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 2807 by tim, Wed Jun 7 01:49:15 2006 UTC vs.
Revision 2882 by tim, Fri Jun 23 21:33:52 2006 UTC

# Line 2 | Line 2 | In order to mimic the experiments, which are usually p
2  
3   \section{\label{methodSection:rigidBodyIntegrators}Integrators for Rigid Body Motion in Molecular Dynamics}
4  
5 < In order to mimic the experiments, which are usually performed under
5 > In order to mimic experiments which are usually performed under
6   constant temperature and/or pressure, extended Hamiltonian system
7   methods have been developed to generate statistical ensembles, such
8 < as canonical ensemble and isobaric-isothermal ensemble \textit{etc}.
9 < In addition to the standard ensemble, specific ensembles have been
10 < developed to account for the anisotropy between the lateral and
11 < normal directions of membranes. The $NPAT$ ensemble, in which the
12 < normal pressure and the lateral surface area of the membrane are
13 < kept constant, and the $NP\gamma T$ ensemble, in which the normal
14 < pressure and the lateral surface tension are kept constant were
15 < proposed to address this issue.
8 > as the canonical and isobaric-isothermal ensembles. In addition to
9 > the standard ensemble, specific ensembles have been developed to
10 > account for the anisotropy between the lateral and normal directions
11 > of membranes. The $NPAT$ ensemble, in which the normal pressure and
12 > the lateral surface area of the membrane are kept constant, and the
13 > $NP\gamma T$ ensemble, in which the normal pressure and the lateral
14 > surface tension are kept constant were proposed to address the
15 > issues.
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
17 > Integration schemes for the rotational motion of the rigid molecules
18 > in the microcanonical ensemble have been extensively studied over
19 > the last two decades. Matubayasi 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
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
32 < proposed by Dullweber, Leimkuhler and McLachlan (DLM) also preserved
33 < the same structural properties of the Hamiltonian flow. In this
34 < section, the integration scheme of DLM method will be reviewed and
35 < extended to other ensembles.
22 > long-time energy conservation than Euler angle methods because of
23 > the time-reversible nature. Extending the Trotter-Suzuki
24 > factorization to general system with a flat phase space, Miller and
25 > his colleagues devised a novel symplectic, time-reversible and
26 > volume-preserving integrator in the quaternion representation, which
27 > was shown to be superior to the Matubayasi's time-reversible
28 > integrator. However, all of the integrators in the quaternion
29 > representation suffer from the computational penalty of constructing
30 > a rotation matrix from quaternions to evolve coordinates and
31 > velocities at every time step. An alternative integration scheme
32 > utilizing the rotation matrix directly proposed by Dullweber,
33 > Leimkuhler and McLachlan (DLM) also preserved the same structural
34 > properties of the Hamiltonian flow. In this section, the integration
35 > scheme of DLM method will be reviewed and extended to other
36 > ensembles.
37  
38   \subsection{\label{methodSection:DLM}Integrating the Equations of Motion: the
39   DLM method}
# Line 111 | Line 112 | torques are calculated at the new positions and orient
112      - \left(\frac{\partial V}{\partial {\bf r}}\right)_{{\bf r}(t + h)}, \\
113   %
114   {\bf \tau}^{s}(t + h) &\leftarrow {\bf u}(t + h)
115 <    \times \frac{\partial V}{\partial {\bf u}}, \\
115 >    \times (\frac{\partial V}{\partial {\bf u}})_{u(t+h)}, \\
116   %
117   {\bf \tau}^{b}(t + h) &\leftarrow \mathsf{A}(t + h)
118      \cdot {\bf \tau}^s(t + h).
119   \end{align*}
120  
121 < {\sc oopse} automatically updates ${\bf u}$ when the rotation matrix
121 > ${\bf u}$ is automatically updated when the rotation matrix
122   $\mathsf{A}$ is calculated in {\tt moveA}.  Once the forces and
123   torques have been obtained at the new time step, the velocities can
124   be advanced to the same time value.
