, which implies the Lagrange
1351 + multiplier $\Lambda$ is absent from the equations of motion. This
1352 + unique property eliminate the requirement of iterations which can
1353 + not be avoided in other methods\cite{}.
1354 +
1355 + Applying hat-map isomorphism, we obtain the equation of motion for
1356 + angular momentum on body frame
1357 + \begin{equation}
1358 + \dot \pi  = \pi  \times I^{ - 1} \pi  + \sum\limits_i {\left( {Q^T
1359 + F_i (r,Q)} \right) \times X_i }.
1360 + \label{introEquation:bodyAngularMotion}
1361 + \end{equation}
1362 + In the same manner, the equation of motion for rotation matrix is
1363 + given by
1364   \[
1365 < L(ax) = aL(x)
1365 > \dot Q = Qskew(I^{ - 1} \pi )
1366   \]
1367  
1368 + \subsection{\label{introSection:SymplecticFreeRB}Symplectic
1369 + Lie-Poisson Integrator for Free Rigid Body}
1370 +
1371 + If there is not external forces exerted on the rigid body, the only
1372 + contribution to the rotational is from the kinetic potential (the
1373 + first term of \ref{ introEquation:bodyAngularMotion}). The free
1374 + rigid body is an example of Lie-Poisson system with Hamiltonian
1375 + function
1376 + \begin{equation}
1377 + T^r (\pi ) = T_1 ^r (\pi _1 ) + T_2^r (\pi _2 ) + T_3^r (\pi _3 )
1378 + \label{introEquation:rotationalKineticRB}
1379 + \end{equation}
1380 + where $T_i^r (\pi _i ) = \frac{{\pi _i ^2 }}{{2I_i }}$ and
1381 + Lie-Poisson structure matrix,
1382 + \begin{equation}
1383 + J(\pi ) = \left( {\begin{array}{*{20}c}
1384 +   0 & {\pi _3 } & { - \pi _2 }  \\
1385 +   { - \pi _3 } & 0 & {\pi _1 }  \\
1386 +   {\pi _2 } & { - \pi _1 } & 0  \\
1387 + \end{array}} \right)
1388 + \end{equation}
1389 + Thus, the dynamics of free rigid body is governed by
1390 + \begin{equation}
1391 + \frac{d}{{dt}}\pi  = J(\pi )\nabla _\pi  T^r (\pi )
1392 + \end{equation}
1393 +
1394 + One may notice that each $T_i^r$ in Equation
1395 + \ref{introEquation:rotationalKineticRB} can be solved exactly. For
1396 + instance, the equations of motion due to $T_1^r$ are given by
1397 + \begin{equation}
1398 + \frac{d}{{dt}}\pi  = R_1 \pi ,\frac{d}{{dt}}Q = QR_1
1399 + \label{introEqaution:RBMotionSingleTerm}
1400 + \end{equation}
1401 + where
1402 + \[ R_1  = \left( {\begin{array}{*{20}c}
1403 +   0 & 0 & 0  \\
1404 +   0 & 0 & {\pi _1 }  \\
1405 +   0 & { - \pi _1 } & 0  \\
1406 + \end{array}} \right).
1407 + \]
1408 + The solutions of Equation \ref{introEqaution:RBMotionSingleTerm} is
1409   \[
1410 < L(\dot x) = pL(x) - px(0)
1410 > \pi (\Delta t) = e^{\Delta tR_1 } \pi (0),Q(\Delta t) =
1411 > Q(0)e^{\Delta tR_1 }
1412   \]
1413 <
1413 > with
1414   \[
1415 < L(\ddot x) = p^2 L(x) - px(0) - \dot x(0)
1415 > e^{\Delta tR_1 }  = \left( {\begin{array}{*{20}c}
1416 >   0 & 0 & 0  \\
1417 >   0 & {\cos \theta _1 } & {\sin \theta _1 }  \\
1418 >   0 & { - \sin \theta _1 } & {\cos \theta _1 }  \\
1419 > \end{array}} \right),\theta _1  = \frac{{\pi _1 }}{{I_1 }}\Delta t.
1420   \]
1421 <
1421 > To reduce the cost of computing expensive functions in $e^{\Delta
1422 > tR_1 }$, we can use Cayley transformation,
1423   \[
1424 < L\left( {\int_0^t {g(t - \tau )h(\tau )d\tau } } \right) = G(p)H(p)
1424 > e^{\Delta tR_1 }  \approx (1 - \Delta tR_1 )^{ - 1} (1 + \Delta tR_1
1425 > )
1426   \]
1427 + The flow maps for $T_2^r$ and $T_3^r$ can be found in the same
1428 + manner.
1429  
1430 < Some relatively important transformation,
1430 > In order to construct a second-order symplectic method, we split the
1431 > angular kinetic Hamiltonian function can into five terms
1432   \[
1433 < L(\cos at) = \frac{p}{{p^2  + a^2 }}
1433 > T^r (\pi ) = \frac{1}{2}T_1 ^r (\pi _1 ) + \frac{1}{2}T_2^r (\pi _2
1434 > ) + T_3^r (\pi _3 ) + \frac{1}{2}T_2^r (\pi _2 ) + \frac{1}{2}T_1 ^r
1435 > (\pi _1 )
1436 > \].
1437 > Concatenating flows corresponding to these five terms, we can obtain
1438 > an symplectic integrator,
1439 > \[
1440 > \varphi _{\Delta t,T^r }  = \varphi _{\Delta t/2,\pi _1 }  \circ
1441 > \varphi _{\Delta t/2,\pi _2 }  \circ \varphi _{\Delta t,\pi _3 }
1442 > \circ \varphi _{\Delta t/2,\pi _2 }  \circ \varphi _{\Delta t/2,\pi
1443 > _1 }.
