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 2720 by tim, Wed Apr 19 19:46:53 2006 UTC vs.
Revision 2882 by tim, Fri Jun 23 21:33:52 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 603 | Line 591 | Instead, we use a approximate map, $\psi_\tau$, which
591   \end{equation}
592  
593   In most cases, it is not easy to find the exact flow $\varphi_\tau$.
594 < Instead, we use a approximate map, $\psi_\tau$, which is usually
594 > Instead, we use an approximate map, $\psi_\tau$, which is usually
595   called integrator. The order of an integrator $\psi_\tau$ is $p$, if
596   the Taylor series of $\psi_\tau$ agree to order $p$,
597   \begin{equation}
598 < \psi_tau(x) = x + \tau f(x) + O(\tau^{p+1})
598 > \psi_\tau(x) = x + \tau f(x) + O(\tau^{p+1})
599   \end{equation}
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 an
604 > ODE and its flow play important roles in numerical studies. Many of
605 > them 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 629 | Line 617 | is the property must be preserved by the integrator.
617   \begin{equation}
618   {\varphi '}^T J \varphi ' = J \circ \varphi
619   \end{equation}
620 < is the property must be preserved by the integrator.
620 > is the property that must be preserved by the integrator.
621  
622   It is possible to construct a \emph{volume-preserving} flow for a
623 < source free($ \nabla \cdot f = 0 $) ODE, if the flow satisfies $
623 > source free ODE ($ \nabla \cdot f = 0 $), if the flow satisfies $
624   \det d\varphi  = 1$. One can show easily that a symplectic flow will
625   be volume-preserving.
626  
627 < Changing the variables $y = h(x)$ in a ODE\ref{introEquation:ODE}
628 < will result in a new system,
627 > Changing the variables $y = h(x)$ in an ODE
628 > (Eq.~\ref{introEquation:ODE}) will result in a new system,
629   \[
630   \dot y = \tilde f(y) = ((dh \cdot f)h^{ - 1} )(y).
631   \]
# 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 method
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}
756  
757 < \subsubsection{\label{introSection:exampleSplittingMethod}Example of Splitting Method}
757 > \subsubsection{\label{introSection:exampleSplittingMethod}\textbf{Examples of the Splitting Method}}
758   The classical equation for a system consisting of interacting
759   particles can be written in Hamiltonian form,
760   \[
761   H = T + V
762   \]
763   where $T$ is the kinetic energy and $V$ is the potential energy.
764 < Setting $H_1 = T, H_2 = V$ and applying Strang splitting, one
764 > Setting $H_1 = T, H_2 = V$ and applying the Strang splitting, one
765   obtains the following:
766   \begin{align}
767   q(\Delta t) &= q(0) + \dot{q}(0)\Delta t +
# Line 799 | Line 788 | q(\Delta t) &= q(0) + \Delta t\, \dot{q}\biggl (\frac{
788      \label{introEquation:Lp9b}\\%
789   %
790   \dot{q}(\Delta t) &= \dot{q}\biggl (\frac{\Delta t}{2}\biggr ) +
791 <    \frac{\Delta t}{2m}\, F[q(0)]. \label{introEquation:Lp9c}
791 >    \frac{\Delta t}{2m}\, F[q(t)]. \label{introEquation:Lp9c}
792   \end{align}
793   From the preceding splitting, one can see that the integration of
794   the equations of motion would follow:
# Line 808 | Line 797 | the equations of motion would follow:
797  
798   \item Use the half step velocities to move positions one whole step, $\Delta t$.
799  
800 < \item Evaluate the forces at the new positions, $\mathbf{r}(\Delta t)$, and use the new forces to complete the velocity move.
800 > \item Evaluate the forces at the new positions, $\mathbf{q}(\Delta t)$, and use the new forces to complete the velocity move.
801  
802   \item Repeat from step 1 with the new position, velocities, and forces assuming the roles of the initial values.
