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

Comparing trunk/tengDissertation/Methodology.tex (file contents):
Revision 2849 by tim, Fri Jun 9 02:41:58 2006 UTC vs.
Revision 2850 by tim, Sun Jun 11 01:55:48 2006 UTC

# Line 16 | Line 16 | two decades. Matubayasi and Nakahara developed a time-
16  
17   Integration schemes for rotational motion of the rigid molecules in
18   microcanonical ensemble have been extensively studied in the last
19 < two decades. Matubayasi and Nakahara developed a time-reversible
20 < integrator for rigid bodies in quaternion representation. Although
21 < it is not symplectic, this integrator still demonstrates a better
22 < long-time energy conservation than traditional methods because of
23 < the time-reversible nature. Extending Trotter-Suzuki to general
24 < system with a flat phase space, Miller and his colleagues devised an
25 < novel symplectic, time-reversible and volume-preserving integrator
26 < in quaternion representation, which was shown to be superior to the
27 < time-reversible integrator of Matubayasi and Nakahara. However, all
28 < of the integrators in quaternion representation suffer from the
19 > two decades. Matubayasi developed a time-reversible integrator for
20 > rigid bodies in quaternion representation. Although it is not
21 > symplectic, this integrator still demonstrates a better long-time
22 > energy conservation than traditional methods because of the
23 > time-reversible nature. Extending Trotter-Suzuki to general system
24 > with a flat phase space, Miller and his colleagues devised an novel
25 > symplectic, time-reversible and volume-preserving integrator in
26 > quaternion representation, which was shown to be superior to the
27 > Matubayasi's time-reversible integrator. However, all of the
28 > integrators in quaternion representation suffer from the
29   computational penalty of constructing a rotation matrix from
30   quaternions to evolve coordinates and velocities at every time step.
31   An alternative integration scheme utilizing rotation matrix directly
# Line 810 | Line 810 | resonance artifact more efficiently\cite{Sandu1999}.
810   bath, Langevin dynamics has been shown to be able to damp out the
811   resonance artifact more efficiently\cite{Sandu1999}.
812  
813 %review rigid body dynamics
814 Rigid bodies are frequently involved in the modeling of different
815 areas, from engineering, physics, to chemistry. For example,
816 missiles and vehicle are usually modeled by rigid bodies.  The
817 movement of the objects in 3D gaming engine or other physics
818 simulator is governed by the rigid body dynamics. In molecular
819 simulation, rigid body is used to simplify the model in
820 protein-protein docking study\cite{Gray2003}.
821
813   It is very important to develop stable and efficient methods to
814   integrate the equations of motion of orientational degrees of
815   freedom. Euler angles are the nature choice to describe the
# Line 842 | Line 833 | geometric structure of the flow. Matubayasi and Nakaha
833  
834   The break through in geometric literature suggests that, in order to
835   develop a long-term integration scheme, one should preserve the
836 < geometric structure of the flow. Matubayasi and Nakahara developed a
836 > geometric structure of the flow. Matubayasi developed a
837   time-reversible integrator for rigid bodies in quaternion
838   representation. Although it is not symplectic, this integrator still
839   demonstrates a better long-time energy conservation than traditional
# Line 915 | Line 906 | sophisticated rigid body dynamics.
906   estimation of friction tensor from hydrodynamics theory into the
907   sophisticated rigid body dynamics.
908  
909 <
910 < \subsection{Friction Tensor}
911 <
912 < For an arbitrary rigid body moves in a fluid, it may experience
913 < friction force $f_r$ or friction torque $\tau _r$ along the opposite
914 < direction of the velocity $v$ or angular velocity $\omega$ at
915 < arbitrary origin $P$,
916 < \begin{equation}
917 < \left( \begin{array}{l}
918 < f_r  \\
919 < \tau _r  \\
920 < \end{array} \right) =  - \left( {\begin{array}{*{20}c}
921 <   {\Xi _{P,t} } & {\Xi _{P,c}^T }  \\
922 <   {\Xi _{P,c} } & {\Xi _{P,r} }  \\
923 < \end{array}} \right)\left( \begin{array}{l}
924 < \nu  \\
925 < \omega  \\
909 > \subsection{\label{introSection:frictionTensor} Friction Tensor}
910 > Theoretically, the friction kernel can be determined using velocity
911 > autocorrelation function. However, this approach become impractical
912 > when the system become more and more complicate. Instead, various
913 > approaches based on hydrodynamics have been developed to calculate
914 > the friction coefficients. The friction effect is isotropic in
915 > Equation, $\zeta$ can be taken as a scalar. In general, friction
916 > tensor $\Xi$ is a $6\times 6$ matrix given by
917 > \[
918 > \Xi  = \left( {\begin{array}{*{20}c}
919 >   {\Xi _{}^{tt} } & {\Xi _{}^{rt} }  \\
920 >   {\Xi _{}^{tr} } & {\Xi _{}^{rr} }  \\
921 > \end{array}} \right).