1444   \]
1445  
1446 + The non-canonical Lie-Poisson bracket ${F, G}$ of two function
1447 + $F(\pi )$ and $G(\pi )$ is defined by
1448   \[
1449 < L(\sin at) = \frac{a}{{p^2  + a^2 }}
1449 > \{ F,G\} (\pi ) = [\nabla _\pi  F(\pi )]^T J(\pi )\nabla _\pi  G(\pi
1450 > )
1451   \]
1452 <
1452 > If the Poisson bracket of a function $F$ with an arbitrary smooth
1453 > function $G$ is zero, $F$ is a \emph{Casimir}, which is the
1454 > conserved quantity in Poisson system. We can easily verify that the
1455 > norm of the angular momentum, $\parallel \pi
1456 > \parallel$, is a \emph{Casimir}. Let$ F(\pi ) = S(\frac{{\parallel
1457 > \pi \parallel ^2 }}{2})$ for an arbitrary function $ S:R \to R$ ,
1458 > then by the chain rule
1459   \[
1460 < L(1) = \frac{1}{p}
1460 > \nabla _\pi  F(\pi ) = S'(\frac{{\parallel \pi \parallel ^2
1461 > }}{2})\pi
1462   \]
1463 + Thus $ [\nabla _\pi  F(\pi )]^T J(\pi ) =  - S'(\frac{{\parallel \pi
1464 + \parallel ^2 }}{2})\pi  \times \pi  = 0 $. This explicit
1465 + Lie-Poisson integrator is found to be extremely efficient and stable
1466 + which can be explained by the fact the small angle approximation is
1467 + used and the norm of the angular momentum is conserved.
1468  
1469 < First, the bath coordinates,
1469 > \subsection{\label{introSection:RBHamiltonianSplitting} Hamiltonian
1470 > Splitting for Rigid Body}
1471 >
1472 > The Hamiltonian of rigid body can be separated in terms of kinetic
1473 > energy and potential energy,
1474   \[
1475 < p^2 L(x_\alpha  ) - px_\alpha  (0) - \dot x_\alpha  (0) =  - \omega
1104 < _\alpha ^2 L(x_\alpha  ) + \frac{{g_\alpha  }}{{\omega _\alpha
1105 < }}L(x)
1475 > H = T(p,\pi ) + V(q,Q)
1476   \]
1477 < \[
1478 < L(x_\alpha  ) = \frac{{\frac{{g_\alpha  }}{{\omega _\alpha  }}L(x) +
1479 < px_\alpha  (0) + \dot x_\alpha  (0)}}{{p^2  + \omega _\alpha ^2 }}
1477 > The equations of motion corresponding to potential energy and
1478 > kinetic energy are listed in the below table,
1479 > \begin{table}
1480 > \caption{Equations of motion due to Potential and Kinetic Energies}
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 > \end{table}
1494 > A second-order symplectic method is now obtained by the
1495 > composition of the flow maps,
1496 > \[
1497 > \varphi _{\Delta t}  = \varphi _{\Delta t/2,V}  \circ \varphi
1498 > _{\Delta t,T}  \circ \varphi _{\Delta t/2,V}.
1499   \]
1500 < Then, the system coordinates,
1501 < \begin{align}
1502 < mL(\ddot x) &=  - \frac{1}{p}\frac{{\partial W(x)}}{{\partial x}} -
1503 < \sum\limits_{\alpha  = 1}^N {\left\{ {\frac{{\frac{{g_\alpha
1504 < }}{{\omega _\alpha  }}L(x) + px_\alpha  (0) + \dot x_\alpha
1505 < (0)}}{{p^2  + \omega _\alpha ^2 }} - \frac{{g_\alpha ^2 }}{{m_\alpha
1506 < }}\omega _\alpha ^2 L(x)} \right\}}
1507 < %
1508 < &= - \frac{1}{p}\frac{{\partial W(x)}}{{\partial x}} -
1120 < \sum\limits_{\alpha  = 1}^N {\left\{ { - \frac{{g_\alpha ^2 }}{{m_\alpha  \omega _\alpha ^2 }}\frac{p}{{p^2  + \omega _\alpha ^2 }}pL(x)
1121 < - \frac{p}{{p^2  + \omega _\alpha ^2 }}g_\alpha  x_\alpha  (0)
1122 < - \frac{1}{{p^2  + \omega _\alpha ^2 }}g_\alpha  \dot x_\alpha  (0)} \right\}}
1123 < \end{align}
1124 < Then, the inverse transform,
1500 > Moreover, $\varphi _{\Delta t/2,V}$ can be divided into two
1501 > sub-flows which corresponding to force and torque respectively,
1502 > \[
1503 > \varphi _{\Delta t/2,V}  = \varphi _{\Delta t/2,F}  \circ \varphi
1504 > _{\Delta t/2,\tau }.
1505 > \]
1506 > Since the associated operators of $\varphi _{\Delta t/2,F} $ and
1507 > $\circ \varphi _{\Delta t/2,\tau }$ are commuted, the composition
1508 > order inside $\varphi _{\Delta t/2,V}$ does not matter.
1509  
1510 + Furthermore, kinetic potential can be separated to translational
1511 + kinetic term, $T^t (p)$, and rotational kinetic term, $T^r (\pi )$,
1512 + \begin{equation}
1513 + T(p,\pi ) =T^t (p) + T^r (\pi ).
1514 + \end{equation}
1515 + where $ T^t (p) = \frac{1}{2}p^T m^{ - 1} p $ and $T^r (\pi )$ is
1516 + defined by \ref{introEquation:rotationalKineticRB}. Therefore, the
1517 + corresponding flow maps are given by
1518 + \[
1519 + \varphi _{\Delta t,T}  = \varphi _{\Delta t,T^t }  \circ \varphi
1520 + _{\Delta t,T^r }.