803   \end{enumerate}
804  
805 < Simply switching the order of splitting and composing, a new
806 < integrator, the \emph{position verlet} integrator, can be generated,
805 > By simply switching the order of the propagators in the splitting
806 > and composing a new integrator, the \emph{position verlet}
807 > integrator, can be generated,
808   \begin{align}
809   \dot q(\Delta t) &= \dot q(0) + \Delta tF(q(0))\left[ {q(0) +
810   \frac{{\Delta t}}{{2m}}\dot q(0)} \right], %
# Line 825 | Line 815 | q(\Delta t)} \right]. %
815   \label{introEquation:positionVerlet2}
816   \end{align}
817  
818 < \subsubsection{\label{introSection:errorAnalysis}Error Analysis and Higher Order Methods}
818 > \subsubsection{\label{introSection:errorAnalysis}\textbf{Error Analysis and Higher Order Methods}}
819  
820 < Baker-Campbell-Hausdorff formula can be used to determine the local
821 < error of splitting method in terms of commutator of the
820 > The Baker-Campbell-Hausdorff formula can be used to determine the
821 > local error of splitting method in terms of the commutator of the
822   operators(\ref{introEquation:exponentialOperator}) associated with
823 < the sub-flow. For operators $hX$ and $hY$ which are associate to
824 < $\varphi_1(t)$ and $\varphi_2(t$ respectively , we have
823 > the sub-flow. For operators $hX$ and $hY$ which are associated with
824 > $\varphi_1(t)$ and $\varphi_2(t)$ respectively , we have
825   \begin{equation}
826   \exp (hX + hY) = \exp (hZ)
827   \end{equation}
# Line 844 | Line 834 | Applying Baker-Campbell-Hausdorff formula to Sprang sp
834   \[
835   [X,Y] = XY - YX .
836   \]
837 < Applying Baker-Campbell-Hausdorff formula to Sprang splitting, we
838 < can obtain
837 > Applying the Baker-Campbell-Hausdorff formula\cite{Varadarajan1974}
838 > to the Sprang splitting, we can obtain
839   \begin{eqnarray*}
840 < \exp (h X/2)\exp (h Y)\exp (h X/2) & = & \exp (h X + h Y + h^2
841 < [X,Y]/4 + h^2 [Y,X]/4 \\ & & \mbox{} + h^2 [X,X]/8 + h^2 [Y,Y]/8 \\
842 < & & \mbox{} + h^3 [Y,[Y,X]]/12 - h^3 [X,[X,Y]]/24 & & \mbox{} +
853 < \ldots )
840 > \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 \\
841 >                                   &   & \mbox{} + h^2 [X,X]/8 + h^2 [Y,Y]/8 \\
842 >                                   &   & \mbox{} + h^3 [Y,[Y,X]]/12 - h^3[X,[X,Y]]/24 + \ldots )
843   \end{eqnarray*}
844 < Since \[ [X,Y] + [Y,X] = 0\] and \[ [X,X] = 0\], the dominant local
844 > Since \[ [X,Y] + [Y,X] = 0\] and \[ [X,X] = 0,\] the dominant local
845   error of Spring splitting is proportional to $h^3$. The same
846 < procedure can be applied to general splitting,  of the form
846 > procedure can be applied to a general splitting,  of the form
847   \begin{equation}
848   \varphi _{b_m h}^2  \circ \varphi _{a_m h}^1  \circ \varphi _{b_{m -
849   1} h}^2  \circ  \ldots  \circ \varphi _{a_1 h}^1 .
850   \end{equation}
851 < Careful choice of coefficient $a_1 ,\ldot , b_m$ will lead to higher
852 < order method. Yoshida proposed an elegant way to compose higher
853 < order methods based on symmetric splitting. Given a symmetric second
854 < order base method $ \varphi _h^{(2)} $, a fourth-order symmetric
855 < method can be constructed by composing,
851 > A careful choice of coefficient $a_1 \ldots b_m$ will lead to higher
852 > order methods. Yoshida proposed an elegant way to compose higher
853 > order methods based on symmetric splitting\cite{Yoshida1990}. Given
854 > a symmetric second order base method $ \varphi _h^{(2)} $, a
855 > fourth-order symmetric method can be constructed by composing,
856   \[
857   \varphi _h^{(4)}  = \varphi _{\alpha h}^{(2)}  \circ \varphi _{\beta
858   h}^{(2)}  \circ \varphi _{\alpha h}^{(2)}
# Line 873 | Line 862 | _{\beta h}^{(2n)}  \circ \varphi _{\alpha h}^{(2n)}
862   integrator $ \varphi _h^{(2n + 2)}$ can be composed by
863   \begin{equation}
864   \varphi _h^{(2n + 2)}  = \varphi _{\alpha h}^{(2n)}  \circ \varphi
865 < _{\beta h}^{(2n)}  \circ \varphi _{\alpha h}^{(2n)}
865 > _{\beta h}^{(2n)}  \circ \varphi _{\alpha h}^{(2n)},
866   \end{equation}
867 < , if the weights are chosen as
867 > if the weights are chosen as
868   \[
869   \alpha  =  - \frac{{2^{1/(2n + 1)} }}{{2 - 2^{1/(2n + 1)} }},\beta =
870   \frac{{2^{1/(2n + 1)} }}{{2 - 2^{1/(2n + 1)} }} .