# Line 198 | Line 199 | and $K$ is the total kinetic energy,
199   \begin{equation}
200   f = 3 N + 3 N_{\mathrm{orient}} - N_{\mathrm{constraints}},
201   \end{equation}
202 < and $K$ is the total kinetic energy,
202 > where $N_{\mathrm{orient}}$ is the number of molecules with
203 > orientational degrees of freedom, and $K$ is the total kinetic
204 > energy,
205   \begin{equation}
206   K = \sum_{i=1}^{N} \frac{1}{2} m_i {\bf v}_i^T \cdot {\bf v}_i +
207   \sum_{i=1}^{N_{\mathrm{orient}}}  \frac{1}{2} {\bf j}_i^T \cdot
# Line 206 | Line 209 | relaxation of the temperature to the target value.  To
209   \end{equation}
210  
211   In eq.(\ref{eq:nosehooverext}), $\tau_T$ is the time constant for
212 < relaxation of the temperature to the target value.  To set values
213 < for $\tau_T$ or $T_{\mathrm{target}}$ in a simulation, one would use
214 < the {\tt tauThermostat} and {\tt targetTemperature} keywords in the
212 < {\tt .bass} file.  The units for {\tt tauThermostat} are fs, and the
213 < units for the {\tt targetTemperature} are degrees K.   The
214 < integration of the equations of motion is carried out in a
215 < velocity-Verlet style 2 part algorithm:
212 > relaxation of the temperature to the target value. The integration
213 > of the equations of motion is carried out in a velocity-Verlet style
214 > 2 part algorithm:
215  
216   {\tt moveA:}
217   \begin{align*}
# Line 278 | Line 277 | self-consistent.  The relative tolerance for the self-
277   caclculate $T(t + h)$ as well as $\chi(t + h)$, they indirectly
278   depend on their own values at time $t + h$.  {\tt moveB} is
279   therefore done in an iterative fashion until $\chi(t + h)$ becomes
280 < self-consistent.  The relative tolerance for the self-consistency
282 < check defaults to a value of $\mbox{10}^{-6}$, but {\sc oopse} will
283 < terminate the iteration after 4 loops even if the consistency check
284 < has not been satisfied.
280 > self-consistent.
281  
282   The Nos\'e-Hoover algorithm is known to conserve a Hamiltonian for
283   the extended system that is, to within a constant, identical to the
# Line 299 | Line 295 | To carry out isobaric-isothermal ensemble calculations
295   \subsection{\label{methodSection:NPTi}Constant-pressure integration with
296   isotropic box deformations (NPTi)}
297  
298 < To carry out isobaric-isothermal ensemble calculations {\sc oopse}
299 < implements the Melchionna modifications to the
298 > We can used an isobaric-isothermal ensemble integrator which is
299 > implemented using the Melchionna modifications to the
300   Nos\'e-Hoover-Andersen equations of motion,\cite{Melchionna1993}
301  
302   \begin{eqnarray}
# Line 356 | Line 352 | relaxation of the pressure to the target value.  To se
352   \end{equation}
353  
354   In eq.(\ref{eq:melchionna1}), $\tau_B$ is the time constant for
355 < relaxation of the pressure to the target value.  To set values for
360 < $\tau_B$ or $P_{\mathrm{target}}$ in a simulation, one would use the
361 < {\tt tauBarostat} and {\tt targetPressure} keywords in the {\tt
362 < .bass} file.  The units for {\tt tauBarostat} are fs, and the units
363 < for the {\tt targetPressure} are atmospheres.  Like in the NVT
355 > relaxation of the pressure to the target value. Like in the NVT
356   integrator, the integration of the equations of motion is carried
357   out in a velocity-Verlet style 2 part algorithm:
358  
# Line 401 | Line 393 | depends on the positions at the same time.  {\sc oopse
393  
394   Most of these equations are identical to their counterparts in the
395   NVT integrator, but the propagation of positions to time $t + h$
396 < depends on the positions at the same time.  {\sc oopse} carries out
397 < this step iteratively (with a limit of 5 passes through the
398 < iterative loop).  Also, the simulation box $\mathsf{H}$ is scaled
399 < uniformly for one full time step by an exponential factor that
400 < depends on the value of $\eta$ at time $t + h / 2$.  Reshaping the
409 < box uniformly also scales the volume of the box by
396 > depends on the positions at the same time. The simulation box
397 > $\mathsf{H}$ is scaled uniformly for one full time step by an
398 > exponential factor that depends on the value of $\eta$ at time $t +
399 > h / 2$.  Reshaping the box uniformly also scales the volume of the
400 > box by
401   \begin{equation}
402   \mathcal{V}(t + h) \leftarrow e^{ - 3 h \eta(t + h /2)}.
403   \mathcal{V}(t)
# Line 448 | Line 439 | + h)$ and $\eta(t + h)$ become self-consistent.  The r
439   to caclculate $T(t + h)$, $P(t + h)$, $\chi(t + h)$, and $\eta(t +
440   h)$, they indirectly depend on their own values at time $t + h$.