1521 + \]
1522 + Finally, we obtain the overall symplectic flow maps for free moving
1523 + rigid body
1524 + \begin{equation}
1525 + \begin{array}{c}
1526 + \varphi _{\Delta t}  = \varphi _{\Delta t/2,F}  \circ \varphi _{\Delta t/2,\tau }  \\
1527 +  \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 }  \\
1528 +  \circ \varphi _{\Delta t/2,\tau }  \circ \varphi _{\Delta t/2,F}  .\\
1529 + \end{array}
1530 + \label{introEquation:overallRBFlowMaps}
1531 + \end{equation}
1532 +
1533 + \section{\label{introSection:langevinDynamics}Langevin Dynamics}
1534 + As an alternative to newtonian dynamics, Langevin dynamics, which
1535 + mimics a simple heat bath with stochastic and dissipative forces,
1536 + has been applied in a variety of studies. This section will review
1537 + the theory of Langevin dynamics simulation. A brief derivation of
1538 + generalized Langevin equation will be given first. Follow that, we
1539 + will discuss the physical meaning of the terms appearing in the
1540 + equation as well as the calculation of friction tensor from
1541 + hydrodynamics theory.
1542 +
1543 + \subsection{\label{introSection:generalizedLangevinDynamics}Derivation of Generalized Langevin Equation}
1544 +
1545 + Harmonic bath model, in which an effective set of harmonic
1546 + oscillators are used to mimic the effect of a linearly responding
1547 + environment, has been widely used in quantum chemistry and
1548 + statistical mechanics. One of the successful applications of
1549 + Harmonic bath model is the derivation of Deriving Generalized
1550 + Langevin Dynamics. Lets consider a system, in which the degree of
1551 + freedom $x$ is assumed to couple to the bath linearly, giving a
1552 + Hamiltonian of the form
1553 + \begin{equation}
1554 + H = \frac{{p^2 }}{{2m}} + U(x) + H_B  + \Delta U(x,x_1 , \ldots x_N)
1555 + \label{introEquation:bathGLE}.
1556 + \end{equation}
1557 + Here $p$ is a momentum conjugate to $q$, $m$ is the mass associated
1558 + with this degree of freedom, $H_B$ is harmonic bath Hamiltonian,
1559 + \[
1560 + H_B  = \sum\limits_{\alpha  = 1}^N {\left\{ {\frac{{p_\alpha ^2
1561 + }}{{2m_\alpha  }} + \frac{1}{2}m_\alpha  \omega _\alpha ^2 }
1562 + \right\}}
1563 + \]
1564 + where the index $\alpha$ runs over all the bath degrees of freedom,
1565 + $\omega _\alpha$ are the harmonic bath frequencies, $m_\alpha$ are
1566 + the harmonic bath masses, and $\Delta U$ is bilinear system-bath
1567 + coupling,
1568 + \[
1569 + \Delta U =  - \sum\limits_{\alpha  = 1}^N {g_\alpha  x_\alpha  x}
1570 + \]
1571 + where $g_\alpha$ are the coupling constants between the bath and the
1572 + coordinate $x$. Introducing
1573 + \[
1574 + W(x) = U(x) - \sum\limits_{\alpha  = 1}^N {\frac{{g_\alpha ^2
1575 + }}{{2m_\alpha  w_\alpha ^2 }}} x^2
1576 + \] and combining the last two terms in Equation
1577 + \ref{introEquation:bathGLE}, we may rewrite the Harmonic bath
1578 + Hamiltonian as
1579 + \[
1580 + H = \frac{{p^2 }}{{2m}} + W(x) + \sum\limits_{\alpha  = 1}^N
1581 + {\left\{ {\frac{{p_\alpha ^2 }}{{2m_\alpha  }} + \frac{1}{2}m_\alpha
1582 + w_\alpha ^2 \left( {x_\alpha   - \frac{{g_\alpha  }}{{m_\alpha
1583 + w_\alpha ^2 }}x} \right)^2 } \right\}}
1584 + \]
1585 + Since the first two terms of the new Hamiltonian depend only on the
1586 + system coordinates, we can get the equations of motion for
1587 + Generalized Langevin Dynamics by Hamilton's equations
1588 + \ref{introEquation:motionHamiltonianCoordinate,
1589 + introEquation:motionHamiltonianMomentum},
1590 + \begin{equation}
1591 + m\ddot x =  - \frac{{\partial W(x)}}{{\partial x}} -
1592 + \sum\limits_{\alpha  = 1}^N {g_\alpha  \left( {x_\alpha   -
1593 + \frac{{g_\alpha  }}{{m_\alpha  w_\alpha ^2 }}x} \right)},
1594 + \label{introEquation:coorMotionGLE}
1595 + \end{equation}
1596 + and
1597 + \begin{equation}
1598 + m\ddot x_\alpha   =  - m_\alpha  w_\alpha ^2 \left( {x_\alpha   -
1599 + \frac{{g_\alpha  }}{{m_\alpha  w_\alpha ^2 }}x} \right).
1600 + \label{introEquation:bathMotionGLE}
1601 + \end{equation}
1602 +
1603 + In order to derive an equation for $x$, the dynamics of the bath
1604 + variables $x_\alpha$ must be solved exactly first. As an integral
1605 + transform which is particularly useful in solving linear ordinary
1606 + differential equations, Laplace transform is the appropriate tool to
1607 + solve this problem. The basic idea is to transform the difficult
1608 + differential equations into simple algebra problems which can be
1609 + solved easily. Then applying inverse Laplace transform, also known
1610 + as the Bromwich integral, we can retrieve the solutions of the
1611 + original problems.