# Line 892 | Line 881 | T(t) = \sum\limits_{i = 1}^N {\frac{{m_i v_i^2 }}{{fk_
881   simulations. For instance, instantaneous temperature of an
882   Hamiltonian system of $N$ particle can be measured by
883   \[
884 < T(t) = \sum\limits_{i = 1}^N {\frac{{m_i v_i^2 }}{{fk_B }}}
884 > T = \sum\limits_{i = 1}^N {\frac{{m_i v_i^2 }}{{fk_B }}}
885   \]
886   where $m_i$ and $v_i$ are the mass and velocity of $i$th particle
887   respectively, $f$ is the number of degrees of freedom, and $k_B$ is
# Line 912 | Line 901 | initialization of a simulation. Sec.~\ref{introSec:pro
901   \end{enumerate}
902   These three individual steps will be covered in the following
903   sections. Sec.~\ref{introSec:initialSystemSettings} deals with the
904 < initialization of a simulation. Sec.~\ref{introSec:production} will
905 < discusses issues in production run, including the force evaluation
917 < and the numerical integration schemes of the equations of motion .
904 > initialization of a simulation. Sec.~\ref{introSection:production}
905 > will discusse issues in production run.
906   Sec.~\ref{introSection:Analysis} provides the theoretical tools for
907   trajectory analysis.
908  
909   \subsection{\label{introSec:initialSystemSettings}Initialization}
910  
911 < \subsubsection{Preliminary preparation}
911 > \subsubsection{\textbf{Preliminary preparation}}
912  
913   When selecting the starting structure of a molecule for molecular
914   simulation, one may retrieve its Cartesian coordinates from public
915   databases, such as RCSB Protein Data Bank \textit{etc}. Although
916   thousands of crystal structures of molecules are discovered every
917   year, many more remain unknown due to the difficulties of
918 < purification and crystallization. Even for the molecule with known
919 < structure, some important information is missing. For example, the
918 > purification and crystallization. Even for molecules with known
919 > structure, some important information is missing. For example, a
920   missing hydrogen atom which acts as donor in hydrogen bonding must
921   be added. Moreover, in order to include electrostatic interaction,
922   one may need to specify the partial charges for individual atoms.
923   Under some circumstances, we may even need to prepare the system in
924 < a special setup. For instance, when studying transport phenomenon in
925 < membrane system, we may prepare the lipids in bilayer structure
926 < instead of placing lipids randomly in solvent, since we are not
927 < interested in self-aggregation and it takes a long time to happen.
924 > a special configuration. For instance, when studying transport
925 > phenomenon in membrane systems, we may prepare the lipids in a
926 > bilayer structure instead of placing lipids randomly in solvent,
927 > since we are not interested in the slow self-aggregation process.
928  
929 < \subsubsection{Minimization}
929 > \subsubsection{\textbf{Minimization}}
930  
931   It is quite possible that some of molecules in the system from
932 < preliminary preparation may be overlapped with each other. This
933 < close proximity leads to high potential energy which consequently
934 < jeopardizes any molecular dynamics simulations. To remove these
935 < steric overlaps, one typically performs energy minimization to find
936 < a more reasonable conformation. Several energy minimization methods
937 < have been developed to exploit the energy surface and to locate the
938 < local minimum. While converging slowly near the minimum, steepest
939 < descent method is extremely robust when systems are far from
940 < harmonic. Thus, it is often used to refine structure from
941 < crystallographic data. Relied on the gradient or hessian, advanced
942 < methods like conjugate gradient and Newton-Raphson converge rapidly
943 < to a local minimum, while become unstable if the energy surface is
944 < far from quadratic. Another factor must be taken into account, when
932 > preliminary preparation may be overlapping with each other. This
933 > close proximity leads to high initial potential energy which
934 > consequently jeopardizes any molecular dynamics simulations. To
935 > remove these steric overlaps, one typically performs energy
936 > minimization to find a more reasonable conformation. Several energy
937 > minimization methods have been developed to exploit the energy
938 > surface and to locate the local minimum. While converging slowly
939 > near the minimum, steepest descent method is extremely robust when
940 > systems are strongly anharmonic. Thus, it is often used to refine
941 > structure from crystallographic data. Relied on the gradient or
942 > hessian, advanced methods like Newton-Raphson converge rapidly to a
943 > local minimum, but become unstable if the energy surface is far from
944 > quadratic. Another factor that must be taken into account, when
945   choosing energy minimization method, is the size of the system.