441   {\tt moveB} is therefore done in an iterative fashion until $\chi(t
442 < + h)$ and $\eta(t + h)$ become self-consistent.  The relative
452 < tolerance for the self-consistency check defaults to a value of
453 < $\mbox{10}^{-6}$, but {\sc oopse} will terminate the iteration after
454 < 4 loops even if the consistency check has not been satisfied.
442 > + h)$ and $\eta(t + h)$ become self-consistent.
443  
444   The Melchionna modification of the Nos\'e-Hoover-Andersen algorithm
445   is known to conserve a Hamiltonian for the extended system that is,
# Line 473 | Line 461 | Bond constraints are applied at the end of both the {\
461   P_{\mathrm{target}} \mathcal{V}(t).
462   \end{equation}
463  
476 Bond constraints are applied at the end of both the {\tt moveA} and
477 {\tt moveB} portions of the algorithm.
478
464   \subsection{\label{methodSection:NPTf}Constant-pressure integration with a
465   flexible box (NPTf)}
466  
# Line 549 | Line 534 | r}(t)\right\},
534   \mathsf{H}(t + h) &\leftarrow \mathsf{H}(t) \cdot e^{-h
535      \overleftrightarrow{\eta}(t + h / 2)} .
536   \end{align*}
537 < {\sc oopse} uses a power series expansion truncated at second order
538 < for the exponential operation which scales the simulation box.
537 > Here, a power series expansion truncated at second order for the
538 > exponential operation is used to scale the simulation box.
539  
540   The {\tt moveB} portion of the algorithm is largely unchanged from
541   the NPTi integrator:
# Line 589 | Line 574 | The NPTf integrator is known to conserve the following
574   identical to those described for the NPTi integrator.
575  
576   The NPTf integrator is known to conserve the following Hamiltonian:
577 < \begin{equation}
578 < H_{\mathrm{NPTf}} = V + K + f k_B T_{\mathrm{target}} \left(
577 > \begin{eqnarray*}
578 > H_{\mathrm{NPTf}} & = & V + K + f k_B T_{\mathrm{target}} \left(
579   \frac{\tau_{T}^2 \chi^2(t)}{2} + \int_{0}^{t} \chi(t^\prime)
580 < dt^\prime \right) + P_{\mathrm{target}} \mathcal{V}(t) + \frac{f k_B
580 > dt^\prime \right) \\
581 > & & + P_{\mathrm{target}} \mathcal{V}(t) + \frac{f k_B
582   T_{\mathrm{target}}}{2}
583   \mathrm{Tr}\left[\overleftrightarrow{\eta}(t)\right]^2 \tau_B^2.
584 < \end{equation}
584 > \end{eqnarray*}
585  
586   This integrator must be used with care, particularly in liquid
587   simulations.  Liquids have very small restoring forces in the
# Line 605 | Line 591 | assume non-orthorhombic geometries.
591   finds most use in simulating crystals or liquid crystals which
592   assume non-orthorhombic geometries.
593  
594 < \subsection{\label{methodSection:otherSpecialEnsembles}Other Special Ensembles}
594 > \subsection{\label{methodSection:NPAT}NPAT Ensemble}
595  
596 < \subsubsection{\label{methodSection:NPAT}NPAT Ensemble}
597 <
598 < A comprehensive understanding of structure¨Cfunction relations of
599 < biological membrane system ultimately relies on structure and
600 < dynamics of lipid bilayer, which are strongly affected by the
601 < interfacial interaction between lipid molecules and surrounding
602 < media. One quantity to describe the interfacial interaction is so
603 < called the average surface area per lipid. Constat area and constant
604 < lateral pressure simulation can be achieved by extending the
619 < standard NPT ensemble with a different pressure control strategy
596 > A comprehensive understanding of relations between structures and
597 > functions in biological membrane system ultimately relies on
598 > structure and dynamics of lipid bilayers, which are strongly
599 > affected by the interfacial interaction between lipid molecules and
600 > surrounding media. One quantity to describe the interfacial
601 > interaction is so called the average surface area per lipid.
602 > Constant area and constant lateral pressure simulations can be
603 > achieved by extending the standard NPT ensemble with a different
604 > pressure control strategy
605  
606   \begin{equation}
607   \dot {\overleftrightarrow{\eta}} _{\alpha \beta}=\left\{\begin{array}{ll}
# Line 630 | Line 615 | described for the NPTi integrator.
615   Note that the iterative schemes for NPAT are identical to those
616   described for the NPTi integrator.
