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 2776 by tim, Thu May 25 21:32:14 2006 UTC vs.
Revision 2850 by tim, Sun Jun 11 01:55:48 2006 UTC

# Line 6 | Line 6 | behind classical mechanics. Firstly, One can determine
6   Closely related to Classical Mechanics, Molecular Dynamics
7   simulations are carried out by integrating the equations of motion
8   for a given system of particles. There are three fundamental ideas
9 < behind classical mechanics. Firstly, One can determine the state of
9 > behind classical mechanics. Firstly, one can determine the state of
10   a mechanical system at any time of interest; Secondly, all the
11   mechanical properties of the system at that time can be determined
12   by combining the knowledge of the properties of the system with the
# Line 17 | Line 17 | Newton¡¯s first law defines a class of inertial frames
17   \subsection{\label{introSection:newtonian}Newtonian Mechanics}
18   The discovery of Newton's three laws of mechanics which govern the
19   motion of particles is the foundation of the classical mechanics.
20 < Newton¡¯s first law defines a class of inertial frames. Inertial
20 > Newton's first law defines a class of inertial frames. Inertial
21   frames are reference frames where a particle not interacting with
22   other bodies will move with constant speed in the same direction.
23 < With respect to inertial frames Newton¡¯s second law has the form
23 > With respect to inertial frames, Newton's second law has the form
24   \begin{equation}
25 < F = \frac {dp}{dt} = \frac {mv}{dt}
25 > F = \frac {dp}{dt} = \frac {mdv}{dt}
26   \label{introEquation:newtonSecondLaw}
27   \end{equation}
28   A point mass interacting with other bodies moves with the
29   acceleration along the direction of the force acting on it. Let
30   $F_{ij}$ be the force that particle $i$ exerts on particle $j$, and
31   $F_{ji}$ be the force that particle $j$ exerts on particle $i$.
32 < Newton¡¯s third law states that
32 > Newton's third law states that
33   \begin{equation}
34   F_{ij} = -F_{ji}
35   \label{introEquation:newtonThirdLaw}
# Line 46 | Line 46 | N \equiv r \times F \label{introEquation:torqueDefinit
46   \end{equation}
47   The torque $\tau$ with respect to the same origin is defined to be
48   \begin{equation}
49 < N \equiv r \times F \label{introEquation:torqueDefinition}
49 > \tau \equiv r \times F \label{introEquation:torqueDefinition}
50   \end{equation}
51   Differentiating Eq.~\ref{introEquation:angularMomentumDefinition},
52   \[
# Line 59 | Line 59 | thus,
59   \]
60   thus,
61   \begin{equation}
62 < \dot L = r \times \dot p = N
62 > \dot L = r \times \dot p = \tau
63   \end{equation}
64   If there are no external torques acting on a body, the angular
65   momentum of it is conserved. The last conservation theorem state
# Line 68 | Line 68 | scheme for rigid body \cite{Dullweber1997}.
68   \end{equation}
69   is conserved. All of these conserved quantities are
70   important factors to determine the quality of numerical integration
71 < scheme for rigid body \cite{Dullweber1997}.
71 > schemes for rigid bodies \cite{Dullweber1997}.
72  
73   \subsection{\label{introSection:lagrangian}Lagrangian Mechanics}
74  
75 < Newtonian Mechanics suffers from two important limitations: it
76 < describes their motion in special cartesian coordinate systems.
77 < Another limitation of Newtonian mechanics becomes obvious when we
78 < try to describe systems with large numbers of particles. It becomes
79 < very difficult to predict the properties of the system by carrying
80 < out calculations involving the each individual interaction between
81 < all the particles, even if we know all of the details of the
82 < interaction. In order to overcome some of the practical difficulties
83 < which arise in attempts to apply Newton's equation to complex
84 < system, alternative procedures may be developed.
75 > Newtonian Mechanics suffers from two important limitations: motions
76 > can only be described in cartesian coordinate systems. Moreover, It
77 > become impossible to predict analytically the properties of the
78 > system even if we know all of the details of the interaction. In
79 > order to overcome some of the practical difficulties which arise in
80 > attempts to apply Newton's equation to complex system, approximate
81 > numerical procedures may be developed.
82  
83 < \subsubsection{\label{introSection:halmiltonPrinciple}Hamilton's
84 < Principle}
83 > \subsubsection{\label{introSection:halmiltonPrinciple}\textbf{Hamilton's
84 > Principle}}
85  
86   Hamilton introduced the dynamical principle upon which it is
87 < possible to base all of mechanics and, indeed, most of classical
88 < physics. Hamilton's Principle may be stated as follow,
87 > possible to base all of mechanics and most of classical physics.
88 > Hamilton's Principle may be stated as follows,
89  
90   The actual trajectory, along which a dynamical system may move from
91   one point to another within a specified time, is derived by finding
92   the path which minimizes the time integral of the difference between
93 < the kinetic, $K$, and potential energies, $U$ \cite{tolman79}.
93 > the kinetic, $K$, and potential energies, $U$.
94   \begin{equation}
95   \delta \int_{t_1 }^{t_2 } {(K - U)dt = 0} ,
96   \label{introEquation:halmitonianPrinciple1}
97   \end{equation}
98  
99   For simple mechanical systems, where the forces acting on the
100 < different part are derivable from a potential and the velocities are
101 < small compared with that of light, the Lagrangian function $L$ can
102 < be define as the difference between the kinetic energy of the system
106 < and its potential energy,
100 > different parts are derivable from a potential, the Lagrangian
101 > function $L$ can be defined as the difference between the kinetic
102 > energy of the system and its potential energy,
103   \begin{equation}
104   L \equiv K - U = L(q_i ,\dot q_i ) ,
105   \label{introEquation:lagrangianDef}
# Line 114 | Line 110 | then Eq.~\ref{introEquation:halmitonianPrinciple1} bec
110   \label{introEquation:halmitonianPrinciple2}
111   \end{equation}
112  
113 < \subsubsection{\label{introSection:equationOfMotionLagrangian}The
114 < Equations of Motion in Lagrangian Mechanics}
113 > \subsubsection{\label{introSection:equationOfMotionLagrangian}\textbf{The
114 > Equations of Motion in Lagrangian Mechanics}}
115  
116 < For a holonomic system of $f$ degrees of freedom, the equations of
117 < motion in the Lagrangian form is
116 > For a system of $f$ degrees of freedom, the equations of motion in
117 > the Lagrangian form is
118   \begin{equation}
119   \frac{d}{{dt}}\frac{{\partial L}}{{\partial \dot q_i }} -
120   \frac{{\partial L}}{{\partial q_i }} = 0,{\rm{ }}i = 1, \ldots,f
# Line 132 | Line 128 | independent of generalized velocities, the generalized
128   Arising from Lagrangian Mechanics, Hamiltonian Mechanics was
129   introduced by William Rowan Hamilton in 1833 as a re-formulation of
130   classical mechanics. If the potential energy of a system is
131 < independent of generalized velocities, the generalized momenta can
136 < be defined as
131 > independent of velocities, the momenta can be defined as
132   \begin{equation}
133   p_i = \frac{\partial L}{\partial \dot q_i}
134   \label{introEquation:generalizedMomenta}
# Line 172 | Line 167 | find
167   By identifying the coefficients of $dq_k$, $dp_k$ and dt, we can
168   find
169   \begin{equation}
170 < \frac{{\partial H}}{{\partial p_k }} = q_k
170 > \frac{{\partial H}}{{\partial p_k }} = \dot {q_k}
171   \label{introEquation:motionHamiltonianCoordinate}
172   \end{equation}
173   \begin{equation}
174 < \frac{{\partial H}}{{\partial q_k }} =  - p_k
174 > \frac{{\partial H}}{{\partial q_k }} =  - \dot {p_k}
175   \label{introEquation:motionHamiltonianMomentum}
176   \end{equation}
177   and
# Line 189 | Line 184 | known as the canonical equations of motions \cite{Gold
184   Eq.~\ref{introEquation:motionHamiltonianCoordinate} and
185   Eq.~\ref{introEquation:motionHamiltonianMomentum} are Hamilton's
186   equation of motion. Due to their symmetrical formula, they are also
187 < known as the canonical equations of motions \cite{Goldstein01}.
187 > known as the canonical equations of motions \cite{Goldstein2001}.
188  
189   An important difference between Lagrangian approach and the
190   Hamiltonian approach is that the Lagrangian is considered to be a
191 < function of the generalized velocities $\dot q_i$ and the
192 < generalized coordinates $q_i$, while the Hamiltonian is considered
193 < to be a function of the generalized momenta $p_i$ and the conjugate
194 < generalized coordinate $q_i$. Hamiltonian Mechanics is more
195 < appropriate for application to statistical mechanics and quantum
196 < mechanics, since it treats the coordinate and its time derivative as
197 < independent variables and it only works with 1st-order differential
203 < equations\cite{Marion90}.
191 > function of the generalized velocities $\dot q_i$ and coordinates
192 > $q_i$, while the Hamiltonian is considered to be a function of the
193 > generalized momenta $p_i$ and the conjugate coordinates $q_i$.
194 > Hamiltonian Mechanics is more appropriate for application to
195 > statistical mechanics and quantum mechanics, since it treats the
196 > coordinate and its time derivative as independent variables and it
197 > only works with 1st-order differential equations\cite{Marion1990}.
198  
199   In Newtonian Mechanics, a system described by conservative forces
200   conserves the total energy \ref{introEquation:energyConservation}.
# Line 230 | Line 224 | momentum variables. Consider a dynamic system in a car
224   possible states. Each possible state of the system corresponds to
225   one unique point in the phase space. For mechanical systems, the
226   phase space usually consists of all possible values of position and
227 < momentum variables. Consider a dynamic system in a cartesian space,
228 < where each of the $6f$ coordinates and momenta is assigned to one of
229 < $6f$ mutually orthogonal axes, the phase space of this system is a
230 < $6f$ dimensional space. A point, $x = (q_1 , \ldots ,q_f ,p_1 ,
231 < \ldots ,p_f )$, with a unique set of values of $6f$ coordinates and
232 < momenta is a phase space vector.