946   Steepest descent and conjugate gradient can deal with models of any
947 < size. Because of the limit of computation power to calculate hessian
948 < matrix and insufficient storage capacity to store them, most
949 < Newton-Raphson methods can not be used with very large models.
947 > size. Because of the limits on computer memory to store the hessian
948 > matrix and the computing power needed to diagonalized these
949 > matrices, most Newton-Raphson methods can not be used with very
950 > large systems.
951  
952 < \subsubsection{Heating}
952 > \subsubsection{\textbf{Heating}}
953  
954   Typically, Heating is performed by assigning random velocities
955 < according to a Gaussian distribution for a temperature. Beginning at
956 < a lower temperature and gradually increasing the temperature by
957 < assigning greater random velocities, we end up with setting the
958 < temperature of the system to a final temperature at which the
959 < simulation will be conducted. In heating phase, we should also keep
960 < the system from drifting or rotating as a whole. Equivalently, the
961 < net linear momentum and angular momentum of the system should be
962 < shifted to zero.
955 > according to a Maxwell-Boltzman distribution for a desired
956 > temperature. Beginning at a lower temperature and gradually
957 > increasing the temperature by assigning larger random velocities, we
958 > end up with setting the temperature of the system to a final
959 > temperature at which the simulation will be conducted. In heating
960 > phase, we should also keep the system from drifting or rotating as a
961 > whole. To do this, the net linear momentum and angular momentum of
962 > the system is shifted to zero after each resampling from the Maxwell
963 > -Boltzman distribution.
964  
965 < \subsubsection{Equilibration}
965 > \subsubsection{\textbf{Equilibration}}
966  
967   The purpose of equilibration is to allow the system to evolve
968   spontaneously for a period of time and reach equilibrium. The
# Line 986 | Line 976 | way.
976  
977   \subsection{\label{introSection:production}Production}
978  
979 < \subsubsection{\label{introSec:forceCalculation}The Force Calculation}
979 > The production run is the most important step of the simulation, in
980 > which the equilibrated structure is used as a starting point and the
981 > motions of the molecules are collected for later analysis. In order
982 > to capture the macroscopic properties of the system, the molecular
983 > dynamics simulation must be performed by sampling correctly and
984 > efficiently from the relevant thermodynamic ensemble.
985  
986 < \subsubsection{\label{introSection:integrationSchemes} Integration
987 < Schemes}
986 > The most expensive part of a molecular dynamics simulation is the
987 > calculation of non-bonded forces, such as van der Waals force and
988 > Coulombic forces \textit{etc}. For a system of $N$ particles, the
989 > complexity of the algorithm for pair-wise interactions is $O(N^2 )$,
990 > which making large simulations prohibitive in the absence of any
991 > algorithmic tricks.
992  
993 + A natural approach to avoid system size issues is to represent the
994 + bulk behavior by a finite number of the particles. However, this
995 + approach will suffer from the surface effect at the edges of the
996 + simulation. To offset this, \textit{Periodic boundary conditions}
997 + (see Fig.~\ref{introFig:pbc}) is developed to simulate bulk
998 + properties with a relatively small number of particles. In this
999 + method, the simulation box is replicated throughout space to form an
1000 + infinite lattice. During the simulation, when a particle moves in
1001 + the primary cell, its image in other cells move in exactly the same
1002 + direction with exactly the same orientation. Thus, as a particle
1003 + leaves the primary cell, one of its images will enter through the
1004 + opposite face.