1612 +
1613 + Let $f(t)$ be a function defined on $ [0,\infty ) $. The Laplace
1614 + transform of f(t) is a new function defined as
1615 + \[
1616 + L(f(t)) \equiv F(p) = \int_0^\infty  {f(t)e^{ - pt} dt}
1617 + \]
1618 + where  $p$ is real and  $L$ is called the Laplace Transform
1619 + Operator. Below are some important properties of Laplace transform
1620 + \begin{equation}
1621 + \begin{array}{c}
1622 + L(x + y) = L(x) + L(y) \\
1623 + L(ax) = aL(x) \\
1624 + L(\dot x) = pL(x) - px(0) \\
1625 + L(\ddot x) = p^2 L(x) - px(0) - \dot x(0) \\
1626 + L\left( {\int_0^t {g(t - \tau )h(\tau )d\tau } } \right) = G(p)H(p) \\
1627 + \end{array}
1628 + \end{equation}
1629 +
1630 + Applying Laplace transform to the bath coordinates, we obtain
1631 + \[
1632 + \begin{array}{c}
1633 + p^2 L(x_\alpha  ) - px_\alpha  (0) - \dot x_\alpha  (0) =  - \omega _\alpha ^2 L(x_\alpha  ) + \frac{{g_\alpha  }}{{\omega _\alpha  }}L(x) \\
1634 + L(x_\alpha  ) = \frac{{\frac{{g_\alpha  }}{{\omega _\alpha  }}L(x) + px_\alpha  (0) + \dot x_\alpha  (0)}}{{p^2  + \omega _\alpha ^2 }} \\
1635 + \end{array}
1636 + \]
1637 + By the same way, the system coordinates become
1638 + \[
1639 + \begin{array}{c}
1640 + mL(\ddot x) =  - \frac{1}{p}\frac{{\partial W(x)}}{{\partial x}} \\
1641 +  - \sum\limits_{\alpha  = 1}^N {\left\{ { - \frac{{g_\alpha ^2 }}{{m_\alpha  \omega _\alpha ^2 }}\frac{p}{{p^2  + \omega _\alpha ^2 }}pL(x) - \frac{p}{{p^2  + \omega _\alpha ^2 }}g_\alpha  x_\alpha  (0) - \frac{1}{{p^2  + \omega _\alpha ^2 }}g_\alpha  \dot x_\alpha  (0)} \right\}}  \\
1642 + \end{array}
1643 + \]
1644 +
1645 + With the help of some relatively important inverse Laplace
1646 + transformations:
1647 + \[
1648 + \begin{array}{c}
1649 + L(\cos at) = \frac{p}{{p^2  + a^2 }} \\
1650 + L(\sin at) = \frac{a}{{p^2  + a^2 }} \\
1651 + L(1) = \frac{1}{p} \\
1652 + \end{array}
1653 + \]
1654 + , we obtain
1655   \begin{align}
1656   m\ddot x &=  - \frac{{\partial W(x)}}{{\partial x}} -
1657   \sum\limits_{\alpha  = 1}^N {\left\{ {\left( { - \frac{{g_\alpha ^2
# Line 1142 | Line 1671 | t)\dot x(t - \tau )d} \tau }  + \sum\limits_{\alpha  =
1671   (\omega _\alpha  t)} \right\}}
1672   \end{align}
1673  
1674 + Introducing a \emph{dynamic friction kernel}
1675   \begin{equation}
1676 + \xi (t) = \sum\limits_{\alpha  = 1}^N {\left( { - \frac{{g_\alpha ^2
1677 + }}{{m_\alpha  \omega _\alpha ^2 }}} \right)\cos (\omega _\alpha  t)}
1678 + \label{introEquation:dynamicFrictionKernelDefinition}
1679 + \end{equation}
1680 + and \emph{a random force}
1681 + \begin{equation}
1682 + R(t) = \sum\limits_{\alpha  = 1}^N {\left( {g_\alpha  x_\alpha  (0)
1683 + - \frac{{g_\alpha ^2 }}{{m_\alpha  \omega _\alpha ^2 }}x(0)}
1684 + \right)\cos (\omega _\alpha  t)}  + \frac{{\dot x_\alpha
1685 + (0)}}{{\omega _\alpha  }}\sin (\omega _\alpha  t),
1686 + \label{introEquation:randomForceDefinition}
1687 + \end{equation}
1688 + the equation of motion can be rewritten as
1689 + \begin{equation}
1690   m\ddot x =  - \frac{{\partial W}}{{\partial x}} - \int_0^t {\xi
1691   (t)\dot x(t - \tau )d\tau }  + R(t)
1692   \label{introEuqation:GeneralizedLangevinDynamics}
1693   \end{equation}
1694 < %where $ {\xi (t)}$ is friction kernel, $R(t)$ is random force and
1695 < %$W$ is the potential of mean force. $W(x) =  - kT\ln p(x)$
1694 > which is known as the \emph{generalized Langevin equation}.
1695 >
1696 > \subsubsection{\label{introSection:randomForceDynamicFrictionKernel}Random Force and Dynamic Friction Kernel}
1697 >
1698 > One may notice that $R(t)$ depends only on initial conditions, which
1699 > implies it is completely deterministic within the context of a
1700 > harmonic bath. However, it is easy to verify that $R(t)$ is totally
1701 > uncorrelated to $x$ and $\dot x$,
1702   \[
1703 < \xi (t) = \sum\limits_{\alpha  = 1}^N {\left( { - \frac{{g_\alpha ^2
1704 < }}{{m_\alpha  \omega _\alpha ^2 }}} \right)\cos (\omega _\alpha  t)}
1703 > \begin{array}{l}
1704 > \left\langle {x(t)R(t)} \right\rangle  = 0, \\
1705 > \left\langle {\dot x(t)R(t)} \right\rangle  = 0. \\
1706 > \end{array}
1707   \]
1708 < For an infinite harmonic bath, we can use the spectral density and
1709 < an integral over frequencies.