227 > momentum variables. Consider a dynamic system of $f$ particles in a
228 > cartesian space, where each of the $6f$ coordinates and momenta is
229 > assigned to one of $6f$ mutually orthogonal axes, the phase space of
230 > this system is a $6f$ dimensional space. A point, $x = (q_1 , \ldots
231 > ,q_f ,p_1 , \ldots ,p_f )$, with a unique set of values of $6f$
232 > coordinates and momenta is a phase space vector.
233  
234 + %%%fix me
235   A microscopic state or microstate of a classical system is
236   specification of the complete phase space vector of a system at any
237   instant in time. An ensemble is defined as a collection of systems
# Line 257 | Line 252 | space. The density of distribution for an ensemble wit
252   regions of the phase space. The condition of an ensemble at any time
253   can be regarded as appropriately specified by the density $\rho$
254   with which representative points are distributed over the phase
255 < space. The density of distribution for an ensemble with $f$ degrees
256 < of freedom is defined as,
255 > space. The density distribution for an ensemble with $f$ degrees of
256 > freedom is defined as,
257   \begin{equation}
258   \rho  = \rho (q_1 , \ldots ,q_f ,p_1 , \ldots ,p_f ,t).
259   \label{introEquation:densityDistribution}
260   \end{equation}
261   Governed by the principles of mechanics, the phase points change
262 < their value which would change the density at any time at phase
263 < space. Hence, the density of distribution is also to be taken as a
262 > their locations which would change the density at any time at phase
263 > space. Hence, the density distribution is also to be taken as a
264   function of the time.
265  
266   The number of systems $\delta N$ at time $t$ can be determined by,
# Line 273 | Line 268 | Assuming a large enough population of systems are expl
268   \delta N = \rho (q,p,t)dq_1  \ldots dq_f dp_1  \ldots dp_f.
269   \label{introEquation:deltaN}
270   \end{equation}
271 < Assuming a large enough population of systems are exploited, we can
272 < sufficiently approximate $\delta N$ without introducing
273 < discontinuity when we go from one region in the phase space to
274 < another. By integrating over the whole phase space,
271 > Assuming a large enough population of systems, we can sufficiently
272 > approximate $\delta N$ without introducing discontinuity when we go
273 > from one region in the phase space to another. By integrating over
274 > the whole phase space,
275   \begin{equation}
276   N = \int { \ldots \int {\rho (q,p,t)dq_1 } ...dq_f dp_1 } ...dp_f
277   \label{introEquation:totalNumberSystem}
# Line 288 | Line 283 | With the help of Equation(\ref{introEquation:unitProba
283   {\rho (q,p,t)dq_1 } ...dq_f dp_1 } ...dp_f }}.
284   \label{introEquation:unitProbability}
285   \end{equation}
286 < With the help of Equation(\ref{introEquation:unitProbability}) and
287 < the knowledge of the system, it is possible to calculate the average
286 > With the help of Eq.~\ref{introEquation:unitProbability} and the
287 > knowledge of the system, it is possible to calculate the average
288   value of any desired quantity which depends on the coordinates and
289   momenta of the system. Even when the dynamics of the real system is
290   complex, or stochastic, or even discontinuous, the average
291 < properties of the ensemble of possibilities as a whole may still
292 < remain well defined. For a classical system in thermal equilibrium
293 < with its environment, the ensemble average of a mechanical quantity,
294 < $\langle A(q , p) \rangle_t$, takes the form of an integral over the
295 < phase space of the system,
291 > properties of the ensemble of possibilities as a whole remaining
292 > well defined. For a classical system in thermal equilibrium with its
293 > environment, the ensemble average of a mechanical quantity, $\langle
294 > A(q , p) \rangle_t$, takes the form of an integral over the phase
295 > space of the system,
296   \begin{equation}
297   \langle  A(q , p) \rangle_t = \frac{{\int { \ldots \int {A(q,p)\rho
298   (q,p,t)dq_1 } ...dq_f dp_1 } ...dp_f }}{{\int { \ldots \int {\rho
# Line 307 | Line 302 | parameters, such as temperature \textit{etc}, partitio
302  
303   There are several different types of ensembles with different
304   statistical characteristics. As a function of macroscopic
305 < parameters, such as temperature \textit{etc}, partition function can
306 < be used to describe the statistical properties of a system in
305 > parameters, such as temperature \textit{etc}, the partition function
306 > can be used to describe the statistical properties of a system in
307   thermodynamic equilibrium.
308  
309   As an ensemble of systems, each of which is known to be thermally
310 < isolated and conserve energy, Microcanonical ensemble(NVE) has a
311 < partition function like,
310 > isolated and conserve energy, the Microcanonical ensemble (NVE) has
311 > a partition function like,
312   \begin{equation}
313   \Omega (N,V,E) = e^{\beta TS} \label{introEquation:NVEPartition}.
314   \end{equation}
315 < A canonical ensemble(NVT)is an ensemble of systems, each of which
315 > A canonical ensemble (NVT)is an ensemble of systems, each of which
316   can share its energy with a large heat reservoir. The distribution
317   of the total energy amongst the possible dynamical states is given
318   by the partition function,
# Line 326 | Line 321 | TS$. Since most experiment are carried out under const
321   \label{introEquation:NVTPartition}
322   \end{equation}
323   Here, $A$ is the Helmholtz free energy which is defined as $ A = U -
324 < TS$. Since most experiment are carried out under constant pressure
325 < condition, isothermal-isobaric ensemble(NPT) play a very important
326 < role in molecular simulation. The isothermal-isobaric ensemble allow
327 < the system to exchange energy with a heat bath of temperature $T$
328 < and to change the volume as well. Its partition function is given as
324 > TS$. Since most experiments are carried out under constant pressure
325 > condition, the isothermal-isobaric ensemble (NPT) plays a very
326 > important role in molecular simulations. The isothermal-isobaric
327 > ensemble allow the system to exchange energy with a heat bath of
328 > temperature $T$ and to change the volume as well. Its partition
329 > function is given as
330   \begin{equation}
331   \Delta (N,P,T) =  - e^{\beta G}.
332   \label{introEquation:NPTPartition}
# Line 339 | Line 335 | The Liouville's theorem is the foundation on which sta
335  
336   \subsection{\label{introSection:liouville}Liouville's theorem}
337  
338 < The Liouville's theorem is the foundation on which statistical
339 < mechanics rests. It describes the time evolution of phase space
338 > Liouville's theorem is the foundation on which statistical mechanics
339 > rests. It describes the time evolution of the phase space
340   distribution function. In order to calculate the rate of change of
341 < $\rho$, we begin from Equation(\ref{introEquation:deltaN}). If we
342 < consider the two faces perpendicular to the $q_1$ axis, which are
343 < located at $q_1$ and $q_1 + \delta q_1$, the number of phase points
344 < leaving the opposite face is given by the expression,
341 > $\rho$, we begin from Eq.~\ref{introEquation:deltaN}. If we consider
342 > the two faces perpendicular to the $q_1$ axis, which are located at
343 > $q_1$ and $q_1 + \delta q_1$, the number of phase points leaving the
344 > opposite face is given by the expression,
345   \begin{equation}
346   \left( {\rho  + \frac{{\partial \rho }}{{\partial q_1 }}\delta q_1 }
347   \right)\left( {\dot q_1  + \frac{{\partial \dot q_1 }}{{\partial q_1
# Line 369 | Line 365 | divining $ \delta q_1  \ldots \delta q_f \delta p_1  \
365   + \frac{{\partial \dot p_i }}{{\partial p_i }}} \right)}  = 0 ,
366   \end{equation}
367   which cancels the first terms of the right hand side. Furthermore,
368 < divining $ \delta q_1  \ldots \delta q_f \delta p_1  \ldots \delta
368 > dividing $ \delta q_1  \ldots \delta q_f \delta p_1  \ldots \delta
369   p_f $ in both sides, we can write out Liouville's theorem in a
370   simple form,
371   \begin{equation}
# Line 381 | Line 377 | statistical mechanics, since the number of particles i
377  
378   Liouville's theorem states that the distribution function is
379   constant along any trajectory in phase space. In classical
380 < statistical mechanics, since the number of particles in the system
381 < is huge, we may be able to believe the system is stationary,
380 > statistical mechanics, since the number of members in an ensemble is
381 > huge and constant, we can assume the local density has no reason
382 > (other than classical mechanics) to change,
383   \begin{equation}
384   \frac{{\partial \rho }}{{\partial t}} = 0.
385   \label{introEquation:stationary}
# Line 395 | Line 392 | distribution,
392   \label{introEquation:densityAndHamiltonian}
393   \end{equation}
394  
395 < \subsubsection{\label{introSection:phaseSpaceConservation}Conservation of Phase Space}
395 > \subsubsection{\label{introSection:phaseSpaceConservation}\textbf{Conservation of Phase Space}}
396   Lets consider a region in the phase space,
397   \begin{equation}
398   \delta v = \int { \ldots \int {dq_1 } ...dq_f dp_1 } ..dp_f .
399   \end{equation}
400   If this region is small enough, the density $\rho$ can be regarded
401 < as uniform over the whole phase space. Thus, the number of phase
402 < points inside this region is given by,
401 > as uniform over the whole integral. Thus, the number of phase points
402 > inside this region is given by,
403   \begin{equation}
404   \delta N = \rho \delta v = \rho \int { \ldots \int {dq_1 } ...dq_f
405   dp_1 } ..dp_f.
# Line 414 | Line 411 | With the help of stationary assumption
411   \end{equation}
412   With the help of stationary assumption
413   (\ref{introEquation:stationary}), we obtain the principle of the
414 < \emph{conservation of extension in phase space},
414 > \emph{conservation of volume in phase space},
415   \begin{equation}
416   \frac{d}{{dt}}(\delta v) = \frac{d}{{dt}}\int { \ldots \int {dq_1 }
417   ...dq_f dp_1 } ..dp_f  = 0.