922 > \]
923 > Here, $ {\Xi^{tt} }$ and $ {\Xi^{rr} }$ are translational friction
924 > tensor and rotational resistance (friction) tensor respectively,
925 > while ${\Xi^{tr} }$ is translation-rotation coupling tensor and $
926 > {\Xi^{rt} }$ is rotation-translation coupling tensor. When a
927 > particle moves in a fluid, it may experience friction force or
928 > torque along the opposite direction of the velocity or angular
929 > velocity,
930 > \[
931 > \left( \begin{array}{l}
932 > F_R  \\
933 > \tau _R  \\
934 > \end{array} \right) =  - \left( {\begin{array}{*{20}c}
935 >   {\Xi ^{tt} } & {\Xi ^{rt} }  \\
936 >   {\Xi ^{tr} } & {\Xi ^{rr} }  \\
937 > \end{array}} \right)\left( \begin{array}{l}
938 > v \\
939 > w \\
940   \end{array} \right)
941 < \end{equation}
942 < where $\Xi _{P,t}t$ is the translation friction tensor, $\Xi _{P,r}$
943 < is the rotational friction tensor and $\Xi _{P,c}$ is the
939 < translation-rotation coupling tensor. The procedure of calculating
940 < friction tensor using hydrodynamic tensor and comparison between
941 < bead model and shell model were elaborated by Carrasco \textit{et
942 < al}\cite{Carrasco1999}. An important property of the friction tensor
943 < is that the translational friction tensor is independent of origin
944 < while the rotational and coupling are sensitive to the choice of the
945 < origin \cite{Brenner1967}, which can be described by
946 < \begin{equation}
947 < \begin{array}{c}
948 < \Xi _{P,t}  = \Xi _{O,t}  = \Xi _t  \\
949 < \Xi _{P,c}  = \Xi _{O,c}  - r_{OP}  \times \Xi _t  \\
950 < \Xi _{P,r}  = \Xi _{O,r}  - r_{OP}  \times \Xi _t  \times r_{OP}  + \Xi _{O,c}  \times r_{OP}  - r_{OP}  \times \Xi _{O,c}^T  \\
951 < \end{array}
952 < \end{equation}
953 < Where $O$ is another origin and $r_{OP}$ is the vector joining $O$
954 < and $P$. It is also worthy of mention that both of translational and
955 < rotational frictional tensors are always symmetric. In contrast,
956 < coupling tensor is only symmetric at center of reaction:
957 < \begin{equation}
958 < \Xi _{R,c}  = \Xi _{R,c}^T
959 < \end{equation}
960 < The proper location for applying friction force is the center of
961 < reaction, at which the trace of rotational resistance tensor reaches
962 < minimum.
941 > \]
942 > where $F_r$ is the friction force and $\tau _R$ is the friction
943 > toque.
944  
945 < \subsection{Rigid body dynamics}
945 > \subsubsection{\label{introSection:resistanceTensorRegular}\textbf{The Resistance Tensor for Regular Shape}}
946  
947 < The Hamiltonian of rigid body can be separated in terms of potential
948 < energy $V(r,A)$ and kinetic energy $T(p,\pi)$,
947 > For a spherical particle, the translational and rotational friction
948 > constant can be calculated from Stoke's law,
949   \[
950 < H = V(r,A) + T(v,\pi )
951 < \]
952 < A second-order symplectic method is now obtained by the composition
953 < of the flow maps,
950 > \Xi ^{tt}  = \left( {\begin{array}{*{20}c}
951 >   {6\pi \eta R} & 0 & 0  \\
952 >   0 & {6\pi \eta R} & 0  \\
953 >   0 & 0 & {6\pi \eta R}  \\
954 > \end{array}} \right)
955 > \]
956 > and
957   \[
958 < \varphi _{\Delta t}  = \varphi _{\Delta t/2,V}  \circ \varphi
959 < _{\Delta t,T}  \circ \varphi _{\Delta t/2,V}.