1005 + \begin{figure}
1006 + \centering
1007 + \includegraphics[width=\linewidth]{pbc.eps}
1008 + \caption[An illustration of periodic boundary conditions]{A 2-D
1009 + illustration of periodic boundary conditions. As one particle leaves
1010 + the left of the simulation box, an image of it enters the right.}
1011 + \label{introFig:pbc}
1012 + \end{figure}
1013 +
1014 + %cutoff and minimum image convention
1015 + Another important technique to improve the efficiency of force
1016 + evaluation is to apply spherical cutoff where particles farther than
1017 + a predetermined distance are not included in the calculation
1018 + \cite{Frenkel1996}. The use of a cutoff radius will cause a
1019 + discontinuity in the potential energy curve. Fortunately, one can
1020 + shift simple radial potential to ensure the potential curve go
1021 + smoothly to zero at the cutoff radius. The cutoff strategy works
1022 + well for Lennard-Jones interaction because of its short range
1023 + nature. However, simply truncating the electrostatic interaction
1024 + with the use of cutoffs has been shown to lead to severe artifacts
1025 + in simulations. The Ewald summation, in which the slowly decaying
1026 + Coulomb potential is transformed into direct and reciprocal sums
1027 + with rapid and absolute convergence, has proved to minimize the
1028 + periodicity artifacts in liquid simulations. Taking the advantages
1029 + of the fast Fourier transform (FFT) for calculating discrete Fourier
1030 + transforms, the particle mesh-based
1031 + methods\cite{Hockney1981,Shimada1993, Luty1994} are accelerated from
1032 + $O(N^{3/2})$ to $O(N logN)$. An alternative approach is the
1033 + \emph{fast multipole method}\cite{Greengard1987, Greengard1994},
1034 + which treats Coulombic interactions exactly at short range, and
1035 + approximate the potential at long range through multipolar
1036 + expansion. In spite of their wide acceptance at the molecular
1037 + simulation community, these two methods are difficult to implement
1038 + correctly and efficiently. Instead, we use a damped and
1039 + charge-neutralized Coulomb potential method developed by Wolf and
1040 + his coworkers\cite{Wolf1999}. The shifted Coulomb potential for
1041 + particle $i$ and particle $j$ at distance $r_{rj}$ is given by:
1042 + \begin{equation}
1043 + V(r_{ij})= \frac{q_i q_j \textrm{erfc}(\alpha
1044 + r_{ij})}{r_{ij}}-\lim_{r_{ij}\rightarrow
1045 + R_\textrm{c}}\left\{\frac{q_iq_j \textrm{erfc}(\alpha
1046 + r_{ij})}{r_{ij}}\right\}. \label{introEquation:shiftedCoulomb}
1047 + \end{equation}
1048 + where $\alpha$ is the convergence parameter. Due to the lack of
1049 + inherent periodicity and rapid convergence,this method is extremely
1050 + efficient and easy to implement.
1051 + \begin{figure}
1052 + \centering
1053 + \includegraphics[width=\linewidth]{shifted_coulomb.eps}
1054 + \caption[An illustration of shifted Coulomb potential]{An
1055 + illustration of shifted Coulomb potential.}
1056 + \label{introFigure:shiftedCoulomb}
1057 + \end{figure}
1058 +
1059 + %multiple time step
1060 +
1061   \subsection{\label{introSection:Analysis} Analysis}
1062  
1063 + Recently, advanced visualization technique have become applied to
1064 + monitor the motions of molecules. Although the dynamics of the
1065 + system can be described qualitatively from animation, quantitative
1066 + trajectory analysis are more useful. According to the principles of
1067 + Statistical Mechanics, Sec.~\ref{introSection:statisticalMechanics},
1068 + one can compute thermodynamic properties, analyze fluctuations of
1069 + structural parameters, and investigate time-dependent processes of
1070 + the molecule from the trajectories.
1071 +
1072 + \subsubsection{\label{introSection:thermodynamicsProperties}\textbf{Thermodynamic Properties}}
1073 +
1074 + Thermodynamic properties, which can be expressed in terms of some
1075 + function of the coordinates and momenta of all particles in the
1076 + system, can be directly computed from molecular dynamics. The usual
1077 + way to measure the pressure is based on virial theorem of Clausius
1078 + which states that the virial is equal to $-3Nk_BT$. For a system
1079 + with forces between particles, the total virial, $W$, contains the
1080 + contribution from external pressure and interaction between the
1081 + particles:
1082 + \[
1083 + W =  - 3PV + \left\langle {\sum\limits_{i < j} {r{}_{ij} \cdot
1084 + f_{ij} } } \right\rangle
1085 + \]
1086 + where $f_{ij}$ is the force between particle $i$ and $j$ at a
1087 + distance $r_{ij}$. Thus, the expression for the pressure is given
1088 + by:
1089 + \begin{equation}
1090 + P = \frac{{Nk_B T}}{V} - \frac{1}{{3V}}\left\langle {\sum\limits_{i
1091 + < j} {r{}_{ij} \cdot f_{ij} } } \right\rangle
1092 + \end{equation}
1093 +
1094 + \subsubsection{\label{introSection:structuralProperties}\textbf{Structural Properties}}
1095 +
1096 + Structural Properties of a simple fluid can be described by a set of
1097 + distribution functions. Among these functions,the \emph{pair
1098 + distribution function}, also known as \emph{radial distribution
1099 + function}, is of most fundamental importance to liquid theory.