418   \label{introEquation:volumePreserving}
419   \end{equation}
420  
421 < \subsubsection{\label{introSection:liouvilleInOtherForms}Liouville's Theorem in Other Forms}
421 > \subsubsection{\label{introSection:liouvilleInOtherForms}\textbf{Liouville's Theorem in Other Forms}}
422  
423   Liouville's theorem can be expresses in a variety of different forms
424   which are convenient within different contexts. For any two function
# Line 435 | Line 432 | Substituting equations of motion in Hamiltonian formal
432   \label{introEquation:poissonBracket}
433   \end{equation}
434   Substituting equations of motion in Hamiltonian formalism(
435 < \ref{introEquation:motionHamiltonianCoordinate} ,
436 < \ref{introEquation:motionHamiltonianMomentum} ) into
437 < (\ref{introEquation:liouvilleTheorem}), we can rewrite Liouville's
438 < theorem using Poisson bracket notion,
435 > Eq.~\ref{introEquation:motionHamiltonianCoordinate} ,
436 > Eq.~\ref{introEquation:motionHamiltonianMomentum} ) into
437 > (Eq.~\ref{introEquation:liouvilleTheorem}), we can rewrite
438 > Liouville's theorem using Poisson bracket notion,
439   \begin{equation}
440   \left( {\frac{{\partial \rho }}{{\partial t}}} \right) =  - \left\{
441   {\rho ,H} \right\}.
# Line 463 | Line 460 | simulation and the quality of the underlying model. Ho
460   Various thermodynamic properties can be calculated from Molecular
461   Dynamics simulation. By comparing experimental values with the
462   calculated properties, one can determine the accuracy of the
463 < simulation and the quality of the underlying model. However, both of
464 < experiment and computer simulation are usually performed during a
463 > simulation and the quality of the underlying model. However, both
464 > experiments and computer simulations are usually performed during a
465   certain time interval and the measurements are averaged over a
466   period of them which is different from the average behavior of
467 < many-body system in Statistical Mechanics. Fortunately, Ergodic
468 < Hypothesis is proposed to make a connection between time average and
469 < ensemble average. It states that time average and average over the
470 < statistical ensemble are identical \cite{Frenkel1996, leach01:mm}.
467 > many-body system in Statistical Mechanics. Fortunately, the Ergodic
468 > Hypothesis makes a connection between time average and the ensemble
469 > average. It states that the time average and average over the
470 > statistical ensemble are identical \cite{Frenkel1996, Leach2001}.
471   \begin{equation}
472   \langle A(q , p) \rangle_t = \mathop {\lim }\limits_{t \to \infty }
473   \frac{1}{t}\int\limits_0^t {A(q(t),p(t))dt = \int\limits_\Gamma
# Line 484 | Line 481 | reasonable, the Monte Carlo techniques\cite{metropolis
481   a properly weighted statistical average. This allows the researcher
482   freedom of choice when deciding how best to measure a given
483   observable. In case an ensemble averaged approach sounds most
484 < reasonable, the Monte Carlo techniques\cite{metropolis:1949} can be
484 > reasonable, the Monte Carlo techniques\cite{Metropolis1949} can be
485   utilized. Or if the system lends itself to a time averaging
486   approach, the Molecular Dynamics techniques in
487   Sec.~\ref{introSection:molecularDynamics} will be the best
488   choice\cite{Frenkel1996}.
489  
490   \section{\label{introSection:geometricIntegratos}Geometric Integrators}
491 < A variety of numerical integrators were proposed to simulate the
492 < motions. They usually begin with an initial conditionals and move
493 < the objects in the direction governed by the differential equations.
494 < However, most of them ignore the hidden physical law contained
495 < within the equations. Since 1990, geometric integrators, which
496 < preserve various phase-flow invariants such as symplectic structure,
497 < volume and time reversal symmetry, are developed to address this
498 < issue. The velocity verlet method, which happens to be a simple
499 < example of symplectic integrator, continues to gain its popularity
500 < in molecular dynamics community. This fact can be partly explained
501 < by its geometric nature.
491 > A variety of numerical integrators have been proposed to simulate
492 > the motions of atoms in MD simulation. They usually begin with
493 > initial conditionals and move the objects in the direction governed
494 > by the differential equations. However, most of them ignore the
495 > hidden physical laws contained within the equations. Since 1990,
496 > geometric integrators, which preserve various phase-flow invariants
497 > such as symplectic structure, volume and time reversal symmetry, are
498 > developed to address this issue\cite{Dullweber1997, McLachlan1998,
499 > Leimkuhler1999}. The velocity verlet method, which happens to be a
500 > simple example of symplectic integrator, continues to gain
501 > popularity in the molecular dynamics community. This fact can be
502 > partly explained by its geometric nature.
503  
504 < \subsection{\label{introSection:symplecticManifold}Symplectic Manifold}
505 < A \emph{manifold} is an abstract mathematical space. It locally
506 < looks like Euclidean space, but when viewed globally, it may have
507 < more complicate structure. A good example of manifold is the surface
508 < of Earth. It seems to be flat locally, but it is round if viewed as
509 < a whole. A \emph{differentiable manifold} (also known as
510 < \emph{smooth manifold}) is a manifold with an open cover in which
511 < the covering neighborhoods are all smoothly isomorphic to one
512 < another. In other words,it is possible to apply calculus on
515 < \emph{differentiable manifold}. A \emph{symplectic manifold} is
516 < defined as a pair $(M, \omega)$ which consisting of a
504 > \subsection{\label{introSection:symplecticManifold}Symplectic Manifolds}
505 > A \emph{manifold} is an abstract mathematical space. It looks
506 > locally like Euclidean space, but when viewed globally, it may have
507 > more complicated structure. A good example of manifold is the
508 > surface of Earth. It seems to be flat locally, but it is round if
509 > viewed as a whole. A \emph{differentiable manifold} (also known as
510 > \emph{smooth manifold}) is a manifold on which it is possible to
511 > apply calculus on \emph{differentiable manifold}. A \emph{symplectic
512 > manifold} is defined as a pair $(M, \omega)$ which consists of a
513   \emph{differentiable manifold} $M$ and a close, non-degenerated,
514   bilinear symplectic form, $\omega$. A symplectic form on a vector
515   space $V$ is a function $\omega(x, y)$ which satisfies
516   $\omega(\lambda_1x_1+\lambda_2x_2, y) = \lambda_1\omega(x_1, y)+
517   \lambda_2\omega(x_2, y)$, $\omega(x, y) = - \omega(y, x)$ and
518 < $\omega(x, x) = 0$. Cross product operation in vector field is an
519 < example of symplectic form.
518 > $\omega(x, x) = 0$. The cross product operation in vector field is
519 > an example of symplectic form.
520  
521 < One of the motivations to study \emph{symplectic manifold} in
521 > One of the motivations to study \emph{symplectic manifolds} in
522   Hamiltonian Mechanics is that a symplectic manifold can represent
523   all possible configurations of the system and the phase space of the
524   system can be described by it's cotangent bundle. Every symplectic
525   manifold is even dimensional. For instance, in Hamilton equations,
526   coordinate and momentum always appear in pairs.
527  
532 Let  $(M,\omega)$ and $(N, \eta)$ be symplectic manifolds. A map
533 \[
534 f : M \rightarrow N
535 \]
536 is a \emph{symplectomorphism} if it is a \emph{diffeomorphims} and
537 the \emph{pullback} of $\eta$ under f is equal to $\omega$.
538 Canonical transformation is an example of symplectomorphism in
539 classical mechanics.
540
528   \subsection{\label{introSection:ODE}Ordinary Differential Equations}
529  
530 < For a ordinary differential system defined as
530 > For an ordinary differential system defined as
531   \begin{equation}
532   \dot x = f(x)
533   \end{equation}
534 < where $x = x(q,p)^T$, this system is canonical Hamiltonian, if
534 > where $x = x(q,p)^T$, this system is a canonical Hamiltonian, if
535   \begin{equation}
536   f(r) = J\nabla _x H(r).
537   \end{equation}
# Line 565 | Line 552 | Another generalization of Hamiltonian dynamics is Pois
552   \end{equation}In this case, $f$ is
553   called a \emph{Hamiltonian vector field}.
554  
555 < Another generalization of Hamiltonian dynamics is Poisson Dynamics,
555 > Another generalization of Hamiltonian dynamics is Poisson
556 > Dynamics\cite{Olver1986},
557   \begin{equation}
558   \dot x = J(x)\nabla _x H \label{introEquation:poissonHamiltonian}
559   \end{equation}
# Line 612 | Line 600 | The hidden geometric properties of ODE and its flow pl
600  
601   \subsection{\label{introSection:geometricProperties}Geometric Properties}
602  
603 < The hidden geometric properties of ODE and its flow play important
604 < roles in numerical studies. Many of them can be found in systems
605 < which occur naturally in applications.
603 > The hidden geometric properties\cite{Budd1999, Marsden1998} of ODE
604 > and its flow play important roles in numerical studies. Many of them
605 > can be found in systems which occur naturally in applications.
606  
607   Let $\varphi$ be the flow of Hamiltonian vector field, $\varphi$ is
608   a \emph{symplectic} flow if it satisfies,
# Line 658 | Line 646 | smooth function $G$ is given by,
646   which is the condition for conserving \emph{first integral}. For a
647   canonical Hamiltonian system, the time evolution of an arbitrary
648   smooth function $G$ is given by,
649 < \begin{equation}
650 < \begin{array}{c}
651 < \frac{{dG(x(t))}}{{dt}} = [\nabla _x G(x(t))]^T \dot x(t) \\
652 <  = [\nabla _x G(x(t))]^T J\nabla _x H(x(t)). \\
665 < \end{array}
649 >
650 > \begin{eqnarray}
651 > \frac{{dG(x(t))}}{{dt}} & = & [\nabla _x G(x(t))]^T \dot x(t) \\
652 >                        & = & [\nabla _x G(x(t))]^T J\nabla _x H(x(t)). \\
653   \label{introEquation:firstIntegral1}
654 < \end{equation}
654 > \end{eqnarray}
655 >
656 >
657   Using poisson bracket notion, Equation
658   \ref{introEquation:firstIntegral1} can be rewritten as
659   \[
# Line 679 | Line 668 | is a \emph{first integral}, which is due to the fact $
668   is a \emph{first integral}, which is due to the fact $\{ H,H\}  =
669   0$.
670  
671 <
683 < When designing any numerical methods, one should always try to
671 > When designing any numerical methods, one should always try to
672   preserve the structural properties of the original ODE and its flow.
673  
674   \subsection{\label{introSection:constructionSymplectic}Construction of Symplectic Methods}
675   A lot of well established and very effective numerical methods have
676   been successful precisely because of their symplecticities even
677   though this fact was not recognized when they were first
678 < constructed. The most famous example is leapfrog methods in
679 < molecular dynamics. In general, symplectic integrators can be
678 > constructed. The most famous example is the Verlet-leapfrog methods
679 > in molecular dynamics. In general, symplectic integrators can be
680   constructed using one of four different methods.