958 > \Xi ^{rr}  = \left( {\begin{array}{*{20}c}
959 >   {8\pi \eta R^3 } & 0 & 0  \\
960 >   0 & {8\pi \eta R^3 } & 0  \\
961 >   0 & 0 & {8\pi \eta R^3 }  \\
962 > \end{array}} \right)
963   \]
964 < Moreover, $\varphi _{\Delta t/2,V}$ can be divided into two
965 < sub-flows which corresponding to force and torque respectively,
964 > where $\eta$ is the viscosity of the solvent and $R$ is the
965 > hydrodynamics radius.
966 >
967 > Other non-spherical shape, such as cylinder and ellipsoid
968 > \textit{etc}, are widely used as reference for developing new
969 > hydrodynamics theory, because their properties can be calculated
970 > exactly. In 1936, Perrin extended Stokes's law to general ellipsoid,
971 > also called a triaxial ellipsoid, which is given in Cartesian
972 > coordinates by\cite{Perrin1934, Perrin1936}
973   \[
974 < \varphi _{\Delta t/2,V}  = \varphi _{\Delta t/2,F}  \circ \varphi
975 < _{\Delta t/2,\tau }.
974 > \frac{{x^2 }}{{a^2 }} + \frac{{y^2 }}{{b^2 }} + \frac{{z^2 }}{{c^2
975 > }} = 1
976   \]
977 < Since the associated operators of $\varphi _{\Delta t/2,F} $ and
978 < $\circ \varphi _{\Delta t/2,\tau }$ are commuted, the composition
979 < order inside $\varphi _{\Delta t/2,V}$ does not matter.
977 > where the semi-axes are of lengths $a$, $b$, and $c$. Unfortunately,
978 > due to the complexity of the elliptic integral, only the ellipsoid
979 > with the restriction of two axes having to be equal, \textit{i.e.}
980 > prolate($ a \ge b = c$) and oblate ($ a < b = c $), can be solved
981 > exactly. Introducing an elliptic integral parameter $S$ for prolate,
982 > \[
983 > S = \frac{2}{{\sqrt {a^2  - b^2 } }}\ln \frac{{a + \sqrt {a^2  - b^2
984 > } }}{b},
985 > \]
986 > and oblate,
987 > \[
988 > S = \frac{2}{{\sqrt {b^2  - a^2 } }}arctg\frac{{\sqrt {b^2  - a^2 }
989 > }}{a}
990 > \],
991 > one can write down the translational and rotational resistance
992 > tensors
993 > \[
994 > \begin{array}{l}
995 > \Xi _a^{tt}  = 16\pi \eta \frac{{a^2  - b^2 }}{{(2a^2  - b^2 )S - 2a}} \\
996 > \Xi _b^{tt}  = \Xi _c^{tt}  = 32\pi \eta \frac{{a^2  - b^2 }}{{(2a^2  - 3b^2 )S + 2a}} \\
997 > \end{array},
998 > \]
999 > and
1000 > \[
1001 > \begin{array}{l}
1002 > \Xi _a^{rr}  = \frac{{32\pi }}{3}\eta \frac{{(a^2  - b^2 )b^2 }}{{2a - b^2 S}} \\
1003 > \Xi _b^{rr}  = \Xi _c^{rr}  = \frac{{32\pi }}{3}\eta \frac{{(a^4  - b^4 )}}{{(2a^2  - b^2 )S - 2a}} \\
1004 > \end{array}.
1005 > \]
1006  
1007 < Furthermore, kinetic potential can be separated to translational
1008 < kinetic term, $T^t (p)$, and rotational kinetic term, $T^r (\pi )$,
1009 < \begin{equation}
1010 < T(p,\pi ) =T^t (p) + T^r (\pi ).
