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 2850 by tim, Sun Jun 11 01:55:48 2006 UTC vs.
Revision 2851 by tim, Sun Jun 11 02:06:01 2006 UTC

# Line 756 | Line 756 | F_{z_{\text{Harmonic}}}(t)=-\frac{\partial U(t)}{\part
756   \begin{equation}
757   F_{z_{\text{Harmonic}}}(t)=-\frac{\partial U(t)}{\partial z(t)}=
758      -k_{\text{Harmonic}}(z(t)-z_{\text{cons}}).
759 \end{equation}
760
761
762 \section{\label{methodSection:langevin}Integrators for Langevin Dynamics of Rigid Bodies}
763
764 %\subsection{\label{methodSection:temperature}Temperature Control}
765
766 %\subsection{\label{methodSection:pressureControl}Pressure Control}
767
768 %\section{\label{methodSection:hydrodynamics}Hydrodynamics}
769
770 %applications of langevin dynamics
771 As an excellent alternative to newtonian dynamics, Langevin
772 dynamics, which mimics a simple heat bath with stochastic and
773 dissipative forces, has been applied in a variety of studies. The
774 stochastic treatment of the solvent enables us to carry out
775 substantially longer time simulation. Implicit solvent Langevin
776 dynamics simulation of met-enkephalin not only outperforms explicit
777 solvent simulation on computation efficiency, but also agrees very
778 well with explicit solvent simulation on dynamics
779 properties\cite{Shen2002}. Recently, applying Langevin dynamics with
780 UNRES model, Liow and his coworkers suggest that protein folding
781 pathways can be possibly exploited within a reasonable amount of
782 time\cite{Liwo2005}. The stochastic nature of the Langevin dynamics
783 also enhances the sampling of the system and increases the
784 probability of crossing energy barrier\cite{Banerjee2004, Cui2003}.
785 Combining Langevin dynamics with Kramers's theory, Klimov and
786 Thirumalai identified the free-energy barrier by studying the
787 viscosity dependence of the protein folding rates\cite{Klimov1997}.
788 In order to account for solvent induced interactions missing from
789 implicit solvent model, Kaya incorporated desolvation free energy
790 barrier into implicit coarse-grained solvent model in protein
791 folding/unfolding study and discovered a higher free energy barrier
792 between the native and denatured states. Because of its stability
793 against noise, Langevin dynamics is very suitable for studying
794 remagnetization processes in various
795 systems\cite{Palacios1998,Berkov2002,Denisov2003}. For instance, the
796 oscillation power spectrum of nanoparticles from Langevin dynamics
797 simulation has the same peak frequencies for different wave
798 vectors,which recovers the property of magnetic excitations in small
799 finite structures\cite{Berkov2005a}. In an attempt to reduce the
800 computational cost of simulation, multiple time stepping (MTS)
801 methods have been introduced and have been of great interest to
802 macromolecule and protein community\cite{Tuckerman1992}. Relying on
803 the observation that forces between distant atoms generally
804 demonstrate slower fluctuations than forces between close atoms, MTS
805 method are generally implemented by evaluating the slowly
806 fluctuating forces less frequently than the fast ones.
807 Unfortunately, nonlinear instability resulting from increasing
808 timestep in MTS simulation have became a critical obstruction
809 preventing the long time simulation. Due to the coupling to the heat
810 bath, Langevin dynamics has been shown to be able to damp out the
811 resonance artifact more efficiently\cite{Sandu1999}.
812
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
816 rotational degrees of freedom. However, due to its singularity, the
817 numerical integration of corresponding equations of motion is very
818 inefficient and inaccurate. Although an alternative integrator using
819 different sets of Euler angles can overcome this
820 difficulty\cite{Ryckaert1977, Andersen1983}, the computational
821 penalty and the lost of angular momentum conservation still remain.