681   \begin{enumerate}
682   \item Generating functions
# Line 697 | Line 685 | Generating function tends to lead to methods which are
685   \item Splitting methods
686   \end{enumerate}
687  
688 < Generating function tends to lead to methods which are cumbersome
689 < and difficult to use. In dissipative systems, variational methods
690 < can capture the decay of energy accurately. Since their
691 < geometrically unstable nature against non-Hamiltonian perturbations,
692 < ordinary implicit Runge-Kutta methods are not suitable for
693 < Hamiltonian system. Recently, various high-order explicit
694 < Runge--Kutta methods have been developed to overcome this
688 > Generating function\cite{Channell1990} tends to lead to methods
689 > which are cumbersome and difficult to use. In dissipative systems,
690 > variational methods can capture the decay of energy
691 > accurately\cite{Kane2000}. Since their geometrically unstable nature
692 > against non-Hamiltonian perturbations, ordinary implicit Runge-Kutta
693 > methods are not suitable for Hamiltonian system. Recently, various
694 > high-order explicit Runge-Kutta methods
695 > \cite{Owren1992,Chen2003}have been developed to overcome this
696   instability. However, due to computational penalty involved in
697 < implementing the Runge-Kutta methods, they do not attract too much
698 < attention from Molecular Dynamics community. Instead, splitting have
699 < been widely accepted since they exploit natural decompositions of
700 < the system\cite{Tuckerman92}.
697 > implementing the Runge-Kutta methods, they have not attracted much
698 > attention from the Molecular Dynamics community. Instead, splitting
699 > methods have been widely accepted since they exploit natural
700 > decompositions of the system\cite{Tuckerman1992, McLachlan1998}.
701  
702 < \subsubsection{\label{introSection:splittingMethod}Splitting Method}
702 > \subsubsection{\label{introSection:splittingMethod}\textbf{Splitting Methods}}
703  
704   The main idea behind splitting methods is to decompose the discrete
705   $\varphi_h$ as a composition of simpler flows,
# Line 731 | Line 720 | order is then given by the Lie-Trotter formula
720   energy respectively, which is a natural decomposition of the
721   problem. If $H_1$ and $H_2$ can be integrated using exact flows
722   $\varphi_1(t)$ and $\varphi_2(t)$, respectively, a simple first
723 < order is then given by the Lie-Trotter formula
723 > order expression is then given by the Lie-Trotter formula
724   \begin{equation}
725   \varphi _h  = \varphi _{1,h}  \circ \varphi _{2,h},
726   \label{introEquation:firstOrderSplitting}
# Line 757 | Line 746 | which has a local error proportional to $h^3$. Sprang
746   \varphi _h  = \varphi _{1,h/2}  \circ \varphi _{2,h}  \circ \varphi
747   _{1,h/2} , \label{introEquation:secondOrderSplitting}
748   \end{equation}
749 < which has a local error proportional to $h^3$. Sprang splitting's
750 < popularity in molecular simulation community attribute to its
751 < symmetric property,
749 > which has a local error proportional to $h^3$. The Sprang
750 > splitting's popularity in molecular simulation community attribute
751 > to its symmetric property,
752   \begin{equation}
753   \varphi _h^{ - 1} = \varphi _{ - h}.
754   \label{introEquation:timeReversible}
755 < \end{equation}
755 > \end{equation},appendixFig:architecture
756  
757 < \subsubsection{\label{introSection:exampleSplittingMethod}Example of Splitting Method}
757 > \subsubsection{\label{introSection:exampleSplittingMethod}\textbf{Example of Splitting Method}}
758   The classical equation for a system consisting of interacting
759   particles can be written in Hamiltonian form,
760   \[
# Line 825 | Line 814 | q(\Delta t)} \right]. %
814   \label{introEquation:positionVerlet2}
815   \end{align}
816  
817 < \subsubsection{\label{introSection:errorAnalysis}Error Analysis and Higher Order Methods}
817 > \subsubsection{\label{introSection:errorAnalysis}\textbf{Error Analysis and Higher Order Methods}}
818  
819   Baker-Campbell-Hausdorff formula can be used to determine the local
820   error of splitting method in terms of commutator of the
# Line 844 | Line 833 | Applying Baker-Campbell-Hausdorff formula to Sprang sp
833   \[
834   [X,Y] = XY - YX .
835   \]
836 < Applying Baker-Campbell-Hausdorff formula to Sprang splitting, we
837 < can obtain
838 < \begin{equation}
839 < \exp (h X/2)\exp (h Y)\exp (h X/2) & = & \exp (h X + h Y + h^2
840 < [X,Y]/4 + h^2 [Y,X]/4 \\ & & \mbox{} + h^2 [X,X]/8 + h^2 [Y,Y]/8 \\
841 < & & \mbox{} + h^3 [Y,[Y,X]]/12 - h^3 [X,[X,Y]]/24 & & \mbox{} +
842 < \ldots )
854 < \end{equation}
836 > Applying Baker-Campbell-Hausdorff formula\cite{Varadarajan1974} to
837 > Sprang splitting, we can obtain
838 > \begin{eqnarray*}
839 > \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 \\
840 >                                   &   & \mbox{} + h^2 [X,X]/8 + h^2 [Y,Y]/8 \\
841 >                                   &   & \mbox{} + h^3 [Y,[Y,X]]/12 - h^3[X,[X,Y]]/24 + \ldots )
842 > \end{eqnarray*}
843   Since \[ [X,Y] + [Y,X] = 0\] and \[ [X,X] = 0\], the dominant local
844   error of Spring splitting is proportional to $h^3$. The same
845   procedure can be applied to general splitting,  of the form
# Line 859 | Line 847 | Careful choice of coefficient $a_1 ,\ldot , b_m$ will
847   \varphi _{b_m h}^2  \circ \varphi _{a_m h}^1  \circ \varphi _{b_{m -
848   1} h}^2  \circ  \ldots  \circ \varphi _{a_1 h}^1 .
849   \end{equation}
850 < Careful choice of coefficient $a_1 ,\ldot , b_m$ will lead to higher
850 > Careful choice of coefficient $a_1 \ldots b_m$ will lead to higher
851   order method. Yoshida proposed an elegant way to compose higher
852 < order methods based on symmetric splitting. Given a symmetric second
853 < order base method $ \varphi _h^{(2)} $, a fourth-order symmetric
854 < method can be constructed by composing,
852 > order methods based on symmetric splitting\cite{Yoshida1990}. Given
853 > a symmetric second order base method $ \varphi _h^{(2)} $, a
854 > fourth-order symmetric method can be constructed by composing,
855   \[
856   \varphi _h^{(4)}  = \varphi _{\alpha h}^{(2)}  \circ \varphi _{\beta
857   h}^{(2)}  \circ \varphi _{\alpha h}^{(2)}
# Line 912 | Line 900 | initialization of a simulation. Sec.~\ref{introSec:pro
900   \end{enumerate}
901   These three individual steps will be covered in the following
902   sections. Sec.~\ref{introSec:initialSystemSettings} deals with the
903 < initialization of a simulation. Sec.~\ref{introSec:production} will
904 < discusses issues in production run. Sec.~\ref{introSection:Analysis}
905 < provides the theoretical tools for trajectory analysis.
903 > initialization of a simulation. Sec.~\ref{introSection:production}
904 > will discusses issues in production run.
905 > Sec.~\ref{introSection:Analysis} provides the theoretical tools for
906 > trajectory analysis.
907  
908   \subsection{\label{introSec:initialSystemSettings}Initialization}
909  
910 < \subsubsection{Preliminary preparation}
910 > \subsubsection{\textbf{Preliminary preparation}}
911  
912   When selecting the starting structure of a molecule for molecular
913   simulation, one may retrieve its Cartesian coordinates from public
# Line 936 | Line 925 | interested in self-aggregation and it takes a long tim
925   instead of placing lipids randomly in solvent, since we are not
926   interested in self-aggregation and it takes a long time to happen.
927  
928 < \subsubsection{Minimization}
928 > \subsubsection{\textbf{Minimization}}
929  
930   It is quite possible that some of molecules in the system from
931   preliminary preparation may be overlapped with each other. This
# Line 958 | Line 947 | Newton-Raphson methods can not be used with very large
947   matrix and insufficient storage capacity to store them, most
948   Newton-Raphson methods can not be used with very large models.
949  
950 < \subsubsection{Heating}
950 > \subsubsection{\textbf{Heating}}
951  
952   Typically, Heating is performed by assigning random velocities
953   according to a Gaussian distribution for a temperature. Beginning at
# Line 970 | Line 959 | shifted to zero.
959   net linear momentum and angular momentum of the system should be
960   shifted to zero.
961  
962 < \subsubsection{Equilibration}
962 > \subsubsection{\textbf{Equilibration}}
963  
964   The purpose of equilibration is to allow the system to evolve
965   spontaneously for a period of time and reach equilibrium. The
# Line 984 | Line 973 | Production run is the most important steps of the simu
973  
974   \subsection{\label{introSection:production}Production}
975  
976 < Production run is the most important steps of the simulation, in
976 > Production run is the most important step of the simulation, in
977   which the equilibrated structure is used as a starting point and the
978   motions of the molecules are collected for later analysis. In order
979   to capture the macroscopic properties of the system, the molecular
# Line 1000 | Line 989 | approach will suffer from the surface effect. To offse
989   A natural approach to avoid system size issue is to represent the
990   bulk behavior by a finite number of the particles. However, this
991   approach will suffer from the surface effect. To offset this,
992 < \textit{Periodic boundary condition} is developed to simulate bulk
993 < properties with a relatively small number of particles. In this
994 < method, the simulation box is replicated throughout space to form an
995 < infinite lattice. During the simulation, when a particle moves in
996 < the primary cell, its image in other cells move in exactly the same
997 < direction with exactly the same orientation. Thus, as a particle
998 < leaves the primary cell, one of its images will enter through the
999 < opposite face.