617  
618 < \subsubsection{\label{methodSection:NPrT}NP$\gamma$T Ensemble}
618 > \subsection{\label{methodSection:NPrT}NP$\gamma$T
619 > Ensemble}
620  
621   Theoretically, the surface tension $\gamma$ of a stress free
622   membrane system should be zero since its surface free energy $G$ is
# Line 640 | Line 626 | the membrane simulation, a special ensemble, NP$\gamma
626   \]
627   However, a surface tension of zero is not appropriate for relatively
628   small patches of membrane. In order to eliminate the edge effect of
629 < the membrane simulation, a special ensemble, NP$\gamma$T, is
629 > the membrane simulation, a special ensemble, NP$\gamma$T, has been
630   proposed to maintain the lateral surface tension and normal
631 < pressure. The equation of motion for cell size control tensor,
631 > pressure. The equation of motion for the cell size control tensor,
632   $\eta$, in $NP\gamma T$ is
633   \begin{equation}
634   \dot {\overleftrightarrow{\eta}} _{\alpha \beta}=\left\{\begin{array}{ll}
# Line 666 | Line 652 | $\gamma$ is set to zero.
652   axes are allowed to fluctuate independently, but the angle between
653   the box axes does not change. It should be noted that the NPTxyz
654   integrator is a special case of $NP\gamma T$ if the surface tension
655 < $\gamma$ is set to zero.
655 > $\gamma$ is set to zero, and if $x$ and $y$ can move independently.
656  
657 < \section{\label{methodSection:zcons}Z-Constraint Method}
657 > \section{\label{methodSection:zcons}The Z-Constraint Method}
658  
659   Based on the fluctuation-dissipation theorem, a force
660   auto-correlation method was developed by Roux and Karplus to
# Line 706 | Line 692 | After the force calculation, define $G_\alpha$ as
692   forces from the rest of the system after the force calculation at
693   each time step instead of resetting the coordinate.
694  
695 < After the force calculation, define $G_\alpha$ as
695 > After the force calculation, we define $G_\alpha$ as
696   \begin{equation}
697   G_{\alpha} = \sum_i F_{\alpha i}, \label{oopseEq:zc1}
698   \end{equation}
# Line 757 | Line 743 | F_{z_{\text{Harmonic}}}(t)=-\frac{\partial U(t)}{\part
743   F_{z_{\text{Harmonic}}}(t)=-\frac{\partial U(t)}{\partial z(t)}=
744      -k_{\text{Harmonic}}(z(t)-z_{\text{cons}}).
745   \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 %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
822 It is very important to develop stable and efficient methods to
823 integrate the equations of motion of orientational degrees of
824 freedom. Euler angles are the nature choice to describe the
825 rotational degrees of freedom. However, due to its singularity, the
826 numerical integration of corresponding equations of motion is very
827 inefficient and inaccurate. Although an alternative integrator using
828 different sets of Euler angles can overcome this
829 difficulty\cite{Ryckaert1977, Andersen1983}, the computational
830 penalty and the lost of angular momentum conservation still remain.
831 In 1977, a singularity free representation utilizing quaternions was
832 developed by Evans\cite{Evans1977}. Unfortunately, this approach
833 suffer from the nonseparable Hamiltonian resulted from quaternion
834 representation, which prevents the symplectic algorithm to be
835 utilized. Another different approach is to apply holonomic
836 constraints to the atoms belonging to the rigid
837 body\cite{Barojas1973}. Each atom moves independently under the
838 normal forces deriving from potential energy and constraint forces
839 which are used to guarantee the rigidness. However, due to their
840 iterative nature, SHAKE and Rattle algorithm converge very slowly
841 when the number of constraint increases.
842
843 The break through in geometric literature suggests that, in order to
844 develop a long-term integration scheme, one should preserve the
845 geometric structure of the flow. Matubayasi and Nakahara developed a
846 time-reversible integrator for rigid bodies in quaternion
847 representation. Although it is not symplectic, this integrator still
848 demonstrates a better long-time energy conservation than traditional
849 methods because of the time-reversible nature. Extending
850 Trotter-Suzuki to general system with a flat phase space, Miller and
851 his colleagues devised an novel symplectic, time-reversible and
852 volume-preserving integrator in quaternion representation. However,
853 all of the integrators in quaternion representation suffer from the
854 computational penalty of constructing a rotation matrix from
855 quaternions to evolve coordinates and velocities at every time step.