1011 < \end{equation}
1012 < where $ T^t (p) = \frac{1}{2}v^T m v $ and $T^r (\pi )$ is defined
1013 < by \ref{introEquation:rotationalKineticRB}. Therefore, the
1014 < corresponding flow maps are given by
1007 > \subsubsection{\label{introSection:resistanceTensorRegularArbitrary}\textbf{The Resistance Tensor for Arbitrary Shape}}
1008 >
1009 > Unlike spherical and other regular shaped molecules, there is not
1010 > analytical solution for friction tensor of any arbitrary shaped
1011 > rigid molecules. The ellipsoid of revolution model and general
1012 > triaxial ellipsoid model have been used to approximate the
1013 > hydrodynamic properties of rigid bodies. However, since the mapping
1014 > from all possible ellipsoidal space, $r$-space, to all possible
1015 > combination of rotational diffusion coefficients, $D$-space is not
1016 > unique\cite{Wegener1979} as well as the intrinsic coupling between
1017 > translational and rotational motion of rigid body, general ellipsoid
1018 > is not always suitable for modeling arbitrarily shaped rigid
1019 > molecule. A number of studies have been devoted to determine the
1020 > friction tensor for irregularly shaped rigid bodies using more
1021 > advanced method where the molecule of interest was modeled by
1022 > combinations of spheres(beads)\cite{Carrasco1999} and the
1023 > hydrodynamics properties of the molecule can be calculated using the
1024 > hydrodynamic interaction tensor. Let us consider a rigid assembly of
1025 > $N$ beads immersed in a continuous medium. Due to hydrodynamics
1026 > interaction, the ``net'' velocity of $i$th bead, $v'_i$ is different
1027 > than its unperturbed velocity $v_i$,
1028   \[
1029 < \varphi _{\Delta t,T}  = \varphi _{\Delta t,T^t }  \circ \varphi
997 < _{\Delta t,T^r }.
1029 > v'_i  = v_i  - \sum\limits_{j \ne i} {T_{ij} F_j }
1030   \]
1031 < The free rigid body is an example of Lie-Poisson system with
1032 < Hamiltonian function
1031 > where $F_i$ is the frictional force, and $T_{ij}$ is the
1032 > hydrodynamic interaction tensor. The friction force of $i$th bead is
1033 > proportional to its ``net'' velocity
1034   \begin{equation}
1035 < T^r (\pi ) = T_1 ^r (\pi _1 ) + T_2^r (\pi _2 ) + T_3^r (\pi _3 )
1036 < \label{introEquation:rotationalKineticRB}
1035 > F_i  = \zeta _i v_i  - \zeta _i \sum\limits_{j \ne i} {T_{ij} F_j }.
1036 > \label{introEquation:tensorExpression}
1037   \end{equation}
1038 < where $T_i^r (\pi _i ) = \frac{{\pi _i ^2 }}{{2I_i }}$ and
1039 < Lie-Poisson structure matrix,
1038 > This equation is the basis for deriving the hydrodynamic tensor. In
1039 > 1930, Oseen and Burgers gave a simple solution to Equation
1040 > \ref{introEquation:tensorExpression}
1041   \begin{equation}
1042 < J(\pi ) = \left( {\begin{array}{*{20}c}
1043 <   0 & {\pi _3 } & { - \pi _2 }  \\
1010 <   { - \pi _3 } & 0 & {\pi _1 }  \\
1011 <   {\pi _2 } & { - \pi _1 } & 0  \\
1012 < \end{array}} \right)
1042 > T_{ij}  = \frac{1}{{8\pi \eta r_{ij} }}\left( {I + \frac{{R_{ij}
1043 > R_{ij}^T }}{{R_{ij}^2 }}} \right). \label{introEquation:oseenTensor}
1044   \end{equation}
1045 < Thus, the dynamics of free rigid body is governed by
1045 > Here $R_{ij}$ is the distance vector between bead $i$ and bead $j$.