1000 < %\begin{figure}
1001 < %\centering
1002 < %\includegraphics[width=\linewidth]{pbcFig.eps}
1003 < %\caption[An illustration of periodic boundary conditions]{A 2-D
1004 < %illustration of periodic boundary conditions. As one particle leaves
1005 < %the right of the simulation box, an image of it enters the left.}
1006 < %\label{introFig:pbc}
1007 < %\end{figure}
992 > \textit{Periodic boundary condition} (see Fig.~\ref{introFig:pbc})
993 > is developed to simulate bulk properties with a relatively small
994 > number of particles. In this method, the simulation box is
995 > replicated throughout space to form an infinite lattice. During the
996 > simulation, when a particle moves in the primary cell, its image in
997 > other cells move in exactly the same direction with exactly the same
998 > orientation. Thus, as a particle leaves the primary cell, one of its
999 > images will enter through the opposite face.
1000 > \begin{figure}
1001 > \centering
1002 > \includegraphics[width=\linewidth]{pbc.eps}
1003 > \caption[An illustration of periodic boundary conditions]{A 2-D
1004 > illustration of periodic boundary conditions. As one particle leaves
1005 > the left of the simulation box, an image of it enters the right.}
1006 > \label{introFig:pbc}
1007 > \end{figure}
1008  
1009   %cutoff and minimum image convention
1010   Another important technique to improve the efficiency of force
# Line 1033 | Line 1022 | discrete Fourier transforms, the particle mesh-based m
1022   reciprocal sums with rapid and absolute convergence, has proved to
1023   minimize the periodicity artifacts in liquid simulations. Taking the
1024   advantages of the fast Fourier transform (FFT) for calculating
1025 < discrete Fourier transforms, the particle mesh-based methods are
1026 < accelerated from $O(N^{3/2})$ to $O(N logN)$. An alternative
1027 < approach is \emph{fast multipole method}, which treats Coulombic
1028 < interaction exactly at short range, and approximate the potential at
1029 < long range through multipolar expansion. In spite of their wide
1030 < acceptances at the molecular simulation community, these two methods
1031 < are hard to be implemented correctly and efficiently. Instead, we
1032 < use a damped and charge-neutralized Coulomb potential method
1033 < developed by Wolf and his coworkers. The shifted Coulomb potential
1034 < for particle $i$ and particle $j$ at distance $r_{rj}$ is given by:
1025 > discrete Fourier transforms, the particle mesh-based
1026 > methods\cite{Hockney1981,Shimada1993, Luty1994} are accelerated from
1027 > $O(N^{3/2})$ to $O(N logN)$. An alternative approach is \emph{fast
1028 > multipole method}\cite{Greengard1987, Greengard1994}, which treats
1029 > Coulombic interaction exactly at short range, and approximate the
1030 > potential at long range through multipolar expansion. In spite of
1031 > their wide acceptances at the molecular simulation community, these
1032 > two methods are hard to be implemented correctly and efficiently.
1033 > Instead, we use a damped and charge-neutralized Coulomb potential
1034 > method developed by Wolf and his coworkers\cite{Wolf1999}. The
1035 > shifted Coulomb potential for particle $i$ and particle $j$ at
1036 > distance $r_{rj}$ is given by:
1037   \begin{equation}
1038   V(r_{ij})= \frac{q_i q_j \textrm{erfc}(\alpha
1039   r_{ij})}{r_{ij}}-\lim_{r_{ij}\rightarrow
# Line 1052 | Line 1043 | efficient and easy to implement.
1043   where $\alpha$ is the convergence parameter. Due to the lack of
1044   inherent periodicity and rapid convergence,this method is extremely
1045   efficient and easy to implement.
1046 < %\begin{figure}
1047 < %\centering
1048 < %\includegraphics[width=\linewidth]{pbcFig.eps}
1049 < %\caption[An illustration of shifted Coulomb potential]{An illustration of shifted Coulomb potential.}
1050 < %\label{introFigure:shiftedCoulomb}
1051 < %\end{figure}
1046 > \begin{figure}
1047 > \centering
1048 > \includegraphics[width=\linewidth]{shifted_coulomb.eps}
1049 > \caption[An illustration of shifted Coulomb potential]{An
1050 > illustration of shifted Coulomb potential.}
1051 > \label{introFigure:shiftedCoulomb}
1052 > \end{figure}
1053  
1054   %multiple time step
1055  
# Line 1073 | Line 1065 | from the trajectories.
1065   parameters, and investigate time-dependent processes of the molecule
1066   from the trajectories.
1067  
1068 < \subsubsection{\label{introSection:thermodynamicsProperties}Thermodynamics Properties}
1068 > \subsubsection{\label{introSection:thermodynamicsProperties}\textbf{Thermodynamics Properties}}
1069  
1070   Thermodynamics properties, which can be expressed in terms of some
1071   function of the coordinates and momenta of all particles in the
# Line 1095 | Line 1087 | P = \frac{{Nk_B T}}{V} - \frac{1}{{3V}}\left\langle {\
1087   < j} {r{}_{ij} \cdot f_{ij} } } \right\rangle
1088   \end{equation}
1089  
1090 < \subsubsection{\label{introSection:structuralProperties}Structural Properties}
1090 > \subsubsection{\label{introSection:structuralProperties}\textbf{Structural Properties}}
1091  
1092   Structural Properties of a simple fluid can be described by a set of
1093   distribution functions. Among these functions,\emph{pair
# Line 1103 | Line 1095 | integrating over the surface factor \cite{Powles73}. T
1095   function}, is of most fundamental importance to liquid-state theory.
1096   Pair distribution function can be gathered by Fourier transforming
1097   raw data from a series of neutron diffraction experiments and
1098 < integrating over the surface factor \cite{Powles73}. The experiment
1099 < result can serve as a criterion to justify the correctness of the
1100 < theory. Moreover, various equilibrium thermodynamic and structural
1101 < properties can also be expressed in terms of radial distribution
1102 < function \cite{allen87:csl}.
1098 > integrating over the surface factor \cite{Powles1973}. The
1099 > experiment result can serve as a criterion to justify the
1100 > correctness of the theory. Moreover, various equilibrium
1101 > thermodynamic and structural properties can also be expressed in
1102 > terms of radial distribution function \cite{Allen1987}.
1103  
1104   A pair distribution functions $g(r)$ gives the probability that a
1105   particle $i$ will be located at a distance $r$ from a another
# Line 1135 | Line 1127 | other is essentially zero.
1127   %\label{introFigure:pairDistributionFunction}
1128   %\end{figure}
1129  
1130 < \subsubsection{\label{introSection:timeDependentProperties}Time-dependent
1131 < Properties}
1130 > \subsubsection{\label{introSection:timeDependentProperties}\textbf{Time-dependent
1131 > Properties}}
1132  
1133   Time-dependent properties are usually calculated using \emph{time
1134   correlation function}, which correlates random variables $A$ and $B$
# Line 1185 | Line 1177 | protein-protein docking study{\cite{Gray03}}.
1177   movement of the objects in 3D gaming engine or other physics
1178   simulator is governed by the rigid body dynamics. In molecular
1179   simulation, rigid body is used to simplify the model in
1180 < protein-protein docking study{\cite{Gray03}}.
1180 > protein-protein docking study\cite{Gray2003}.
1181  
1182   It is very important to develop stable and efficient methods to
1183   integrate the equations of motion of orientational degrees of
# Line 1193 | Line 1185 | different sets of Euler angles can overcome this diffi
1185   rotational degrees of freedom. However, due to its singularity, the
1186   numerical integration of corresponding equations of motion is very
1187   inefficient and inaccurate. Although an alternative integrator using
1188 < different sets of Euler angles can overcome this difficulty\cite{},
1189 < the computational penalty and the lost of angular momentum
1190 < conservation still remain. A singularity free representation
1191 < utilizing quaternions was developed by Evans in 1977. Unfortunately,
1192 < this approach suffer from the nonseparable Hamiltonian resulted from
1193 < quaternion representation, which prevents the symplectic algorithm
1194 < to be utilized. Another different approach is to apply holonomic
1195 < constraints to the atoms belonging to the rigid body. Each atom
1196 < moves independently under the normal forces deriving from potential
1197 < energy and constraint forces which are used to guarantee the
1198 < rigidness. However, due to their iterative nature, SHAKE and Rattle
1199 < algorithm converge very slowly when the number of constraint
1200 < increases.
1188 > different sets of Euler angles can overcome this
1189 > difficulty\cite{Barojas1973}, the computational penalty and the lost
1190 > of angular momentum conservation still remain. A singularity free
1191 > representation utilizing quaternions was developed by Evans in
1192 > 1977\cite{Evans1977}. Unfortunately, this approach suffer from the
1193 > nonseparable Hamiltonian resulted from quaternion representation,
1194 > which prevents the symplectic algorithm to be utilized. Another
1195 > different approach is to apply holonomic constraints to the atoms
1196 > belonging to the rigid body. Each atom moves independently under the
1197 > normal forces deriving from potential energy and constraint forces
1198 > which are used to guarantee the rigidness. However, due to their
1199 > iterative nature, SHAKE and Rattle algorithm converge very slowly
1200 > when the number of constraint increases\cite{Ryckaert1977,
1201 > Andersen1983}.
1202  
1203   The break through in geometric literature suggests that, in order to
1204   develop a long-term integration scheme, one should preserve the
1205   symplectic structure of the flow. Introducing conjugate momentum to
1206   rotation matrix $Q$ and re-formulating Hamiltonian's equation, a
1207 < symplectic integrator, RSHAKE, was proposed to evolve the
1208 < Hamiltonian system in a constraint manifold by iteratively
1207 > symplectic integrator, RSHAKE\cite{Kol1997}, was proposed to evolve
1208 > the Hamiltonian system in a constraint manifold by iteratively
1209   satisfying the orthogonality constraint $Q_T Q = 1$. An alternative
1210 < method using quaternion representation was developed by Omelyan.
1211 < However, both of these methods are iterative and inefficient. In
1212 < this section, we will present a symplectic Lie-Poisson integrator
1213 < for rigid body developed by Dullweber and his
1214 < coworkers\cite{Dullweber1997} in depth.
1210 > method using quaternion representation was developed by
1211 > Omelyan\cite{Omelyan1998}. However, both of these methods are
1212 > iterative and inefficient. In this section, we will present a
1213 > symplectic Lie-Poisson integrator for rigid body developed by
1214 > Dullweber and his coworkers\cite{Dullweber1997} in depth.