856 An alternative integration scheme utilizing rotation matrix directly
857 is RSHAKE\cite{Kol1997}, in which a conjugate momentum to rotation
858 matrix is introduced to re-formulate the Hamiltonian's equation and
859 the Hamiltonian is evolved in a constraint manifold by iteratively
860 satisfying the orthogonality constraint. However, RSHAKE is
861 inefficient because of the iterative procedure. An extremely
862 efficient integration scheme in rotation matrix representation,
863 which also preserves the same structural properties of the
864 Hamiltonian flow as Miller's integrator, is proposed by Dullweber,
865 Leimkuhler and McLachlan (DLM)\cite{Dullweber1997}.
866
867 %review langevin/browninan dynamics for arbitrarily shaped rigid body
868 Combining Langevin or Brownian dynamics with rigid body dynamics,
869 one can study the slow processes in biomolecular systems. Modeling
870 the DNA as a chain of rigid spheres beads, which subject to harmonic
871 potentials as well as excluded volume potentials, Mielke and his
872 coworkers discover rapid superhelical stress generations from the
873 stochastic simulation of twin supercoiling DNA with response to
874 induced torques\cite{Mielke2004}. Membrane fusion is another key
875 biological process which controls a variety of physiological
876 functions, such as release of neurotransmitters \textit{etc}. A
877 typical fusion event happens on the time scale of millisecond, which
878 is impracticable to study using all atomistic model with newtonian
879 mechanics. With the help of coarse-grained rigid body model and
880 stochastic dynamics, the fusion pathways were exploited by many
881 researchers\cite{Noguchi2001,Noguchi2002,Shillcock2005}. Due to the
882 difficulty of numerical integration of anisotropy rotation, most of
883 the rigid body models are simply modeled by sphere, cylinder,
884 ellipsoid or other regular shapes in stochastic simulations. In an
885 effort to account for the diffusion anisotropy of the arbitrary
886 particles, Fernandes and de la Torre improved the original Brownian
887 dynamics simulation algorithm\cite{Ermak1978,Allison1991} by
888 incorporating a generalized $6\times6$ diffusion tensor and
889 introducing a simple rotation evolution scheme consisting of three
890 consecutive rotations\cite{Fernandes2002}. Unfortunately, unexpected
891 error and bias are introduced into the system due to the arbitrary
892 order of applying the noncommuting rotation
893 operators\cite{Beard2003}. Based on the observation the momentum
894 relaxation time is much less than the time step, one may ignore the
895 inertia in Brownian dynamics. However, assumption of the zero
896 average acceleration is not always true for cooperative motion which
897 is common in protein motion. An inertial Brownian dynamics (IBD) was
898 proposed to address this issue by adding an inertial correction
899 term\cite{Beard2003}. As a complement to IBD which has a lower bound
900 in time step because of the inertial relaxation time, long-time-step
901 inertial dynamics (LTID) can be used to investigate the inertial
902 behavior of the polymer segments in low friction
903 regime\cite{Beard2001}. LTID can also deal with the rotational
904 dynamics for nonskew bodies without translation-rotation coupling by
905 separating the translation and rotation motion and taking advantage
906 of the analytical solution of hydrodynamics properties. However,
907 typical nonskew bodies like cylinder and ellipsoid are inadequate to
908 represent most complex macromolecule assemblies. These intricate
909 molecules have been represented by a set of beads and their
910 hydrodynamics properties can be calculated using variant
911 hydrodynamic interaction tensors.
912
913 The goal of the present work is to develop a Langevin dynamics
914 algorithm for arbitrary rigid particles by integrating the accurate
915 estimation of friction tensor from hydrodynamics theory into the
916 sophisticated rigid body dynamics.
917
918
919 \subsection{Friction Tensor}
920
921 For an arbitrary rigid body moves in a fluid, it may experience
922 friction force $f_r$ or friction torque $\tau _r$ along the opposite
923 direction of the velocity $v$ or angular velocity $\omega$ at
924 arbitrary origin $P$,
925 \begin{equation}
926 \left( \begin{array}{l}
927 f_r  \\
928 \tau _r  \\
929 \end{array} \right) =  - \left( {\begin{array}{*{20}c}
930   {\Xi _{P,t} } & {\Xi _{P,c}^T }  \\
931   {\Xi _{P,c} } & {\Xi _{P,r} }  \\
932 \end{array}} \right)\left( \begin{array}{l}
933 \nu  \\
934 \omega  \\
935 \end{array} \right)
936 \end{equation}
937 where $\Xi _{P,t}t$ is the translation friction tensor, $\Xi _{P,r}$
938 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.
963
964 \subsection{Rigid body dynamics}
965
966 The Hamiltonian of rigid body can be separated in terms of potential
967 energy $V(r,A)$ and kinetic energy $T(p,\pi)$,
968 \[
969 H = V(r,A) + T(v,\pi )
970 \]
971 A second-order symplectic method is now obtained by the composition
972 of the flow maps,
973 \[
974 \varphi _{\Delta t}  = \varphi _{\Delta t/2,V}  \circ \varphi
975 _{\Delta t,T}  \circ \varphi _{\Delta t/2,V}.