1100 + Experimentally, pair distribution function can be gathered by
1101 + Fourier transforming raw data from a series of neutron diffraction
1102 + experiments and integrating over the surface factor
1103 + \cite{Powles1973}. The experimental results can serve as a criterion
1104 + to justify the correctness of a liquid model. Moreover, various
1105 + equilibrium thermodynamic and structural properties can also be
1106 + expressed in terms of radial distribution function \cite{Allen1987}.
1107 +
1108 + The pair distribution functions $g(r)$ gives the probability that a
1109 + particle $i$ will be located at a distance $r$ from a another
1110 + particle $j$ in the system
1111 + \[
1112 + g(r) = \frac{V}{{N^2 }}\left\langle {\sum\limits_i {\sum\limits_{j
1113 + \ne i} {\delta (r - r_{ij} )} } } \right\rangle = \frac{\rho
1114 + (r)}{\rho}.
1115 + \]
1116 + Note that the delta function can be replaced by a histogram in
1117 + computer simulation. Peaks in $g(r)$ represent solvent shells, and
1118 + the height of these peaks gradually decreases to 1 as the liquid of
1119 + large distance approaches the bulk density.
1120 +
1121 +
1122 + \subsubsection{\label{introSection:timeDependentProperties}\textbf{Time-dependent
1123 + Properties}}
1124 +
1125 + Time-dependent properties are usually calculated using \emph{time
1126 + correlation functions}, which correlate random variables $A$ and $B$
1127 + at two different times,
1128 + \begin{equation}
1129 + C_{AB} (t) = \left\langle {A(t)B(0)} \right\rangle.
1130 + \label{introEquation:timeCorrelationFunction}
1131 + \end{equation}
1132 + If $A$ and $B$ refer to same variable, this kind of correlation
1133 + function is called an \emph{autocorrelation function}. One example
1134 + of an auto correlation function is the velocity auto-correlation
1135 + function which is directly related to transport properties of
1136 + molecular liquids:
1137 + \[
1138 + D = \frac{1}{3}\int\limits_0^\infty  {\left\langle {v(t) \cdot v(0)}
1139 + \right\rangle } dt
1140 + \]
1141 + where $D$ is diffusion constant. Unlike the velocity autocorrelation
1142 + function, which is averaging over time origins and over all the
1143 + atoms, the dipole autocorrelation functions are calculated for the
1144 + entire system. The dipole autocorrelation function is given by:
1145 + \[
1146 + c_{dipole}  = \left\langle {u_{tot} (t) \cdot u_{tot} (t)}
1147 + \right\rangle
1148 + \]
1149 + Here $u_{tot}$ is the net dipole of the entire system and is given
1150 + by
1151 + \[
1152 + u_{tot} (t) = \sum\limits_i {u_i (t)}
1153 + \]
1154 + In principle, many time correlation functions can be related with
1155 + Fourier transforms of the infrared, Raman, and inelastic neutron
1156 + scattering spectra of molecular liquids. In practice, one can
1157 + extract the IR spectrum from the intensity of dipole fluctuation at
1158 + each frequency using the following relationship:
1159 + \[
1160 + \hat c_{dipole} (v) = \int_{ - \infty }^\infty  {c_{dipole} (t)e^{ -
1161 + i2\pi vt} dt}
1162 + \]
1163 +
1164   \section{\label{introSection:rigidBody}Dynamics of Rigid Bodies}
1165  
1166   Rigid bodies are frequently involved in the modeling of different
1167   areas, from engineering, physics, to chemistry. For example,
1168   missiles and vehicle are usually modeled by rigid bodies.  The
1169   movement of the objects in 3D gaming engine or other physics
1170 < simulator is governed by the rigid body dynamics. In molecular
1171 < simulation, rigid body is used to simplify the model in
1172 < protein-protein docking study{\cite{Gray03}}.
1170 > simulator is governed by rigid body dynamics. In molecular
1171 > simulations, rigid bodies are used to simplify protein-protein
1172 > docking studies\cite{Gray2003}.
1173  
1174   It is very important to develop stable and efficient methods to
1175 < integrate the equations of motion of orientational degrees of
1176 < freedom. Euler angles are the nature choice to describe the
1177 < rotational degrees of freedom. However, due to its singularity, the
1178 < numerical integration of corresponding equations of motion is very
1179 < inefficient and inaccurate. Although an alternative integrator using
1180 < different sets of Euler angles can overcome this difficulty\cite{},
1181 < the computational penalty and the lost of angular momentum
1182 < conservation still remain. A singularity free representation
1183 < utilizing quaternions was developed by Evans in 1977. Unfortunately,
1184 < this approach suffer from the nonseparable Hamiltonian resulted from
1175 > integrate the equations of motion for orientational degrees of
1176 > freedom. Euler angles are the natural choice to describe the
1177 > rotational degrees of freedom. However, due to $\frac {1}{sin
1178 > \theta}$ singularities, the numerical integration of corresponding
1179 > equations of motion is very inefficient and inaccurate. Although an
1180 > alternative integrator using multiple sets of Euler angles can
1181 > overcome this difficulty\cite{Barojas1973}, the computational
1182 > penalty and the loss of angular momentum conservation still remain.