1215  
1216   \subsection{\label{introSection:constrainedHamiltonianRB}Constrained Hamiltonian for Rigid Body}
1217   The motion of the rigid body is Hamiltonian with the Hamiltonian
# Line 1250 | Line 1243 | the equations of motion,
1243   Using Equation (\ref{introEquation:motionHamiltonianCoordinate},
1244   \ref{introEquation:motionHamiltonianMomentum}), one can write down
1245   the equations of motion,
1246 < \[
1247 < \begin{array}{c}
1248 < \frac{{dq}}{{dt}} = \frac{p}{m} \label{introEquation:RBMotionPosition}\\
1249 < \frac{{dp}}{{dt}} =  - \nabla _q V(q,Q) \label{introEquation:RBMotionMomentum}\\
1250 < \frac{{dQ}}{{dt}} = PJ^{ - 1}  \label{introEquation:RBMotionRotation}\\
1251 < \frac{{dP}}{{dt}} =  - \nabla _Q V(q,Q) - 2Q\Lambda . \label{introEquation:RBMotionP}\\
1252 < \end{array}
1260 < \]
1246 >
1247 > \begin{eqnarray}
1248 > \frac{{dq}}{{dt}} & = & \frac{p}{m} \label{introEquation:RBMotionPosition}\\
1249 > \frac{{dp}}{{dt}} & = & - \nabla _q V(q,Q) \label{introEquation:RBMotionMomentum}\\
1250 > \frac{{dQ}}{{dt}} & = & PJ^{ - 1}  \label{introEquation:RBMotionRotation}\\
1251 > \frac{{dP}}{{dt}} & = & - \nabla _Q V(q,Q) - 2Q\Lambda . \label{introEquation:RBMotionP}
1252 > \end{eqnarray}
1253  
1254   In general, there are two ways to satisfy the holonomic constraints.
1255   We can use constraint force provided by lagrange multiplier on the
# Line 1338 | Line 1330 | operations
1330   \[
1331   \hat vu = v \times u
1332   \]
1341
1333   Using \ref{introEqaution:RBMotionPI}, one can construct a skew
1334   matrix,
1335   \begin{equation}
1336 < (\mathop \Pi \limits^ \bullet   - \mathop \Pi \limits^ \bullet  ^T
1336 > (\mathop \Pi \limits^ \bullet   - \mathop \Pi \limits^ {\bullet  ^T}
1337   ){\rm{ }} = {\rm{ }}(\Pi  - \Pi ^T ){\rm{ }}(J^{ - 1} \Pi  + \Pi J^{
1338   - 1} ) + \sum\limits_i {[Q^T F_i (r,Q)X_i^T  - X_i F_i (r,Q)^T Q]} -
1339   (\Lambda  - \Lambda ^T ) . \label{introEquation:skewMatrixPI}
# Line 1351 | Line 1342 | not be avoided in other methods\cite{}.
1342   \ref{introEquation:skewMatrixPI} is zero, which implies the Lagrange
1343   multiplier $\Lambda$ is absent from the equations of motion. This
1344   unique property eliminate the requirement of iterations which can
1345 < not be avoided in other methods\cite{}.
1345 > not be avoided in other methods\cite{Kol1997, Omelyan1998}.
1346  
1347   Applying hat-map isomorphism, we obtain the equation of motion for
1348   angular momentum on body frame
# Line 1371 | Line 1362 | first term of \ref{ introEquation:bodyAngularMotion}).
1362  
1363   If there is not external forces exerted on the rigid body, the only
1364   contribution to the rotational is from the kinetic potential (the
1365 < first term of \ref{ introEquation:bodyAngularMotion}). The free
1366 < rigid body is an example of Lie-Poisson system with Hamiltonian
1376 < function
1365 > first term of \ref{introEquation:bodyAngularMotion}). The free rigid
1366 > body is an example of Lie-Poisson system with Hamiltonian function
1367   \begin{equation}
1368   T^r (\pi ) = T_1 ^r (\pi _1 ) + T_2^r (\pi _2 ) + T_3^r (\pi _3 )
1369   \label{introEquation:rotationalKineticRB}
# Line 1618 | Line 1608 | Operator. Below are some important properties of Lapla
1608   \]
1609   where  $p$ is real and  $L$ is called the Laplace Transform
1610   Operator. Below are some important properties of Laplace transform
1611 < \begin{equation}
1612 < \begin{array}{c}
1613 < L(x + y) = L(x) + L(y) \\
1614 < L(ax) = aL(x) \\
1615 < L(\dot x) = pL(x) - px(0) \\
1616 < L(\ddot x) = p^2 L(x) - px(0) - \dot x(0) \\
1617 < L\left( {\int_0^t {g(t - \tau )h(\tau )d\tau } } \right) = G(p)H(p) \\
1618 < \end{array}
1629 < \end{equation}
1611 >
1612 > \begin{eqnarray*}
1613 > L(x + y)  & = & L(x) + L(y) \\
1614 > L(ax)     & = & aL(x) \\
1615 > L(\dot x) & = & pL(x) - px(0) \\
1616 > L(\ddot x)& = & p^2 L(x) - px(0) - \dot x(0) \\
1617 > L\left( {\int_0^t {g(t - \tau )h(\tau )d\tau } } \right)& = & G(p)H(p) \\
1618 > \end{eqnarray*}
1619  
1620 +
1621   Applying Laplace transform to the bath coordinates, we obtain
1622 < \[
1623 < \begin{array}{c}
1624 < p^2 L(x_\alpha  ) - px_\alpha  (0) - \dot x_\alpha  (0) =  - \omega _\alpha ^2 L(x_\alpha  ) + \frac{{g_\alpha  }}{{\omega _\alpha  }}L(x) \\
1625 < L(x_\alpha  ) = \frac{{\frac{{g_\alpha  }}{{\omega _\alpha  }}L(x) + px_\alpha  (0) + \dot x_\alpha  (0)}}{{p^2  + \omega _\alpha ^2 }} \\
1626 < \end{array}
1637 < \]
1622 > \begin{eqnarray*}
1623 > 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) \\
1624 > L(x_\alpha  ) & = & \frac{{\frac{{g_\alpha  }}{{\omega _\alpha  }}L(x) + px_\alpha  (0) + \dot x_\alpha  (0)}}{{p^2  + \omega _\alpha ^2 }} \\
1625 > \end{eqnarray*}
1626 >
1627   By the same way, the system coordinates become
1628 < \[
1629 < \begin{array}{c}
1630 < mL(\ddot x) =  - \frac{1}{p}\frac{{\partial W(x)}}{{\partial x}} \\
1631 <  - \sum\limits_{\alpha  = 1}^N {\left\{ { - \frac{{g_\alpha ^2 }}{{m_\alpha  \omega _\alpha ^2 }}\frac{p}{{p^2  + \omega _\alpha ^2 }}pL(x) - \frac{p}{{p^2  + \omega _\alpha ^2 }}g_\alpha  x_\alpha  (0) - \frac{1}{{p^2  + \omega _\alpha ^2 }}g_\alpha  \dot x_\alpha  (0)} \right\}}  \\
1643 < \end{array}
1644 < \]
1628 > \begin{eqnarray*}
1629 > mL(\ddot x) & = & - \frac{1}{p}\frac{{\partial W(x)}}{{\partial x}} \\
1630 >  & & \mbox{} - \sum\limits_{\alpha  = 1}^N {\left\{ { - \frac{{g_\alpha ^2 }}{{m_\alpha  \omega _\alpha ^2 }}\frac{p}{{p^2  + \omega _\alpha ^2 }}pL(x) - \frac{p}{{p^2  + \omega _\alpha ^2 }}g_\alpha  x_\alpha  (0) - \frac{1}{{p^2  + \omega _\alpha ^2 }}g_\alpha  \dot x_\alpha  (0)} \right\}}  \\
1631 > \end{eqnarray*}
1632  
1633   With the help of some relatively important inverse Laplace
1634   transformations:
# Line 1653 | Line 1640 | transformations:
1640   \end{array}
1641   \]
1642   , we obtain
1643 < \begin{align}
1644 < m\ddot x &=  - \frac{{\partial W(x)}}{{\partial x}} -
1643 > \begin{eqnarray*}
1644 > m\ddot x & =  & - \frac{{\partial W(x)}}{{\partial x}} -
1645   \sum\limits_{\alpha  = 1}^N {\left\{ {\left( { - \frac{{g_\alpha ^2
1646   }}{{m_\alpha  \omega _\alpha ^2 }}} \right)\int_0^t {\cos (\omega
1647 < _\alpha  t)\dot x(t - \tau )d\tau  - \left[ {g_\alpha  x_\alpha  (0)
1648 < - \frac{{g_\alpha  }}{{m_\alpha  \omega _\alpha  }}} \right]\cos
1649 < (\omega _\alpha  t) - \frac{{g_\alpha  \dot x_\alpha  (0)}}{{\omega
1650 < _\alpha  }}\sin (\omega _\alpha  t)} } \right\}}
1651 < %
1652 < &= - \frac{{\partial W(x)}}{{\partial x}} - \int_0^t
1647 > _\alpha  t)\dot x(t - \tau )d\tau } } \right\}}  \\
1648 > & & + \sum\limits_{\alpha  = 1}^N {\left\{ {\left[ {g_\alpha
1649 > x_\alpha (0) - \frac{{g_\alpha  }}{{m_\alpha  \omega _\alpha  }}}
1650 > \right]\cos (\omega _\alpha  t) + \frac{{g_\alpha  \dot x_\alpha
1651 > (0)}}{{\omega _\alpha  }}\sin (\omega _\alpha  t)} \right\}}
1652 > \end{eqnarray*}
1653 > \begin{eqnarray*}
1654 > m\ddot x & = & - \frac{{\partial W(x)}}{{\partial x}} - \int_0^t
1655   {\sum\limits_{\alpha  = 1}^N {\left( { - \frac{{g_\alpha ^2
1656   }}{{m_\alpha  \omega _\alpha ^2 }}} \right)\cos (\omega _\alpha
1657 < t)\dot x(t - \tau )d} \tau }  + \sum\limits_{\alpha  = 1}^N {\left\{
1658 < {\left[ {g_\alpha  x_\alpha  (0) - \frac{{g_\alpha  }}{{m_\alpha
1659 < \omega _\alpha  }}} \right]\cos (\omega _\alpha  t) +
1660 < \frac{{g_\alpha  \dot x_\alpha  (0)}}{{\omega _\alpha  }}\sin
1661 < (\omega _\alpha  t)} \right\}}
1662 < \end{align}
1674 <
1657 > t)\dot x(t - \tau )d} \tau }  \\
1658 > & & + \sum\limits_{\alpha  = 1}^N {\left\{ {\left[ {g_\alpha
1659 > x_\alpha (0) - \frac{{g_\alpha }}{{m_\alpha \omega _\alpha  }}}
1660 > \right]\cos (\omega _\alpha  t) + \frac{{g_\alpha  \dot x_\alpha
1661 > (0)}}{{\omega _\alpha  }}\sin (\omega _\alpha  t)} \right\}}
1662 > \end{eqnarray*}
1663   Introducing a \emph{dynamic friction kernel}
1664   \begin{equation}
1665   \xi (t) = \sum\limits_{\alpha  = 1}^N {\left( { - \frac{{g_\alpha ^2
# Line 1694 | Line 1682 | which is known as the \emph{generalized Langevin equat
1682   \end{equation}
1683   which is known as the \emph{generalized Langevin equation}.