976 \]
977 Moreover, $\varphi _{\Delta t/2,V}$ can be divided into two
978 sub-flows which corresponding to force and torque respectively,
979 \[
980 \varphi _{\Delta t/2,V}  = \varphi _{\Delta t/2,F}  \circ \varphi
981 _{\Delta t/2,\tau }.
982 \]
983 Since the associated operators of $\varphi _{\Delta t/2,F} $ and
984 $\circ \varphi _{\Delta t/2,\tau }$ are commuted, the composition
985 order inside $\varphi _{\Delta t/2,V}$ does not matter.
986
987 Furthermore, kinetic potential can be separated to translational
988 kinetic term, $T^t (p)$, and rotational kinetic term, $T^r (\pi )$,
989 \begin{equation}
990 T(p,\pi ) =T^t (p) + T^r (\pi ).
991 \end{equation}
992 where $ T^t (p) = \frac{1}{2}v^T m v $ and $T^r (\pi )$ is defined
993 by \ref{introEquation:rotationalKineticRB}. Therefore, the
994 corresponding flow maps are given by
995 \[
996 \varphi _{\Delta t,T}  = \varphi _{\Delta t,T^t }  \circ \varphi
997 _{\Delta t,T^r }.
998 \]
999 The free rigid body is an example of Lie-Poisson system with
1000 Hamiltonian function
1001 \begin{equation}
1002 T^r (\pi ) = T_1 ^r (\pi _1 ) + T_2^r (\pi _2 ) + T_3^r (\pi _3 )
1003 \label{introEquation:rotationalKineticRB}
1004 \end{equation}
1005 where $T_i^r (\pi _i ) = \frac{{\pi _i ^2 }}{{2I_i }}$ and
1006 Lie-Poisson structure matrix,
1007 \begin{equation}
1008 J(\pi ) = \left( {\begin{array}{*{20}c}
1009   0 & {\pi _3 } & { - \pi _2 }  \\
1010   { - \pi _3 } & 0 & {\pi _1 }  \\
1011   {\pi _2 } & { - \pi _1 } & 0  \\
1012 \end{array}} \right)
1013 \end{equation}
1014 Thus, the dynamics of free rigid body is governed by
1015 \begin{equation}
1016 \frac{d}{{dt}}\pi  = J(\pi )\nabla _\pi  T^r (\pi )
1017 \end{equation}
1018 One may notice that each $T_i^r$ in Equation
1019 \ref{introEquation:rotationalKineticRB} can be solved exactly. For
1020 instance, the equations of motion due to $T_1^r$ are given by
1021 \begin{equation}
1022 \frac{d}{{dt}}\pi  = R_1 \pi ,\frac{d}{{dt}}A = AR_1
1023 \label{introEqaution:RBMotionSingleTerm}
1024 \end{equation}
1025 where
1026 \[ R_1  = \left( {\begin{array}{*{20}c}
1027   0 & 0 & 0  \\
1028   0 & 0 & {\pi _1 }  \\
1029   0 & { - \pi _1 } & 0  \\
1030 \end{array}} \right).
1031 \]
1032 The solutions of Equation \ref{introEqaution:RBMotionSingleTerm} is
1033 \[
1034 \pi (\Delta t) = e^{\Delta tR_1 } \pi (0),A(\Delta t) =
1035 A(0)e^{\Delta tR_1 }
1036 \]
1037 with
1038 \[
1039 e^{\Delta tR_1 }  = \left( {\begin{array}{*{20}c}
1040   0 & 0 & 0  \\
1041   0 & {\cos \theta _1 } & {\sin \theta _1 }  \\
1042   0 & { - \sin \theta _1 } & {\cos \theta _1 }  \\
1043 \end{array}} \right),\theta _1  = \frac{{\pi _1 }}{{I_1 }}\Delta t.
1044 \]
1045 To reduce the cost of computing expensive functions in $e^{\Delta
1046 tR_1 }$, we can use Cayley transformation,
1047 \[
1048 e^{\Delta tR_1 }  \approx (1 - \Delta tR_1 )^{ - 1} (1 + \Delta tR_1
1049 )
1050 \]
1051 The flow maps for $T_2^r$ and $T_3^r$ can be found in the same
1052 manner.