822 In 1977, a singularity free representation utilizing quaternions was
823 developed by Evans\cite{Evans1977}. Unfortunately, this approach
824 suffer from the nonseparable Hamiltonian resulted from quaternion
825 representation, which prevents the symplectic algorithm to be
826 utilized. Another different approach is to apply holonomic
827 constraints to the atoms belonging to the rigid
828 body\cite{Barojas1973}. Each atom moves independently under the
829 normal forces deriving from potential energy and constraint forces
830 which are used to guarantee the rigidness. However, due to their
831 iterative nature, SHAKE and Rattle algorithm converge very slowly
832 when the number of constraint increases.
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 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
840 methods because of the time-reversible nature. Extending
841 Trotter-Suzuki to general system with a flat phase space, Miller and
842 his colleagues devised an novel symplectic, time-reversible and
843 volume-preserving integrator in quaternion representation. However,
844 all of the integrators in quaternion representation suffer from the
845 computational penalty of constructing a rotation matrix from
846 quaternions to evolve coordinates and velocities at every time step.
847 An alternative integration scheme utilizing rotation matrix directly
848 is RSHAKE\cite{Kol1997}, in which a conjugate momentum to rotation
849 matrix is introduced to re-formulate the Hamiltonian's equation and
850 the Hamiltonian is evolved in a constraint manifold by iteratively
851 satisfying the orthogonality constraint. However, RSHAKE is
852 inefficient because of the iterative procedure. An extremely
853 efficient integration scheme in rotation matrix representation,
854 which also preserves the same structural properties of the
855 Hamiltonian flow as Miller's integrator, is proposed by Dullweber,
856 Leimkuhler and McLachlan (DLM)\cite{Dullweber1997}.
857
858 %review langevin/browninan dynamics for arbitrarily shaped rigid body
859 Combining Langevin or Brownian dynamics with rigid body dynamics,
860 one can study the slow processes in biomolecular systems. Modeling
861 the DNA as a chain of rigid spheres beads, which subject to harmonic
862 potentials as well as excluded volume potentials, Mielke and his
863 coworkers discover rapid superhelical stress generations from the
864 stochastic simulation of twin supercoiling DNA with response to
865 induced torques\cite{Mielke2004}. Membrane fusion is another key
866 biological process which controls a variety of physiological
867 functions, such as release of neurotransmitters \textit{etc}. A
868 typical fusion event happens on the time scale of millisecond, which
869 is impracticable to study using all atomistic model with newtonian
870 mechanics. With the help of coarse-grained rigid body model and
871 stochastic dynamics, the fusion pathways were exploited by many
872 researchers\cite{Noguchi2001,Noguchi2002,Shillcock2005}. Due to the
873 difficulty of numerical integration of anisotropy rotation, most of
874 the rigid body models are simply modeled by sphere, cylinder,
875 ellipsoid or other regular shapes in stochastic simulations. In an
876 effort to account for the diffusion anisotropy of the arbitrary
877 particles, Fernandes and de la Torre improved the original Brownian
878 dynamics simulation algorithm\cite{Ermak1978,Allison1991} by
879 incorporating a generalized $6\times6$ diffusion tensor and
880 introducing a simple rotation evolution scheme consisting of three
881 consecutive rotations\cite{Fernandes2002}. Unfortunately, unexpected
882 error and bias are introduced into the system due to the arbitrary
883 order of applying the noncommuting rotation
884 operators\cite{Beard2003}. Based on the observation the momentum
885 relaxation time is much less than the time step, one may ignore the
886 inertia in Brownian dynamics. However, assumption of the zero
887 average acceleration is not always true for cooperative motion which
888 is common in protein motion. An inertial Brownian dynamics (IBD) was
889 proposed to address this issue by adding an inertial correction
890 term\cite{Beard2003}. As a complement to IBD which has a lower bound
891 in time step because of the inertial relaxation time, long-time-step
892 inertial dynamics (LTID) can be used to investigate the inertial
893 behavior of the polymer segments in low friction
894 regime\cite{Beard2003}. LTID can also deal with the rotational
895 dynamics for nonskew bodies without translation-rotation coupling by
896 separating the translation and rotation motion and taking advantage
897 of the analytical solution of hydrodynamics properties. However,
898 typical nonskew bodies like cylinder and ellipsoid are inadequate to
899 represent most complex macromolecule assemblies. These intricate
900 molecules have been represented by a set of beads and their
901 hydrodynamics properties can be calculated using variant
902 hydrodynamic interaction tensors.