1046 > A second order expression for element of different size was
1047 > introduced by Rotne and Prager\cite{Rotne1969} and improved by
1048 > Garc\'{i}a de la Torre and Bloomfield\cite{Torre1977},
1049   \begin{equation}
1050 < \frac{d}{{dt}}\pi  = J(\pi )\nabla _\pi  T^r (\pi )
1050 > T_{ij}  = \frac{1}{{8\pi \eta R_{ij} }}\left[ {\left( {I +
1051 > \frac{{R_{ij} R_{ij}^T }}{{R_{ij}^2 }}} \right) + R\frac{{\sigma
1052 > _i^2  + \sigma _j^2 }}{{r_{ij}^2 }}\left( {\frac{I}{3} -
1053 > \frac{{R_{ij} R_{ij}^T }}{{R_{ij}^2 }}} \right)} \right].
1054 > \label{introEquation:RPTensorNonOverlapped}
1055   \end{equation}
1056 < One may notice that each $T_i^r$ in Equation
1057 < \ref{introEquation:rotationalKineticRB} can be solved exactly. For
1058 < instance, the equations of motion due to $T_1^r$ are given by
1056 > Both of the Equation \ref{introEquation:oseenTensor} and Equation
1057 > \ref{introEquation:RPTensorNonOverlapped} have an assumption $R_{ij}
1058 > \ge \sigma _i  + \sigma _j$. An alternative expression for
1059 > overlapping beads with the same radius, $\sigma$, is given by
1060   \begin{equation}
1061 < \frac{d}{{dt}}\pi  = R_1 \pi ,\frac{d}{{dt}}A = AR_1
1062 < \label{introEqaution:RBMotionSingleTerm}
1061 > T_{ij}  = \frac{1}{{6\pi \eta R_{ij} }}\left[ {\left( {1 -
1062 > \frac{2}{{32}}\frac{{R_{ij} }}{\sigma }} \right)I +
1063 > \frac{2}{{32}}\frac{{R_{ij} R_{ij}^T }}{{R_{ij} \sigma }}} \right]
1064 > \label{introEquation:RPTensorOverlapped}
1065   \end{equation}
1066 < where
1067 < \[ R_1  = \left( {\begin{array}{*{20}c}
1068 <   0 & 0 & 0  \\
1069 <   0 & 0 & {\pi _1 }  \\
1070 <   0 & { - \pi _1 } & 0  \\
1071 < \end{array}} \right).
1072 < \]
1073 < The solutions of Equation \ref{introEqaution:RBMotionSingleTerm} is
1066 >
1067 > To calculate the resistance tensor at an arbitrary origin $O$, we
1068 > construct a $3N \times 3N$ matrix consisting of $N \times N$
1069 > $B_{ij}$ blocks
1070 > \begin{equation}
1071 > B = \left( {\begin{array}{*{20}c}
1072 >   {B_{11} } &  \ldots  & {B_{1N} }  \\
1073 >    \vdots  &  \ddots  &  \vdots   \\
1074 >   {B_{N1} } &  \cdots  & {B_{NN} }  \\
1075 > \end{array}} \right),
1076 > \end{equation}
1077 > where $B_{ij}$ is given by
1078   \[
1079 < \pi (\Delta t) = e^{\Delta tR_1 } \pi (0),A(\Delta t) =
1080 < A(0)e^{\Delta tR_1 }
1079 > B_{ij}  = \delta _{ij} \frac{I}{{6\pi \eta R}} + (1 - \delta _{ij}
1080 > )T_{ij}
1081   \]
1082 < with
1082 > where $\delta _{ij}$ is Kronecker delta function. Inverting matrix
1083 > $B$, we obtain
1084 >
1085   \[
1086 < e^{\Delta tR_1 }  = \left( {\begin{array}{*{20}c}
1087 <   0 & 0 & 0  \\
1088 <   0 & {\cos \theta _1 } & {\sin \theta _1 }  \\
1089 <   0 & { - \sin \theta _1 } & {\cos \theta _1 }  \\
1090 < \end{array}} \right),\theta _1  = \frac{{\pi _1 }}{{I_1 }}\Delta t.