1183 > A singularity-free representation utilizing quaternions was
1184 > developed by Evans in 1977\cite{Evans1977}. Unfortunately, this
1185 > approach uses a nonseparable Hamiltonian resulting from the
1186   quaternion representation, which prevents the symplectic algorithm
1187   to be utilized. Another different approach is to apply holonomic
1188   constraints to the atoms belonging to the rigid body. Each atom
1189   moves independently under the normal forces deriving from potential
1190   energy and constraint forces which are used to guarantee the
1191 < rigidness. However, due to their iterative nature, SHAKE and Rattle
1192 < algorithm converge very slowly when the number of constraint
1193 < increases.
1191 > rigidness. However, due to their iterative nature, the SHAKE and
1192 > Rattle algorithms also converge very slowly when the number of
1193 > constraints increases\cite{Ryckaert1977, Andersen1983}.
1194  
1195 < The break through in geometric literature suggests that, in order to
1195 > A break-through in geometric literature suggests that, in order to
1196   develop a long-term integration scheme, one should preserve the
1197 < symplectic structure of the flow. Introducing conjugate momentum to
1198 < rotation matrix $Q$ and re-formulating Hamiltonian's equation, a
1199 < symplectic integrator, RSHAKE, was proposed to evolve the
1200 < Hamiltonian system in a constraint manifold by iteratively
1201 < satisfying the orthogonality constraint $Q_T Q = 1$. An alternative
1202 < method using quaternion representation was developed by Omelyan.
1203 < However, both of these methods are iterative and inefficient. In
1204 < this section, we will present a symplectic Lie-Poisson integrator
1205 < for rigid body developed by Dullweber and his
1206 < coworkers\cite{Dullweber1997} in depth.
1197 > symplectic structure of the flow. By introducing a conjugate
1198 > momentum to the rotation matrix $Q$ and re-formulating Hamiltonian's
1199 > equation, a symplectic integrator, RSHAKE\cite{Kol1997}, was
1200 > proposed to evolve the Hamiltonian system in a constraint manifold
1201 > by iteratively satisfying the orthogonality constraint $Q^T Q = 1$.
1202 > An alternative method using the quaternion representation was
1203 > developed by Omelyan\cite{Omelyan1998}. However, both of these
1204 > methods are iterative and inefficient. In this section, we descibe a
1205 > symplectic Lie-Poisson integrator for rigid body developed by
1206 > Dullweber and his coworkers\cite{Dullweber1997} in depth.
1207  
1208 < \subsection{\label{introSection:constrainedHamiltonianRB}Constrained Hamiltonian for Rigid Body}
1209 < The motion of the rigid body is Hamiltonian with the Hamiltonian
1208 > \subsection{\label{introSection:constrainedHamiltonianRB}Constrained Hamiltonian for Rigid Bodies}
1209 > The motion of a rigid body is Hamiltonian with the Hamiltonian
1210   function
1211   \begin{equation}
1212   H = \frac{1}{2}(p^T m^{ - 1} p) + \frac{1}{2}tr(PJ^{ - 1} P) +
# Line 1051 | Line 1220 | constrained Hamiltonian equation subjects to a holonom
1220   I_{ii}^{ - 1}  = \frac{1}{2}\sum\limits_{i \ne j} {J_{jj}^{ - 1} }
1221   \]
1222   where $I_{ii}$ is the diagonal element of the inertia tensor. This
1223 < constrained Hamiltonian equation subjects to a holonomic constraint,
1223 > constrained Hamiltonian equation is subjected to a holonomic
1224 > constraint,
1225   \begin{equation}
1226 < Q^T Q = 1$, \label{introEquation:orthogonalConstraint}
1226 > Q^T Q = 1, \label{introEquation:orthogonalConstraint}
1227   \end{equation}
1228 < which is used to ensure rotation matrix's orthogonality.