903
904 The goal of the present work is to develop a Langevin dynamics
905 algorithm for arbitrary rigid particles by integrating the accurate
906 estimation of friction tensor from hydrodynamics theory into the
907 sophisticated rigid body dynamics.
908
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 \[
931 \left( \begin{array}{l}
932 F_R  \\
933 \tau _R  \\
934 \end{array} \right) =  - \left( {\begin{array}{*{20}c}
935   {\Xi ^{tt} } & {\Xi ^{rt} }  \\
936   {\Xi ^{tr} } & {\Xi ^{rr} }  \\
937 \end{array}} \right)\left( \begin{array}{l}
938 v \\
939 w \\
940 \end{array} \right)
941 \]
942 where $F_r$ is the friction force and $\tau _R$ is the friction
943 toque.
944
945 \subsubsection{\label{introSection:resistanceTensorRegular}\textbf{The Resistance Tensor for Regular Shape}}
946
947 For a spherical particle, the translational and rotational friction
948 constant can be calculated from Stoke's law,
949 \[
950 \Xi ^{tt}  = \left( {\begin{array}{*{20}c}
951   {6\pi \eta R} & 0 & 0  \\
952   0 & {6\pi \eta R} & 0  \\
953   0 & 0 & {6\pi \eta R}  \\
954 \end{array}} \right)
955 \]
956 and
957 \[
958 \Xi ^{rr}  = \left( {\begin{array}{*{20}c}
959   {8\pi \eta R^3 } & 0 & 0  \\
960   0 & {8\pi \eta R^3 } & 0  \\
961   0 & 0 & {8\pi \eta R^3 }  \\
962 \end{array}} \right)
963 \]
964 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 \[
974 \frac{{x^2 }}{{a^2 }} + \frac{{y^2 }}{{b^2 }} + \frac{{z^2 }}{{c^2
975 }} = 1
976 \]
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 \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 \[
1029 v'_i  = v_i  - \sum\limits_{j \ne i} {T_{ij} F_j }
1030 \]
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 F_i  = \zeta _i v_i  - \zeta _i \sum\limits_{j \ne i} {T_{ij} F_j }.
1036 \label{introEquation:tensorExpression}
759   \end{equation}
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 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 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 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 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 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
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 \[
1079 B_{ij}  = \delta _{ij} \frac{I}{{6\pi \eta R}} + (1 - \delta _{ij}
1080 )T_{ij}
1081 \]
1082 where $\delta _{ij}$ is Kronecker delta function. Inverting matrix
1083 $B$, we obtain
1084
1085 \[
1086 C = B^{ - 1}  = \left( {\begin{array}{*{20}c}
1087   {C_{11} } &  \ldots  & {C_{1N} }  \\
1088    \vdots  &  \ddots  &  \vdots   \\
1089   {C_{N1} } &  \cdots  & {C_{NN} }  \\
1090 \end{array}} \right)
1091 \]
1092 , 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 \[
1095 U_i  = \left( {\begin{array}{*{20}c}
1096   0 & { - z_i } & {y_i }  \\
1097   {z_i } & 0 & { - x_i }  \\
1098   { - y_i } & {x_i } & 0  \\
1099 \end{array}} \right)
1100 \]
1101 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 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 \[
1140 U_{OP}  = \left( {\begin{array}{*{20}c}
1141   0 & { - z_{OP} } & {y_{OP} }  \\
1142   {z_i } & 0 & { - x_{OP} }  \\
1143   { - y_{OP} } & {x_{OP} } & 0  \\
1144 \end{array}} \right)
1145 \]
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 where $x_OR$, $y_OR$, $z_OR$ are the components of the vector
1167 joining center of resistance $R$ and origin $O$.