1086 > C = B^{ - 1}  = \left( {\begin{array}{*{20}c}
1087 >   {C_{11} } &  \ldots  & {C_{1N} }  \\
1088 >    \vdots  &  \ddots  &  \vdots   \\
1089 >   {C_{N1} } &  \cdots  & {C_{NN} }  \\
1090 > \end{array}} \right)
1091   \]
1092 < To reduce the cost of computing expensive functions in $e^{\Delta
1093 < tR_1 }$, we can use Cayley transformation,
1092 > , which can be partitioned into $N \times N$ $3 \times 3$ block
1093 > $C_{ij}$. With the help of $C_{ij}$ and skew matrix $U_i$
1094   \[
1095 < e^{\Delta tR_1 }  \approx (1 - \Delta tR_1 )^{ - 1} (1 + \Delta tR_1
1096 < )
1095 > U_i  = \left( {\begin{array}{*{20}c}
1096 >   0 & { - z_i } & {y_i }  \\
1097 >   {z_i } & 0 & { - x_i }  \\
1098 >   { - y_i } & {x_i } & 0  \\
1099 > \end{array}} \right)
1100   \]
1101 < The flow maps for $T_2^r$ and $T_3^r$ can be found in the same
1102 < manner.
1101 > where $x_i$, $y_i$, $z_i$ are the components of the vector joining
1102 > bead $i$ and origin $O$. Hence, the elements of resistance tensor at
1103 > arbitrary origin $O$ can be written as
1104 > \begin{equation}
1105 > \begin{array}{l}
1106 > \Xi _{}^{tt}  = \sum\limits_i {\sum\limits_j {C_{ij} } } , \\
1107 > \Xi _{}^{tr}  = \Xi _{}^{rt}  = \sum\limits_i {\sum\limits_j {U_i C_{ij} } } , \\
1108 > \Xi _{}^{rr}  =  - \sum\limits_i {\sum\limits_j {U_i C_{ij} } } U_j  \\
1109 > \end{array}
1110 > \label{introEquation:ResistanceTensorArbitraryOrigin}
1111 > \end{equation}
1112  
1113 < In order to construct a second-order symplectic method, we split the
1114 < angular kinetic Hamiltonian function into five terms
1113 > The resistance tensor depends on the origin to which they refer. The
1114 > proper location for applying friction force is the center of
1115 > resistance (reaction), at which the trace of rotational resistance
1116 > tensor, $ \Xi ^{rr}$ reaches minimum. Mathematically, the center of
1117 > resistance is defined as an unique point of the rigid body at which
1118 > the translation-rotation coupling tensor are symmetric,
1119 > \begin{equation}
1120 > \Xi^{tr}  = \left( {\Xi^{tr} } \right)^T
1121 > \label{introEquation:definitionCR}
1122 > \end{equation}
1123 > Form Equation \ref{introEquation:ResistanceTensorArbitraryOrigin},
1124 > we can easily find out that the translational resistance tensor is
1125 > origin independent, while the rotational resistance tensor and
1126 > translation-rotation coupling resistance tensor depend on the
1127 > origin. Given resistance tensor at an arbitrary origin $O$, and a
1128 > vector ,$r_{OP}(x_{OP}, y_{OP}, z_{OP})$, from $O$ to $P$, we can
1129 > obtain the resistance tensor at $P$ by
1130 > \begin{equation}
1131 > \begin{array}{l}
1132 > \Xi _P^{tt}  = \Xi _O^{tt}  \\
1133 > \Xi _P^{tr}  = \Xi _P^{rt}  = \Xi _O^{tr}  - U_{OP} \Xi _O^{tt}  \\
1134 > \Xi _P^{rr}  = \Xi _O^{rr}  - U_{OP} \Xi _O^{tt} U_{OP}  + \Xi _O^{tr} U_{OP}  - U_{OP} \Xi _O^{{tr} ^{^T }}  \\
1135 > \end{array}
1136 > \label{introEquation:resistanceTensorTransformation}
1137 > \end{equation}
1138 > where
1139   \[
1140 < T^r (\pi ) = \frac{1}{2}T_1 ^r (\pi _1 ) + \frac{1}{2}T_2^r (\pi _2
1141 < ) + T_3^r (\pi _3 ) + \frac{1}{2}T_2^r (\pi _2 ) + \frac{1}{2}T_1 ^r
1142 < (\pi _1 )
1143 < \].
1144 < Concatenating flows corresponding to these five terms, we can obtain
1062 < the flow map for free rigid body,
1063 < \[
1064 < \varphi _{\Delta t,T^r }  = \varphi _{\Delta t/2,\pi _1 }  \circ
1065 < \varphi _{\Delta t/2,\pi _2 }  \circ \varphi _{\Delta t,\pi _3 }
1066 < \circ \varphi _{\Delta t/2,\pi _2 }  \circ \varphi _{\Delta t/2,\pi
1067 < _1 }.