1053
1054 In order to construct a second-order symplectic method, we split the
1055 angular kinetic Hamiltonian function into five terms
1056 \[
1057 T^r (\pi ) = \frac{1}{2}T_1 ^r (\pi _1 ) + \frac{1}{2}T_2^r (\pi _2
1058 ) + T_3^r (\pi _3 ) + \frac{1}{2}T_2^r (\pi _2 ) + \frac{1}{2}T_1 ^r
1059 (\pi _1 )
1060 \].
1061 Concatenating flows corresponding to these five terms, we can obtain
1062 the flow map for free rigid body,
1063 \[
1064 \varphi _{\Delta t,T^r }  = \varphi _{\Delta t/2,\pi _1 }  \circ
1065 \varphi _{\Delta t/2,\pi _2 }  \circ \varphi _{\Delta t,\pi _3 }
1066 \circ \varphi _{\Delta t/2,\pi _2 }  \circ \varphi _{\Delta t/2,\pi
1067 _1 }.
1068 \]
1069
1070 The equations of motion corresponding to potential energy and
1071 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}
1084
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
1094 \subsection{Langevin dynamics for rigid particles of arbitrary shape}
1095
1096 Consider a Langevin equation of motions in generalized coordinates
1097 \begin{equation}
1098 M_i \dot V_i (t) = F_{s,i} (t) + F_{f,i(t)}  + F_{r,i} (t)
1099 \label{LDGeneralizedForm}
1100 \end{equation}
1101 where $M_i$ is a $6\times6$ generalized diagonal mass (include mass
1102 and moment of inertial) matrix and $V_i$ is a generalized velocity,
1103 $V_i = V_i(v_i,\omega _i)$. The right side of Eq.
1104 (\ref{LDGeneralizedForm}) consists of three generalized forces in
1105 lab-fixed frame, systematic force $F_{s,i}$, dissipative force
1106 $F_{f,i}$ and stochastic force $F_{r,i}$. While the evolution of the
1107 system in Newtownian mechanics typically refers to lab-fixed frame,
1108 it is also convenient to handle the rotation of rigid body in
1109 body-fixed frame. Thus the friction and random forces are calculated
1110 in body-fixed frame and converted back to lab-fixed frame by:
1111 \[
1112 \begin{array}{l}
1113 F_{f,i}^l (t) = A^T F_{f,i}^b (t), \\
1114 F_{r,i}^l (t) = A^T F_{r,i}^b (t) \\
1115 \end{array}.
1116 \]
1117 Here, the body-fixed friction force $F_{r,i}^b$ is proportional to
1118 the body-fixed velocity at center of resistance $v_{R,i}^b$ and
1119 angular velocity $\omega _i$,
1120 \begin{equation}
1121 F_{r,i}^b (t) = \left( \begin{array}{l}
1122 f_{r,i}^b (t) \\
1123 \tau _{r,i}^b (t) \\
1124 \end{array} \right) =  - \left( {\begin{array}{*{20}c}
1125   {\Xi _{R,t} } & {\Xi _{R,c}^T }  \\
1126   {\Xi _{R,c} } & {\Xi _{R,r} }  \\
1127 \end{array}} \right)\left( \begin{array}{l}
1128 v_{R,i}^b (t) \\
1129 \omega _i (t) \\
1130 \end{array} \right),
1131 \end{equation}
1132 while the random force $F_{r,i}^l$ is a Gaussian stochastic variable
1133 with zero mean and variance
1134 \begin{equation}
1135 \left\langle {F_{r,i}^l (t)(F_{r,i}^l (t'))^T } \right\rangle  =
1136 \left\langle {F_{r,i}^b (t)(F_{r,i}^b (t'))^T } \right\rangle  =
1137 2k_B T\Xi _R \delta (t - t').