1708 > This property is what we expect from a truly random process. As long
1709 > as the model, which is gaussian distribution in general, chosen for
1710 > $R(t)$ is a truly random process, the stochastic nature of the GLE
1711 > still remains.
1712  
1713 + %dynamic friction kernel
1714 + The convolution integral
1715   \[
1716 < R(t) = \sum\limits_{\alpha  = 1}^N {\left( {g_\alpha  x_\alpha  (0)
1161 < - \frac{{g_\alpha ^2 }}{{m_\alpha  \omega _\alpha ^2 }}x(0)}
1162 < \right)\cos (\omega _\alpha  t)}  + \frac{{\dot x_\alpha
1163 < (0)}}{{\omega _\alpha  }}\sin (\omega _\alpha  t)
1716 > \int_0^t {\xi (t)\dot x(t - \tau )d\tau }
1717   \]
1718 < The random forces depend only on initial conditions.
1718 > depends on the entire history of the evolution of $x$, which implies
1719 > that the bath retains memory of previous motions. In other words,
1720 > the bath requires a finite time to respond to change in the motion
1721 > of the system. For a sluggish bath which responds slowly to changes
1722 > in the system coordinate, we may regard $\xi(t)$ as a constant
1723 > $\xi(t) = \Xi_0$. Hence, the convolution integral becomes
1724 > \[
1725 > \int_0^t {\xi (t)\dot x(t - \tau )d\tau }  = \xi _0 (x(t) - x(0))
1726 > \]
1727 > and Equation \ref{introEuqation:GeneralizedLangevinDynamics} becomes
1728 > \[
1729 > m\ddot x =  - \frac{\partial }{{\partial x}}\left( {W(x) +
1730 > \frac{1}{2}\xi _0 (x - x_0 )^2 } \right) + R(t),
1731 > \]
1732 > which can be used to describe dynamic caging effect. The other
1733 > extreme is the bath that responds infinitely quickly to motions in
1734 > the system. Thus, $\xi (t)$ can be taken as a $delta$ function in
1735 > time:
1736 > \[
1737 > \xi (t) = 2\xi _0 \delta (t)
1738 > \]
1739 > Hence, the convolution integral becomes
1740 > \[
1741 > \int_0^t {\xi (t)\dot x(t - \tau )d\tau }  = 2\xi _0 \int_0^t
1742 > {\delta (t)\dot x(t - \tau )d\tau }  = \xi _0 \dot x(t),
1743 > \]
1744 > and Equation \ref{introEuqation:GeneralizedLangevinDynamics} becomes
1745 > \begin{equation}
1746 > m\ddot x =  - \frac{{\partial W(x)}}{{\partial x}} - \xi _0 \dot
1747 > x(t) + R(t) \label{introEquation:LangevinEquation}
1748 > \end{equation}
1749 > which is known as the Langevin equation. The static friction
1750 > coefficient $\xi _0$ can either be calculated from spectral density
1751 > or be determined by Stokes' law for regular shaped particles.A
1752 > briefly review on calculating friction tensor for arbitrary shaped
1753 > particles is given in Sec.~\ref{introSection:frictionTensor}.
1754  
1755   \subsubsection{\label{introSection:secondFluctuationDissipation}The Second Fluctuation Dissipation Theorem}
1756 < So we can define a new set of coordinates,
1756 >
1757 > Defining a new set of coordinates,
1758   \[
1759   q_\alpha  (t) = x_\alpha  (t) - \frac{1}{{m_\alpha  \omega _\alpha
1760   ^2 }}x(0)
1761 < \]
1762 < This makes
1761 > \],
1762 > we can rewrite $R(T)$ as
1763   \[
1764 < R(t) = \sum\limits_{\alpha  = 1}^N {g_\alpha  q_\alpha  (t)}
1764 > R(t) = \sum\limits_{\alpha  = 1}^N {g_\alpha  q_\alpha  (t)}.
1765   \]
1766   And since the $q$ coordinates are harmonic oscillators,
1767   \[
1768 < \begin{array}{l}
1768 > \begin{array}{c}
1769 > \left\langle {q_\alpha ^2 } \right\rangle  = \frac{{kT}}{{m_\alpha  \omega _\alpha ^2 }} \\
1770   \left\langle {q_\alpha  (t)q_\alpha  (0)} \right\rangle  = \left\langle {q_\alpha ^2 (0)} \right\rangle \cos (\omega _\alpha  t) \\
1771   \left\langle {q_\alpha  (t)q_\beta  (0)} \right\rangle  = \delta _{\alpha \beta } \left\langle {q_\alpha  (t)q_\alpha  (0)} \right\rangle  \\
1772 + \left\langle {R(t)R(0)} \right\rangle  = \sum\limits_\alpha  {\sum\limits_\beta  {g_\alpha  g_\beta  \left\langle {q_\alpha  (t)q_\beta  (0)} \right\rangle } }  \\
1773 +  = \sum\limits_\alpha  {g_\alpha ^2 \left\langle {q_\alpha ^2 (0)} \right\rangle \cos (\omega _\alpha  t)}  \\
1774 +  = kT\xi (t) \\
1775   \end{array}
1776 + \]
1777 + Thus, we recover the \emph{second fluctuation dissipation theorem}
1778 + \begin{equation}
1779 + \xi (t) = \left\langle {R(t)R(0)} \right\rangle
1780 + \label{introEquation:secondFluctuationDissipation}.
1781 + \end{equation}
1782 + In effect, it acts as a constraint on the possible ways in which one
1783 + can model the random force and friction kernel.