1684  
1685 < \subsubsection{\label{introSection:randomForceDynamicFrictionKernel}Random Force and Dynamic Friction Kernel}
1685 > \subsubsection{\label{introSection:randomForceDynamicFrictionKernel}\textbf{Random Force and Dynamic Friction Kernel}}
1686  
1687   One may notice that $R(t)$ depends only on initial conditions, which
1688   implies it is completely deterministic within the context of a
# Line 1749 | Line 1737 | or be determined by Stokes' law for regular shaped par
1737   \end{equation}
1738   which is known as the Langevin equation. The static friction
1739   coefficient $\xi _0$ can either be calculated from spectral density
1740 < or be determined by Stokes' law for regular shaped particles.A
1740 > or be determined by Stokes' law for regular shaped particles. A
1741   briefly review on calculating friction tensor for arbitrary shaped
1742   particles is given in Sec.~\ref{introSection:frictionTensor}.
1743  
1744 < \subsubsection{\label{introSection:secondFluctuationDissipation}The Second Fluctuation Dissipation Theorem}
1744 > \subsubsection{\label{introSection:secondFluctuationDissipation}\textbf{The Second Fluctuation Dissipation Theorem}}
1745  
1746   Defining a new set of coordinates,
1747   \[
# Line 1765 | Line 1753 | And since the $q$ coordinates are harmonic oscillators
1753   R(t) = \sum\limits_{\alpha  = 1}^N {g_\alpha  q_\alpha  (t)}.
1754   \]
1755   And since the $q$ coordinates are harmonic oscillators,
1756 < \[
1757 < \begin{array}{c}
1758 < \left\langle {q_\alpha ^2 } \right\rangle  = \frac{{kT}}{{m_\alpha  \omega _\alpha ^2 }} \\
1759 < \left\langle {q_\alpha  (t)q_\alpha  (0)} \right\rangle  = \left\langle {q_\alpha ^2 (0)} \right\rangle \cos (\omega _\alpha  t) \\
1760 < \left\langle {q_\alpha  (t)q_\beta  (0)} \right\rangle  = \delta _{\alpha \beta } \left\langle {q_\alpha  (t)q_\alpha  (0)} \right\rangle  \\
1761 < \left\langle {R(t)R(0)} \right\rangle  = \sum\limits_\alpha  {\sum\limits_\beta  {g_\alpha  g_\beta  \left\langle {q_\alpha  (t)q_\beta  (0)} \right\rangle } }  \\
1762 <  = \sum\limits_\alpha  {g_\alpha ^2 \left\langle {q_\alpha ^2 (0)} \right\rangle \cos (\omega _\alpha  t)}  \\
1763 <  = kT\xi (t) \\
1764 < \end{array}
1765 < \]
1756 >
1757 > \begin{eqnarray*}
1758 > \left\langle {q_\alpha ^2 } \right\rangle  & = & \frac{{kT}}{{m_\alpha  \omega _\alpha ^2 }} \\
1759 > \left\langle {q_\alpha  (t)q_\alpha  (0)} \right\rangle & = & \left\langle {q_\alpha ^2 (0)} \right\rangle \cos (\omega _\alpha  t) \\
1760 > \left\langle {q_\alpha  (t)q_\beta  (0)} \right\rangle & = &\delta _{\alpha \beta } \left\langle {q_\alpha  (t)q_\alpha  (0)} \right\rangle  \\
1761 > \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 } }  \\
1762 >  & = &\sum\limits_\alpha  {g_\alpha ^2 \left\langle {q_\alpha ^2 (0)} \right\rangle \cos (\omega _\alpha  t)}  \\
1763 >  & = &kT\xi (t) \\
1764 > \end{eqnarray*}
1765 >
1766   Thus, we recover the \emph{second fluctuation dissipation theorem}
1767   \begin{equation}
1768   \xi (t) = \left\langle {R(t)R(0)} \right\rangle
# Line 1782 | Line 1770 | can model the random force and friction kernel.
1770   \end{equation}
1771   In effect, it acts as a constraint on the possible ways in which one
1772   can model the random force and friction kernel.
1785
1786 \subsection{\label{introSection:frictionTensor} Friction Tensor}
1787 Theoretically, the friction kernel can be determined using velocity
1788 autocorrelation function. However, this approach become impractical
1789 when the system become more and more complicate. Instead, various
1790 approaches based on hydrodynamics have been developed to calculate
1791 the friction coefficients. The friction effect is isotropic in
1792 Equation, $\zeta$ can be taken as a scalar. In general, friction
1793 tensor $\Xi$ is a $6\times 6$ matrix given by
1794 \[
1795 \Xi  = \left( {\begin{array}{*{20}c}
1796   {\Xi _{}^{tt} } & {\Xi _{}^{rt} }  \\
1797   {\Xi _{}^{tr} } & {\Xi _{}^{rr} }  \\
1798 \end{array}} \right).
1799 \]
1800 Here, $ {\Xi^{tt} }$ and $ {\Xi^{rr} }$ are translational friction
1801 tensor and rotational resistance (friction) tensor respectively,
1802 while ${\Xi^{tr} }$ is translation-rotation coupling tensor and $
1803 {\Xi^{rt} }$ is rotation-translation coupling tensor. When a
1804 particle moves in a fluid, it may experience friction force or
1805 torque along the opposite direction of the velocity or angular
1806 velocity,
1807 \[
1808 \left( \begin{array}{l}
1809 F_R  \\
1810 \tau _R  \\
1811 \end{array} \right) =  - \left( {\begin{array}{*{20}c}
1812   {\Xi ^{tt} } & {\Xi ^{rt} }  \\
1813   {\Xi ^{tr} } & {\Xi ^{rr} }  \\
1814 \end{array}} \right)\left( \begin{array}{l}
1815 v \\
1816 w \\
1817 \end{array} \right)
1818 \]
1819 where $F_r$ is the friction force and $\tau _R$ is the friction
1820 toque.
1821
1822 \subsubsection{\label{introSection:resistanceTensorRegular}The Resistance Tensor for Regular Shape}
1823
1824 For a spherical particle, the translational and rotational friction
1825 constant can be calculated from Stoke's law,
1826 \[
1827 \Xi ^{tt}  = \left( {\begin{array}{*{20}c}
1828   {6\pi \eta R} & 0 & 0  \\
1829   0 & {6\pi \eta R} & 0  \\
1830   0 & 0 & {6\pi \eta R}  \\
1831 \end{array}} \right)
1832 \]
1833 and
1834 \[
1835 \Xi ^{rr}  = \left( {\begin{array}{*{20}c}
1836   {8\pi \eta R^3 } & 0 & 0  \\
1837   0 & {8\pi \eta R^3 } & 0  \\
1838   0 & 0 & {8\pi \eta R^3 }  \\
1839 \end{array}} \right)
1840 \]
1841 where $\eta$ is the viscosity of the solvent and $R$ is the
1842 hydrodynamics radius.
1843
1844 Other non-spherical shape, such as cylinder and ellipsoid
1845 \textit{etc}, are widely used as reference for developing new
1846 hydrodynamics theory, because their properties can be calculated
1847 exactly. In 1936, Perrin extended Stokes's law to general ellipsoid,
1848 also called a triaxial ellipsoid, which is given in Cartesian
1849 coordinates by
1850 \[
1851 \frac{{x^2 }}{{a^2 }} + \frac{{y^2 }}{{b^2 }} + \frac{{z^2 }}{{c^2
1852 }} = 1
1853 \]
1854 where the semi-axes are of lengths $a$, $b$, and $c$. Unfortunately,
1855 due to the complexity of the elliptic integral, only the ellipsoid
1856 with the restriction of two axes having to be equal, \textit{i.e.}
1857 prolate($ a \ge b = c$) and oblate ($ a < b = c $), can be solved
1858 exactly. Introducing an elliptic integral parameter $S$ for prolate,
1859 \[
1860 S = \frac{2}{{\sqrt {a^2  - b^2 } }}\ln \frac{{a + \sqrt {a^2  - b^2
1861 } }}{b},
1862 \]
1863 and oblate,
1864 \[
1865 S = \frac{2}{{\sqrt {b^2  - a^2 } }}arctg\frac{{\sqrt {b^2  - a^2 }
1866 }}{a}
1867 \],
1868 one can write down the translational and rotational resistance
1869 tensors
1870 \[
1871 \begin{array}{l}
1872 \Xi _a^{tt}  = 16\pi \eta \frac{{a^2  - b^2 }}{{(2a^2  - b^2 )S - 2a}} \\
1873 \Xi _b^{tt}  = \Xi _c^{tt}  = 32\pi \eta \frac{{a^2  - b^2 }}{{(2a^2  - 3b^2 )S + 2a}} \\
1874 \end{array},
1875 \]
1876 and
1877 \[
1878 \begin{array}{l}
1879 \Xi _a^{rr}  = \frac{{32\pi }}{3}\eta \frac{{(a^2  - b^2 )b^2 }}{{2a - b^2 S}} \\
1880 \Xi _b^{rr}  = \Xi _c^{rr}  = \frac{{32\pi }}{3}\eta \frac{{(a^4  - b^4 )}}{{(2a^2  - b^2 )S - 2a}} \\
1881 \end{array}.