1138 \end{equation}
1139 The equation of motion for $v_i$ can be written as
1140 \begin{equation}
1141 m\dot v_i (t) = f_{t,i} (t) = f_{s,i} (t) + f_{f,i}^l (t) +
1142 f_{r,i}^l (t)
1143 \end{equation}
1144 Since the frictional force is applied at the center of resistance
1145 which generally does not coincide with the center of mass, an extra
1146 torque is exerted at the center of mass. Thus, the net body-fixed
1147 frictional torque at the center of mass, $\tau _{n,i}^b (t)$, is
1148 given by
1149 \begin{equation}
1150 \tau _{r,i}^b = \tau _{r,i}^b +r_{MR} \times f_{r,i}^b
1151 \end{equation}
1152 where $r_{MR}$ is the vector from the center of mass to the center
1153 of the resistance. Instead of integrating angular velocity in
1154 lab-fixed frame, we consider the equation of motion of angular
1155 momentum in body-fixed frame
1156 \begin{equation}
1157 \dot \pi _i (t) = \tau _{t,i} (t) = \tau _{s,i} (t) + \tau _{f,i}^b
1158 (t) + \tau _{r,i}^b(t)
1159 \end{equation}
1160
1161 Embedding the friction terms into force and torque, one can
1162 integrate the langevin equations of motion for rigid body of
1163 arbitrary shape in a velocity-Verlet style 2-part algorithm, where
1164 $h= \delta t$:
1165
1166 {\tt part one:}
1167 \begin{align*}
1168 v_i (t + h/2) &\leftarrow v_i (t) + \frac{{hf_{t,i}^l (t)}}{{2m_i }} \\
1169 \pi _i (t + h/2) &\leftarrow \pi _i (t) + \frac{{h\tau _{t,i}^b (t)}}{2} \\
1170 r_i (t + h) &\leftarrow r_i (t) + hv_i (t + h/2) \\
1171 A_i (t + h) &\leftarrow rotate\left( {h\pi _i (t + h/2)I^{ - 1} } \right) \\
1172 \end{align*}
1173 In this context, the $\mathrm{rotate}$ function is the reversible
1174 product of five consecutive body-fixed rotations,
1175 \begin{equation}
1176 \mathrm{rotate}({\bf a}) = \mathsf{G}_x(a_x / 2) \cdot
1177 \mathsf{G}_y(a_y / 2) \cdot \mathsf{G}_z(a_z) \cdot \mathsf{G}_y(a_y
1178 / 2) \cdot \mathsf{G}_x(a_x /2),
1179 \end{equation}
1180 where each rotational propagator, $\mathsf{G}_\alpha(\theta)$,
1181 rotates both the rotation matrix ($\mathsf{A}$) and the body-fixed
1182 angular momentum ($\pi$) by an angle $\theta$ around body-fixed axis
1183 $\alpha$,
1184 \begin{equation}
1185 \mathsf{G}_\alpha( \theta ) = \left\{
1186 \begin{array}{lcl}
1187 \mathsf{A}(t) & \leftarrow & \mathsf{A}(0) \cdot \mathsf{R}_\alpha(\theta)^T, \\
1188 {\bf j}(t) & \leftarrow & \mathsf{R}_\alpha(\theta) \cdot {\bf
1189 j}(0).
1190 \end{array}
1191 \right.
1192 \end{equation}
1193 $\mathsf{R}_\alpha$ is a quadratic approximation to the single-axis
1194 rotation matrix.  For example, in the small-angle limit, the
1195 rotation matrix around the body-fixed x-axis can be approximated as
1196 \begin{equation}
1197 \mathsf{R}_x(\theta) \approx \left(
1198 \begin{array}{ccc}
1199 1 & 0 & 0 \\
1200 0 & \frac{1-\theta^2 / 4}{1 + \theta^2 / 4}  & -\frac{\theta}{1+
1201 \theta^2 / 4} \\
1202 0 & \frac{\theta}{1+ \theta^2 / 4} & \frac{1-\theta^2 / 4}{1 +
1203 \theta^2 / 4}
1204 \end{array}
1205 \right).
1206 \end{equation}
1207 All other rotations follow in a straightforward manner.
1208
1209 After the first part of the propagation, the friction and random
1210 forces are generated at the center of resistance in body-fixed frame
1211 and converted back into lab-fixed frame
1212 \[
1213 f_{t,i}^l (t + h) =  - \left( {\frac{{\partial V}}{{\partial r_i }}}
1214 \right)_{r_i (t + h)}  + A_i^T (t + h)[F_{f,i}^b (t + h) + F_{r,i}^b
1215 (t + h)],
1216 \]
1217 while the system torque in lab-fixed frame is transformed into
1218 body-fixed frame,
1219 \[
1220 \tau _{t,i}^b (t + h) = A\tau _{s,i}^l (t) + \tau _{n,i}^b (t) +
1221 \tau _{r,i}^b (t).
1222 \]
1223 Once the forces and torques have been obtained at the new time step,
1224 the velocities can be advanced to the same time value.
1225
1226 {\tt part two:}
1227 \begin{align*}
1228 v_i (t) &\leftarrow v_i (t + h/2) + \frac{{hf_{t,i}^l (t + h)}}{{2m_i }} \\
1229 \pi _i (t) &\leftarrow \pi _i (t + h/2) + \frac{{h\tau _{t,i}^b (t + h)}}{2} \\
1230 \end{align*}
1231
1232 \subsection{Results and discussion}

Diff Legend

Removed lines
+ Added lines
< Changed lines
> Changed lines