1784 +
1785 + \subsection{\label{introSection:frictionTensor} Friction Tensor}
1786 + Theoretically, the friction kernel can be determined using velocity
1787 + autocorrelation function. However, this approach become impractical
1788 + when the system become more and more complicate. Instead, various
1789 + approaches based on hydrodynamics have been developed to calculate
1790 + the friction coefficients. The friction effect is isotropic in
1791 + Equation, $\zeta$ can be taken as a scalar. In general, friction
1792 + tensor $\Xi$ is a $6\times 6$ matrix given by
1793 + \[
1794 + \Xi  = \left( {\begin{array}{*{20}c}
1795 +   {\Xi _{}^{tt} } & {\Xi _{}^{rt} }  \\
1796 +   {\Xi _{}^{tr} } & {\Xi _{}^{rr} }  \\
1797 + \end{array}} \right).
1798   \]
1799 + Here, $ {\Xi^{tt} }$ and $ {\Xi^{rr} }$ are translational friction
1800 + tensor and rotational resistance (friction) tensor respectively,
1801 + while ${\Xi^{tr} }$ is translation-rotation coupling tensor and $
1802 + {\Xi^{rt} }$ is rotation-translation coupling tensor. When a
1803 + particle moves in a fluid, it may experience friction force or
1804 + torque along the opposite direction of the velocity or angular
1805 + velocity,
1806 + \[
1807 + \left( \begin{array}{l}
1808 + F_R  \\
1809 + \tau _R  \\
1810 + \end{array} \right) =  - \left( {\begin{array}{*{20}c}
1811 +   {\Xi ^{tt} } & {\Xi ^{rt} }  \\
1812 +   {\Xi ^{tr} } & {\Xi ^{rr} }  \\
1813 + \end{array}} \right)\left( \begin{array}{l}
1814 + v \\
1815 + w \\
1816 + \end{array} \right)
1817 + \]
1818 + where $F_r$ is the friction force and $\tau _R$ is the friction
1819 + toque.
1820  
1821 < \begin{align}
1186 < \left\langle {R(t)R(0)} \right\rangle  &= \sum\limits_\alpha
1187 < {\sum\limits_\beta  {g_\alpha  g_\beta  \left\langle {q_\alpha
1188 < (t)q_\beta  (0)} \right\rangle } }
1189 < %
1190 < &= \sum\limits_\alpha  {g_\alpha ^2 \left\langle {q_\alpha ^2 (0)}
1191 < \right\rangle \cos (\omega _\alpha  t)}
1192 < %
1193 < &= kT\xi (t)
1194 < \end{align}
1821 > \subsubsection{\label{introSection:resistanceTensorRegular}The Resistance Tensor for Regular Shape}
1822  
1823 < \begin{equation}
1824 < \xi (t) = \left\langle {R(t)R(0)} \right\rangle
1825 < \label{introEquation:secondFluctuationDissipation}
1826 < \end{equation}
1823 > For a spherical particle, the translational and rotational friction
1824 > constant can be calculated from Stoke's law,
1825 > \[
1826 > \Xi ^{tt}  = \left( {\begin{array}{*{20}c}
1827 >   {6\pi \eta R} & 0 & 0  \\
1828 >   0 & {6\pi \eta R} & 0  \\
1829 >   0 & 0 & {6\pi \eta R}  \\
1830 > \end{array}} \right)
1831 > \]
1832 > and
1833 > \[
1834 > \Xi ^{rr}  = \left( {\begin{array}{*{20}c}
1835 >   {8\pi \eta R^3 } & 0 & 0  \\
1836 >   0 & {8\pi \eta R^3 } & 0  \\
1837 >   0 & 0 & {8\pi \eta R^3 }  \\
1838 > \end{array}} \right)
1839 > \]
1840 > where $\eta$ is the viscosity of the solvent and $R$ is the
1841 > hydrodynamics radius.
1842  
1843 < \section{\label{introSection:hydroynamics}Hydrodynamics}
1843 > Other non-spherical shape, such as cylinder and ellipsoid
1844 > \textit{etc}, are widely used as reference for developing new
1845 > hydrodynamics theory, because their properties can be calculated
1846 > exactly. In 1936, Perrin extended Stokes's law to general ellipsoid,
1847 > also called a triaxial ellipsoid, which is given in Cartesian
1848 > coordinates by
1849 > \[
1850 > \frac{{x^2 }}{{a^2 }} + \frac{{y^2 }}{{b^2 }} + \frac{{z^2 }}{{c^2
1851 > }} = 1
1852 > \]
1853 > where the semi-axes are of lengths $a$, $b$, and $c$. Unfortunately,
1854 > due to the complexity of the elliptic integral, only the ellipsoid
1855 > with the restriction of two axes having to be equal, \textit{i.e.}
1856 > prolate($ a \ge b = c$) and oblate ($ a < b = c $), can be solved
1857 > exactly. Introducing an elliptic integral parameter $S$ for prolate,
1858 > \[
1859 > S = \frac{2}{{\sqrt {a^2  - b^2 } }}\ln \frac{{a + \sqrt {a^2  - b^2
1860 > } }}{b},
1861 > \]
1862 > and oblate,
1863 > \[
1864 > S = \frac{2}{{\sqrt {b^2  - a^2 } }}arctg\frac{{\sqrt {b^2  - a^2 }
1865 > }}{a}
1866 > \],
1867 > one can write down the translational and rotational resistance
1868 > tensors
1869 > \[
1870 > \begin{array}{l}
1871 > \Xi _a^{tt}  = 16\pi \eta \frac{{a^2  - b^2 }}{{(2a^2  - b^2 )S - 2a}} \\
1872 > \Xi _b^{tt}  = \Xi _c^{tt}  = 32\pi \eta \frac{{a^2  - b^2 }}{{(2a^2  - 3b^2 )S + 2a}} \\
1873 > \end{array},
1874 > \]
1875 > and
1876 > \[
1877 > \begin{array}{l}
1878 > \Xi _a^{rr}  = \frac{{32\pi }}{3}\eta \frac{{(a^2  - b^2 )b^2 }}{{2a - b^2 S}} \\
1879 > \Xi _b^{rr}  = \Xi _c^{rr}  = \frac{{32\pi }}{3}\eta \frac{{(a^4  - b^4 )}}{{(2a^2  - b^2 )S - 2a}} \\
1880 > \end{array}.