1168
1169 \subsection{Langevin dynamics for rigid particles of arbitrary shape}
1170
1171 Consider a Langevin equation of motions in generalized coordinates
1172 \begin{equation}
1173 M_i \dot V_i (t) = F_{s,i} (t) + F_{f,i(t)}  + F_{r,i} (t)
1174 \label{LDGeneralizedForm}
1175 \end{equation}
1176 where $M_i$ is a $6\times6$ generalized diagonal mass (include mass
1177 and moment of inertial) matrix and $V_i$ is a generalized velocity,
1178 $V_i = V_i(v_i,\omega _i)$. The right side of Eq.
1179 (\ref{LDGeneralizedForm}) consists of three generalized forces in
1180 lab-fixed frame, systematic force $F_{s,i}$, dissipative force
1181 $F_{f,i}$ and stochastic force $F_{r,i}$. While the evolution of the
1182 system in Newtownian mechanics typically refers to lab-fixed frame,
1183 it is also convenient to handle the rotation of rigid body in
1184 body-fixed frame. Thus the friction and random forces are calculated
1185 in body-fixed frame and converted back to lab-fixed frame by:
1186 \[
1187 \begin{array}{l}
1188 F_{f,i}^l (t) = A^T F_{f,i}^b (t), \\
1189 F_{r,i}^l (t) = A^T F_{r,i}^b (t) \\
1190 \end{array}.
1191 \]
1192 Here, the body-fixed friction force $F_{r,i}^b$ is proportional to
1193 the body-fixed velocity at center of resistance $v_{R,i}^b$ and
1194 angular velocity $\omega _i$,
1195 \begin{equation}
1196 F_{r,i}^b (t) = \left( \begin{array}{l}
1197 f_{r,i}^b (t) \\
1198 \tau _{r,i}^b (t) \\
1199 \end{array} \right) =  - \left( {\begin{array}{*{20}c}
1200   {\Xi _{R,t} } & {\Xi _{R,c}^T }  \\
1201   {\Xi _{R,c} } & {\Xi _{R,r} }  \\
1202 \end{array}} \right)\left( \begin{array}{l}
1203 v_{R,i}^b (t) \\
1204 \omega _i (t) \\
1205 \end{array} \right),
1206 \end{equation}
1207 while the random force $F_{r,i}^l$ is a Gaussian stochastic variable
1208 with zero mean and variance
1209 \begin{equation}
1210 \left\langle {F_{r,i}^l (t)(F_{r,i}^l (t'))^T } \right\rangle  =
1211 \left\langle {F_{r,i}^b (t)(F_{r,i}^b (t'))^T } \right\rangle  =
1212 2k_B T\Xi _R \delta (t - t').
1213 \end{equation}
1214 The equation of motion for $v_i$ can be written as
1215 \begin{equation}
1216 m\dot v_i (t) = f_{t,i} (t) = f_{s,i} (t) + f_{f,i}^l (t) +
1217 f_{r,i}^l (t)
1218 \end{equation}
1219 Since the frictional force is applied at the center of resistance
1220 which generally does not coincide with the center of mass, an extra
1221 torque is exerted at the center of mass. Thus, the net body-fixed
1222 frictional torque at the center of mass, $\tau _{n,i}^b (t)$, is
1223 given by
1224 \begin{equation}
1225 \tau _{r,i}^b = \tau _{r,i}^b +r_{MR} \times f_{r,i}^b
1226 \end{equation}
1227 where $r_{MR}$ is the vector from the center of mass to the center
1228 of the resistance. Instead of integrating angular velocity in
1229 lab-fixed frame, we consider the equation of motion of angular
1230 momentum in body-fixed frame
1231 \begin{equation}
1232 \dot \pi _i (t) = \tau _{t,i} (t) = \tau _{s,i} (t) + \tau _{f,i}^b
1233 (t) + \tau _{r,i}^b(t)
1234 \end{equation}
1235
1236 Embedding the friction terms into force and torque, one can