1882 \]
1883
1884 \subsubsection{\label{introSection:resistanceTensorRegularArbitrary}The Resistance Tensor for Arbitrary Shape}
1885
1886 Unlike spherical and other regular shaped molecules, there is not
1887 analytical solution for friction tensor of any arbitrary shaped
1888 rigid molecules. The ellipsoid of revolution model and general
1889 triaxial ellipsoid model have been used to approximate the
1890 hydrodynamic properties of rigid bodies. However, since the mapping
1891 from all possible ellipsoidal space, $r$-space, to all possible
1892 combination of rotational diffusion coefficients, $D$-space is not
1893 unique\cite{Wegener79} as well as the intrinsic coupling between
1894 translational and rotational motion of rigid body\cite{}, general
1895 ellipsoid is not always suitable for modeling arbitrarily shaped
1896 rigid molecule. A number of studies have been devoted to determine
1897 the friction tensor for irregularly shaped rigid bodies using more
1898 advanced method\cite{} where the molecule of interest was modeled by
1899 combinations of spheres(beads)\cite{} and the hydrodynamics
1900 properties of the molecule can be calculated using the hydrodynamic
1901 interaction tensor. Let us consider a rigid assembly of $N$ beads
1902 immersed in a continuous medium. Due to hydrodynamics interaction,
1903 the ``net'' velocity of $i$th bead, $v'_i$ is different than its
1904 unperturbed velocity $v_i$,
1905 \[
1906 v'_i  = v_i  - \sum\limits_{j \ne i} {T_{ij} F_j }
1907 \]
1908 where $F_i$ is the frictional force, and $T_{ij}$ is the
1909 hydrodynamic interaction tensor. The friction force of $i$th bead is
1910 proportional to its ``net'' velocity
1911 \begin{equation}
1912 F_i  = \zeta _i v_i  - \zeta _i \sum\limits_{j \ne i} {T_{ij} F_j }.
1913 \label{introEquation:tensorExpression}
1914 \end{equation}
1915 This equation is the basis for deriving the hydrodynamic tensor. In
1916 1930, Oseen and Burgers gave a simple solution to Equation
1917 \ref{introEquation:tensorExpression}
1918 \begin{equation}
1919 T_{ij}  = \frac{1}{{8\pi \eta r_{ij} }}\left( {I + \frac{{R_{ij}
1920 R_{ij}^T }}{{R_{ij}^2 }}} \right).
1921 \label{introEquation:oseenTensor}
1922 \end{equation}
1923 Here $R_{ij}$ is the distance vector between bead $i$ and bead $j$.
1924 A second order expression for element of different size was
1925 introduced by Rotne and Prager\cite{} and improved by Garc\'{i}a de
1926 la Torre and Bloomfield,
1927 \begin{equation}
1928 T_{ij}  = \frac{1}{{8\pi \eta R_{ij} }}\left[ {\left( {I +
1929 \frac{{R_{ij} R_{ij}^T }}{{R_{ij}^2 }}} \right) + R\frac{{\sigma
1930 _i^2  + \sigma _j^2 }}{{r_{ij}^2 }}\left( {\frac{I}{3} -
1931 \frac{{R_{ij} R_{ij}^T }}{{R_{ij}^2 }}} \right)} \right].
1932 \label{introEquation:RPTensorNonOverlapped}
1933 \end{equation}
1934 Both of the Equation \ref{introEquation:oseenTensor} and Equation
1935 \ref{introEquation:RPTensorNonOverlapped} have an assumption $R_{ij}
1936 \ge \sigma _i  + \sigma _j$. An alternative expression for
1937 overlapping beads with the same radius, $\sigma$, is given by
1938 \begin{equation}
1939 T_{ij}  = \frac{1}{{6\pi \eta R_{ij} }}\left[ {\left( {1 -
1940 \frac{2}{{32}}\frac{{R_{ij} }}{\sigma }} \right)I +
1941 \frac{2}{{32}}\frac{{R_{ij} R_{ij}^T }}{{R_{ij} \sigma }}} \right]
1942 \label{introEquation:RPTensorOverlapped}
1943 \end{equation}
1944
1945 To calculate the resistance tensor at an arbitrary origin $O$, we
1946 construct a $3N \times 3N$ matrix consisting of $N \times N$
1947 $B_{ij}$ blocks
1948 \begin{equation}
1949 B = \left( {\begin{array}{*{20}c}
1950   {B_{11} } &  \ldots  & {B_{1N} }  \\
1951    \vdots  &  \ddots  &  \vdots   \\
1952   {B_{N1} } &  \cdots  & {B_{NN} }  \\
1953 \end{array}} \right),
1954 \end{equation}
1955 where $B_{ij}$ is given by
1956 \[
1957 B_{ij}  = \delta _{ij} \frac{I}{{6\pi \eta R}} + (1 - \delta _{ij}
1958 )T_{ij}
1959 \]
1960 where $\delta _{ij}$ is Kronecker delta function. Inverting matrix
1961 $B$, we obtain
1962
1963 \[
1964 C = B^{ - 1}  = \left( {\begin{array}{*{20}c}
1965   {C_{11} } &  \ldots  & {C_{1N} }  \\
1966    \vdots  &  \ddots  &  \vdots   \\
1967   {C_{N1} } &  \cdots  & {C_{NN} }  \\
1968 \end{array}} \right)
1969 \]
1970 , which can be partitioned into $N \times N$ $3 \times 3$ block
1971 $C_{ij}$. With the help of $C_{ij}$ and skew matrix $U_i$
1972 \[
1973 U_i  = \left( {\begin{array}{*{20}c}
1974   0 & { - z_i } & {y_i }  \\
1975   {z_i } & 0 & { - x_i }  \\
1976   { - y_i } & {x_i } & 0  \\
1977 \end{array}} \right)
1978 \]
1979 where $x_i$, $y_i$, $z_i$ are the components of the vector joining
1980 bead $i$ and origin $O$. Hence, the elements of resistance tensor at
1981 arbitrary origin $O$ can be written as
1982 \begin{equation}
1983 \begin{array}{l}
1984 \Xi _{}^{tt}  = \sum\limits_i {\sum\limits_j {C_{ij} } } , \\
1985 \Xi _{}^{tr}  = \Xi _{}^{rt}  = \sum\limits_i {\sum\limits_j {U_i C_{ij} } } , \\
1986 \Xi _{}^{rr}  =  - \sum\limits_i {\sum\limits_j {U_i C_{ij} } } U_j  \\
1987 \end{array}
1988 \label{introEquation:ResistanceTensorArbitraryOrigin}
1989 \end{equation}
1990
1991 The resistance tensor depends on the origin to which they refer. The
1992 proper location for applying friction force is the center of
1993 resistance (reaction), at which the trace of rotational resistance
1994 tensor, $ \Xi ^{rr}$ reaches minimum. Mathematically, the center of
1995 resistance is defined as an unique point of the rigid body at which
1996 the translation-rotation coupling tensor are symmetric,
1997 \begin{equation}
1998 \Xi^{tr}  = \left( {\Xi^{tr} } \right)^T
1999 \label{introEquation:definitionCR}
2000 \end{equation}
2001 Form Equation \ref{introEquation:ResistanceTensorArbitraryOrigin},
2002 we can easily find out that the translational resistance tensor is
2003 origin independent, while the rotational resistance tensor and
2004 translation-rotation coupling resistance tensor depend on the
2005 origin. Given resistance tensor at an arbitrary origin $O$, and a
2006 vector ,$r_{OP}(x_{OP}, y_{OP}, z_{OP})$, from $O$ to $P$, we can
2007 obtain the resistance tensor at $P$ by
2008 \begin{equation}
2009 \begin{array}{l}
2010 \Xi _P^{tt}  = \Xi _O^{tt}  \\
2011 \Xi _P^{tr}  = \Xi _P^{rt}  = \Xi _O^{tr}  - U_{OP} \Xi _O^{tt}  \\
2012 \Xi _P^{rr}  = \Xi _O^{rr}  - U_{OP} \Xi _O^{tt} U_{OP}  + \Xi _O^{tr} U_{OP}  - U_{OP} \Xi _O^{tr} ^{^T }  \\
2013 \end{array}
2014 \label{introEquation:resistanceTensorTransformation}
2015 \end{equation}
2016 where
2017 \[
2018 U_{OP}  = \left( {\begin{array}{*{20}c}
2019   0 & { - z_{OP} } & {y_{OP} }  \\
2020   {z_i } & 0 & { - x_{OP} }  \\
2021   { - y_{OP} } & {x_{OP} } & 0  \\
2022 \end{array}} \right)
2023 \]
2024 Using Equations \ref{introEquation:definitionCR} and
2025 \ref{introEquation:resistanceTensorTransformation}, one can locate
2026 the position of center of resistance,
2027 \[
2028 \left( \begin{array}{l}
2029 x_{OR}  \\
2030 y_{OR}  \\
2031 z_{OR}  \\
2032 \end{array} \right) = \left( {\begin{array}{*{20}c}
2033   {(\Xi _O^{rr} )_{yy}  + (\Xi _O^{rr} )_{zz} } & { - (\Xi _O^{rr} )_{xy} } & { - (\Xi _O^{rr} )_{xz} }  \\
2034   { - (\Xi _O^{rr} )_{xy} } & {(\Xi _O^{rr} )_{zz}  + (\Xi _O^{rr} )_{xx} } & { - (\Xi _O^{rr} )_{yz} }  \\
2035   { - (\Xi _O^{rr} )_{xz} } & { - (\Xi _O^{rr} )_{yz} } & {(\Xi _O^{rr} )_{xx}  + (\Xi _O^{rr} )_{yy} }  \\
2036 \end{array}} \right)^{ - 1} \left( \begin{array}{l}
2037 (\Xi _O^{tr} )_{yz}  - (\Xi _O^{tr} )_{zy}  \\
2038 (\Xi _O^{tr} )_{zx}  - (\Xi _O^{tr} )_{xz}  \\
2039 (\Xi _O^{tr} )_{xy}  - (\Xi _O^{tr} )_{yx}  \\
2040 \end{array} \right).
2041 \]
2042 where $x_OR$, $y_OR$, $z_OR$ are the components of the vector
2043 joining center of resistance $R$ and origin $O$.

Diff Legend

Removed lines
+ Added lines
< Changed lines
> Changed lines