1881 > \]
1882  
1883 < \subsection{\label{introSection:frictionTensor} Friction Tensor}
1204 < \subsection{\label{introSection:analyticalApproach}Analytical
1205 < Approach}
1883 > \subsubsection{\label{introSection:resistanceTensorRegularArbitrary}The Resistance Tensor for Arbitrary Shape}
1884  
1885 < \subsection{\label{introSection:approximationApproach}Approximation
1886 < Approach}
1885 > Unlike spherical and other regular shaped molecules, there is not
1886 > analytical solution for friction tensor of any arbitrary shaped
1887 > rigid molecules. The ellipsoid of revolution model and general
1888 > triaxial ellipsoid model have been used to approximate the
1889 > hydrodynamic properties of rigid bodies. However, since the mapping
1890 > from all possible ellipsoidal space, $r$-space, to all possible
1891 > combination of rotational diffusion coefficients, $D$-space is not
1892 > unique\cite{Wegener79} as well as the intrinsic coupling between
1893 > translational and rotational motion of rigid body\cite{}, general
1894 > ellipsoid is not always suitable for modeling arbitrarily shaped
1895 > rigid molecule. A number of studies have been devoted to determine
1896 > the friction tensor for irregularly shaped rigid bodies using more
1897 > advanced method\cite{} where the molecule of interest was modeled by
1898 > combinations of spheres(beads)\cite{} and the hydrodynamics
1899 > properties of the molecule can be calculated using the hydrodynamic
1900 > interaction tensor. Let us consider a rigid assembly of $N$ beads
1901 > immersed in a continuous medium. Due to hydrodynamics interaction,
1902 > the ``net'' velocity of $i$th bead, $v'_i$ is different than its
1903 > unperturbed velocity $v_i$,
1904 > \[
1905 > v'_i  = v_i  - \sum\limits_{j \ne i} {T_{ij} F_j }
1906 > \]
1907 > where $F_i$ is the frictional force, and $T_{ij}$ is the
1908 > hydrodynamic interaction tensor. The friction force of $i$th bead is
1909 > proportional to its ``net'' velocity
1910 > \begin{equation}
1911 > F_i  = \zeta _i v_i  - \zeta _i \sum\limits_{j \ne i} {T_{ij} F_j }.
1912 > \label{introEquation:tensorExpression}
1913 > \end{equation}
1914 > This equation is the basis for deriving the hydrodynamic tensor. In
1915 > 1930, Oseen and Burgers gave a simple solution to Equation
1916 > \ref{introEquation:tensorExpression}
1917 > \begin{equation}
1918 > T_{ij}  = \frac{1}{{8\pi \eta r_{ij} }}\left( {I + \frac{{R_{ij}
1919 > R_{ij}^T }}{{R_{ij}^2 }}} \right).
1920 > \label{introEquation:oseenTensor}
1921 > \end{equation}
1922 > Here $R_{ij}$ is the distance vector between bead $i$ and bead $j$.
1923 > A second order expression for element of different size was
1924 > introduced by Rotne and Prager\cite{} and improved by Garc\'{i}a de
1925 > la Torre and Bloomfield,
1926 > \begin{equation}
1927 > T_{ij}  = \frac{1}{{8\pi \eta R_{ij} }}\left[ {\left( {I +
1928 > \frac{{R_{ij} R_{ij}^T }}{{R_{ij}^2 }}} \right) + R\frac{{\sigma
1929 > _i^2  + \sigma _j^2 }}{{r_{ij}^2 }}\left( {\frac{I}{3} -
1930 > \frac{{R_{ij} R_{ij}^T }}{{R_{ij}^2 }}} \right)} \right].
1931 > \label{introEquation:RPTensorNonOverlapped}
1932 > \end{equation}
1933 > Both of the Equation \ref{introEquation:oseenTensor} and Equation
1934 > \ref{introEquation:RPTensorNonOverlapped} have an assumption $R_{ij}
1935 > \ge \sigma _i  + \sigma _j$. An alternative expression for
1936 > overlapping beads with the same radius, $\sigma$, is given by
1937 > \begin{equation}
1938 > T_{ij}  = \frac{1}{{6\pi \eta R_{ij} }}\left[ {\left( {1 -
1939 > \frac{2}{{32}}\frac{{R_{ij} }}{\sigma }} \right)I +
1940 > \frac{2}{{32}}\frac{{R_{ij} R_{ij}^T }}{{R_{ij} \sigma }}} \right]
1941 > \label{introEquation:RPTensorOverlapped}
1942 > \end{equation}
1943  
1944 < \subsection{\label{introSection:centersRigidBody}Centers of Rigid
1945 < Body}
1944 > To calculate the resistance tensor at an arbitrary origin $O$, we
1945 > construct a $3N \times 3N$ matrix consisting of $N \times N$
1946 > $B_{ij}$ blocks
1947 > \begin{equation}
1948 > B = \left( {\begin{array}{*{20}c}
1949 >   {B_{11} } &  \ldots  & {B_{1N} }  \\
1950 >    \vdots  &  \ddots  &  \vdots   \\
1951 >   {B_{N1} } &  \cdots  & {B_{NN} }  \\
1952 > \end{array}} \right),
1953 > \end{equation}
1954 > where $B_{ij}$ is given by