1140 > U_{OP}  = \left( {\begin{array}{*{20}c}
1141 >   0 & { - z_{OP} } & {y_{OP} }  \\
1142 >   {z_i } & 0 & { - x_{OP} }  \\
1143 >   { - y_{OP} } & {x_{OP} } & 0  \\
1144 > \end{array}} \right)
1145   \]
1146 + Using Equations \ref{introEquation:definitionCR} and
1147 + \ref{introEquation:resistanceTensorTransformation}, one can locate
1148 + the position of center of resistance,
1149 + \begin{eqnarray*}
1150 + \left( \begin{array}{l}
1151 + x_{OR}  \\
1152 + y_{OR}  \\
1153 + z_{OR}  \\
1154 + \end{array} \right) & = &\left( {\begin{array}{*{20}c}
1155 +   {(\Xi _O^{rr} )_{yy}  + (\Xi _O^{rr} )_{zz} } & { - (\Xi _O^{rr} )_{xy} } & { - (\Xi _O^{rr} )_{xz} }  \\
1156 +   { - (\Xi _O^{rr} )_{xy} } & {(\Xi _O^{rr} )_{zz}  + (\Xi _O^{rr} )_{xx} } & { - (\Xi _O^{rr} )_{yz} }  \\
1157 +   { - (\Xi _O^{rr} )_{xz} } & { - (\Xi _O^{rr} )_{yz} } & {(\Xi _O^{rr} )_{xx}  + (\Xi _O^{rr} )_{yy} }  \\
1158 + \end{array}} \right)^{ - 1}  \\
1159 +  & & \left( \begin{array}{l}
1160 + (\Xi _O^{tr} )_{yz}  - (\Xi _O^{tr} )_{zy}  \\
1161 + (\Xi _O^{tr} )_{zx}  - (\Xi _O^{tr} )_{xz}  \\
1162 + (\Xi _O^{tr} )_{xy}  - (\Xi _O^{tr} )_{yx}  \\
1163 + \end{array} \right) \\
1164 + \end{eqnarray*}
1165  
1166 < The equations of motion corresponding to potential energy and
1167 < kinetic energy are listed in the below table,
1072 < \begin{center}
1073 < \begin{tabular}{|l|l|}
1074 <  \hline
1075 <  % after \\: \hline or \cline{col1-col2} \cline{col3-col4} ...
1076 <  Potential & Kinetic \\
1077 <  $\frac{{dq}}{{dt}} = \frac{p}{m}$ & $\frac{d}{{dt}}q = p$ \\
1078 <  $\frac{d}{{dt}}p =  - \frac{{\partial V}}{{\partial q}}$ & $ \frac{d}{{dt}}p = 0$ \\
1079 <  $\frac{d}{{dt}}Q = 0$ & $ \frac{d}{{dt}}Q = Qskew(I^{ - 1} j)$ \\
1080 <  $ \frac{d}{{dt}}\pi  = \sum\limits_i {\left( {Q^T F_i (r,Q)} \right) \times X_i }$ & $\frac{d}{{dt}}\pi  = \pi  \times I^{ - 1} \pi$\\
1081 <  \hline
1082 < \end{tabular}
1083 < \end{center}
1166 > where $x_OR$, $y_OR$, $z_OR$ are the components of the vector
1167 > joining center of resistance $R$ and origin $O$.
1168  
1085 Finally, we obtain the overall symplectic flow maps for free moving
1086 rigid body
1087 \begin{align*}
1088 \varphi _{\Delta t}  = &\varphi _{\Delta t/2,F}  \circ \varphi _{\Delta t/2,\tau } \circ  \\
1089  &\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 } \circ  \\
1090  &\varphi _{\Delta t/2,\tau }  \circ \varphi _{\Delta t/2,F}  .\\
1091 \label{introEquation:overallRBFlowMaps}
1092 \end{align*}
1093
1169   \subsection{Langevin dynamics for rigid particles of arbitrary shape}
1170  
1171   Consider a Langevin equation of motions in generalized coordinates

Diff Legend

Removed lines
+ Added lines
< Changed lines
> Changed lines