1237 integrate the langevin equations of motion for rigid body of
1238 arbitrary shape in a velocity-Verlet style 2-part algorithm, where
1239 $h= \delta t$:
1240
1241 {\tt part one:}
1242 \begin{align*}
1243 v_i (t + h/2) &\leftarrow v_i (t) + \frac{{hf_{t,i}^l (t)}}{{2m_i }} \\
1244 \pi _i (t + h/2) &\leftarrow \pi _i (t) + \frac{{h\tau _{t,i}^b (t)}}{2} \\
1245 r_i (t + h) &\leftarrow r_i (t) + hv_i (t + h/2) \\
1246 A_i (t + h) &\leftarrow rotate\left( {h\pi _i (t + h/2)I^{ - 1} } \right) \\
1247 \end{align*}
1248 In this context, the $\mathrm{rotate}$ function is the reversible
1249 product of five consecutive body-fixed rotations,
1250 \begin{equation}
1251 \mathrm{rotate}({\bf a}) = \mathsf{G}_x(a_x / 2) \cdot
1252 \mathsf{G}_y(a_y / 2) \cdot \mathsf{G}_z(a_z) \cdot \mathsf{G}_y(a_y
1253 / 2) \cdot \mathsf{G}_x(a_x /2),
1254 \end{equation}
1255 where each rotational propagator, $\mathsf{G}_\alpha(\theta)$,
1256 rotates both the rotation matrix ($\mathsf{A}$) and the body-fixed
1257 angular momentum ($\pi$) by an angle $\theta$ around body-fixed axis
1258 $\alpha$,
1259 \begin{equation}
1260 \mathsf{G}_\alpha( \theta ) = \left\{
1261 \begin{array}{lcl}
1262 \mathsf{A}(t) & \leftarrow & \mathsf{A}(0) \cdot \mathsf{R}_\alpha(\theta)^T, \\
1263 {\bf j}(t) & \leftarrow & \mathsf{R}_\alpha(\theta) \cdot {\bf
1264 j}(0).
1265 \end{array}
1266 \right.
1267 \end{equation}
1268 $\mathsf{R}_\alpha$ is a quadratic approximation to the single-axis
1269 rotation matrix.  For example, in the small-angle limit, the
1270 rotation matrix around the body-fixed x-axis can be approximated as
1271 \begin{equation}
1272 \mathsf{R}_x(\theta) \approx \left(
1273 \begin{array}{ccc}
1274 1 & 0 & 0 \\
1275 0 & \frac{1-\theta^2 / 4}{1 + \theta^2 / 4}  & -\frac{\theta}{1+
1276 \theta^2 / 4} \\
1277 0 & \frac{\theta}{1+ \theta^2 / 4} & \frac{1-\theta^2 / 4}{1 +
1278 \theta^2 / 4}
1279 \end{array}
1280 \right).
1281 \end{equation}
1282 All other rotations follow in a straightforward manner.
1283
1284 After the first part of the propagation, the friction and random
1285 forces are generated at the center of resistance in body-fixed frame
1286 and converted back into lab-fixed frame
1287 \[
1288 f_{t,i}^l (t + h) =  - \left( {\frac{{\partial V}}{{\partial r_i }}}
1289 \right)_{r_i (t + h)}  + A_i^T (t + h)[F_{f,i}^b (t + h) + F_{r,i}^b
1290 (t + h)],
1291 \]
1292 while the system torque in lab-fixed frame is transformed into
1293 body-fixed frame,
1294 \[
1295 \tau _{t,i}^b (t + h) = A\tau _{s,i}^l (t) + \tau _{n,i}^b (t) +
1296 \tau _{r,i}^b (t).
1297 \]
1298 Once the forces and torques have been obtained at the new time step,
1299 the velocities can be advanced to the same time value.
1300
1301 {\tt part two:}
1302 \begin{align*}
1303 v_i (t) &\leftarrow v_i (t + h/2) + \frac{{hf_{t,i}^l (t + h)}}{{2m_i }} \\
1304 \pi _i (t) &\leftarrow \pi _i (t + h/2) + \frac{{h\tau _{t,i}^b (t + h)}}{2} \\
1305 \end{align*}
1306
1307 \subsection{Results and discussion}

Diff Legend

Removed lines
+ Added lines
< Changed lines
> Changed lines