1955 > \[
1956 > B_{ij}  = \delta _{ij} \frac{I}{{6\pi \eta R}} + (1 - \delta _{ij}
1957 > )T_{ij}
1958 > \]
1959 > where $\delta _{ij}$ is Kronecker delta function. Inverting matrix
1960 > $B$, we obtain
1961  
1962 < \section{\label{introSection:correlationFunctions}Correlation Functions}
1962 > \[
1963 > C = B^{ - 1}  = \left( {\begin{array}{*{20}c}
1964 >   {C_{11} } &  \ldots  & {C_{1N} }  \\
1965 >    \vdots  &  \ddots  &  \vdots   \\
1966 >   {C_{N1} } &  \cdots  & {C_{NN} }  \\
1967 > \end{array}} \right)
1968 > \]
1969 > , which can be partitioned into $N \times N$ $3 \times 3$ block
1970 > $C_{ij}$. With the help of $C_{ij}$ and skew matrix $U_i$
1971 > \[
1972 > U_i  = \left( {\begin{array}{*{20}c}
1973 >   0 & { - z_i } & {y_i }  \\
1974 >   {z_i } & 0 & { - x_i }  \\
1975 >   { - y_i } & {x_i } & 0  \\
1976 > \end{array}} \right)
1977 > \]
1978 > where $x_i$, $y_i$, $z_i$ are the components of the vector joining
1979 > bead $i$ and origin $O$. Hence, the elements of resistance tensor at
1980 > arbitrary origin $O$ can be written as
1981 > \begin{equation}
1982 > \begin{array}{l}
1983 > \Xi _{}^{tt}  = \sum\limits_i {\sum\limits_j {C_{ij} } } , \\
1984 > \Xi _{}^{tr}  = \Xi _{}^{rt}  = \sum\limits_i {\sum\limits_j {U_i C_{ij} } } , \\
1985 > \Xi _{}^{rr}  =  - \sum\limits_i {\sum\limits_j {U_i C_{ij} } } U_j  \\
1986 > \end{array}
1987 > \label{introEquation:ResistanceTensorArbitraryOrigin}
1988 > \end{equation}
1989 >
1990 > The resistance tensor depends on the origin to which they refer. The
1991 > proper location for applying friction force is the center of
1992 > resistance (reaction), at which the trace of rotational resistance
1993 > tensor, $ \Xi ^{rr}$ reaches minimum. Mathematically, the center of
1994 > resistance is defined as an unique point of the rigid body at which
1995 > the translation-rotation coupling tensor are symmetric,
1996 > \begin{equation}
1997 > \Xi^{tr}  = \left( {\Xi^{tr} } \right)^T
1998 > \label{introEquation:definitionCR}
1999 > \end{equation}
2000 > Form Equation \ref{introEquation:ResistanceTensorArbitraryOrigin},
2001 > we can easily find out that the translational resistance tensor is
2002 > origin independent, while the rotational resistance tensor and
2003 > translation-rotation coupling resistance tensor depend on the
2004 > origin. Given resistance tensor at an arbitrary origin $O$, and a
2005 > vector ,$r_{OP}(x_{OP}, y_{OP}, z_{OP})$, from $O$ to $P$, we can
2006 > obtain the resistance tensor at $P$ by
2007 > \begin{equation}
2008 > \begin{array}{l}
2009 > \Xi _P^{tt}  = \Xi _O^{tt}  \\
2010 > \Xi _P^{tr}  = \Xi _P^{rt}  = \Xi _O^{tr}  - U_{OP} \Xi _O^{tt}  \\
2011 > \Xi _P^{rr}  = \Xi _O^{rr}  - U_{OP} \Xi _O^{tt} U_{OP}  + \Xi _O^{tr} U_{OP}  - U_{OP} \Xi _O^{tr} ^{^T }  \\
2012 > \end{array}
2013 > \label{introEquation:resistanceTensorTransformation}
2014 > \end{equation}
2015 > where
2016 > \[
2017 > U_{OP}  = \left( {\begin{array}{*{20}c}
2018 >   0 & { - z_{OP} } & {y_{OP} }  \\
2019 >   {z_i } & 0 & { - x_{OP} }  \\
2020 >   { - y_{OP} } & {x_{OP} } & 0  \\
2021 > \end{array}} \right)
2022 > \]
2023 > Using Equations \ref{introEquation:definitionCR} and
2024 > \ref{introEquation:resistanceTensorTransformation}, one can locate
2025 > the position of center of resistance,
2026 > \[
2027 > \left( \begin{array}{l}
2028 > x_{OR}  \\
2029 > y_{OR}  \\
2030 > z_{OR}  \\
2031 > \end{array} \right) = \left( {\begin{array}{*{20}c}
2032 >   {(\Xi _O^{rr} )_{yy}  + (\Xi _O^{rr} )_{zz} } & { - (\Xi _O^{rr} )_{xy} } & { - (\Xi _O^{rr} )_{xz} }  \\
2033 >   { - (\Xi _O^{rr} )_{xy} } & {(\Xi _O^{rr} )_{zz}  + (\Xi _O^{rr} )_{xx} } & { - (\Xi _O^{rr} )_{yz} }  \\
2034 >   { - (\Xi _O^{rr} )_{xz} } & { - (\Xi _O^{rr} )_{yz} } & {(\Xi _O^{rr} )_{xx}  + (\Xi _O^{rr} )_{yy} }  \\
2035 > \end{array}} \right)^{ - 1} \left( \begin{array}{l}
2036 > (\Xi _O^{tr} )_{yz}  - (\Xi _O^{tr} )_{zy}  \\
2037 > (\Xi _O^{tr} )_{zx}  - (\Xi _O^{tr} )_{xz}  \\
2038 > (\Xi _O^{tr} )_{xy}  - (\Xi _O^{tr} )_{yx}  \\
2039 > \end{array} \right).
2040 > \]
2041 > where $x_OR$, $y_OR$, $z_OR$ are the components of the vector
2042 > joining center of resistance $R$ and origin $O$.

Diff Legend

Removed lines
+ Added lines
< Changed lines
> Changed lines