1229 < Differentiating \ref{introEquation:orthogonalConstraint} and using
1230 < Equation \ref{introEquation:RBMotionMomentum}, one may obtain,
1228 > which is used to ensure rotation matrix's unitarity. Differentiating
1229 > \ref{introEquation:orthogonalConstraint} and using Equation
1230 > \ref{introEquation:RBMotionMomentum}, one may obtain,
1231   \begin{equation}
1232   Q^T PJ^{ - 1}  + J^{ - 1} P^T Q = 0 . \\
1233   \label{introEquation:RBFirstOrderConstraint}
# Line 1066 | Line 1236 | the equations of motion,
1236   Using Equation (\ref{introEquation:motionHamiltonianCoordinate},
1237   \ref{introEquation:motionHamiltonianMomentum}), one can write down
1238   the equations of motion,
1069 \[
1070 \begin{array}{c}
1071 \frac{{dq}}{{dt}} = \frac{p}{m} \label{introEquation:RBMotionPosition}\\
1072 \frac{{dp}}{{dt}} =  - \nabla _q V(q,Q) \label{introEquation:RBMotionMomentum}\\
1073 \frac{{dQ}}{{dt}} = PJ^{ - 1}  \label{introEquation:RBMotionRotation}\\
1074 \frac{{dP}}{{dt}} =  - \nabla _Q V(q,Q) - 2Q\Lambda . \label{introEquation:RBMotionP}\\
1075 \end{array}
1076 \]
1239  
1240 + \begin{eqnarray}
1241 + \frac{{dq}}{{dt}} & = & \frac{p}{m} \label{introEquation:RBMotionPosition}\\
1242 + \frac{{dp}}{{dt}} & = & - \nabla _q V(q,Q) \label{introEquation:RBMotionMomentum}\\
1243 + \frac{{dQ}}{{dt}} & = & PJ^{ - 1}  \label{introEquation:RBMotionRotation}\\
1244 + \frac{{dP}}{{dt}} & = & - \nabla _Q V(q,Q) - 2Q\Lambda . \label{introEquation:RBMotionP}
1245 + \end{eqnarray}
1246 +
1247   In general, there are two ways to satisfy the holonomic constraints.
1248 < We can use constraint force provided by lagrange multiplier on the
1249 < normal manifold to keep the motion on constraint space. Or we can
1250 < simply evolve the system in constraint manifold. The two method are
1251 < proved to be equivalent. The holonomic constraint and equations of
1252 < motions define a constraint manifold for rigid body
1248 > We can use a constraint force provided by a Lagrange multiplier on
1249 > the normal manifold to keep the motion on constraint space. Or we
1250 > can simply evolve the system on the constraint manifold. These two
1251 > methods have been proved to be equivalent. The holonomic constraint
1252 > and equations of motions define a constraint manifold for rigid
1253 > bodies
1254   \[
1255   M = \left\{ {(Q,P):Q^T Q = 1,Q^T PJ^{ - 1}  + J^{ - 1} P^T Q = 0}
1256   \right\}.
# Line 1089 | Line 1259 | diffeomorphic. Introducing
1259   Unfortunately, this constraint manifold is not the cotangent bundle
1260   $T_{\star}SO(3)$. However, it turns out that under symplectic
1261   transformation, the cotangent space and the phase space are
1262 < diffeomorphic. Introducing
1262 > diffeomorphic. By introducing
1263   \[
1264   \tilde Q = Q,\tilde P = \frac{1}{2}\left( {P - QP^T Q} \right),
1265   \]
# Line 1121 | Line 1291 | body, angular momentum on body frame $\Pi  = Q^t P$ is
1291   respectively.
1292  
1293   As a common choice to describe the rotation dynamics of the rigid
1294 < body, angular momentum on body frame $\Pi  = Q^t P$ is introduced to
1295 < rewrite the equations of motion,
1294 > body, the angular momentum on the body fixed frame $\Pi  = Q^t P$ is
1295 > introduced to rewrite the equations of motion,
1296   \begin{equation}
1297   \begin{array}{l}
1298   \mathop \Pi \limits^ \bullet   = J^{ - 1} \Pi ^T \Pi  + Q^T \sum\limits_i {F_i (q,Q)X_i^T }  - \Lambda  \\
# Line 1154 | Line 1324 | operations
1324   \[
1325   \hat vu = v \times u
1326   \]
1157
1327   Using \ref{introEqaution:RBMotionPI}, one can construct a skew
1328   matrix,
1329   \begin{equation}
1330 < (\mathop \Pi \limits^ \bullet   - \mathop \Pi \limits^ \bullet  ^T
1330 > (\mathop \Pi \limits^ \bullet   - \mathop \Pi \limits^ {\bullet  ^T}
1331   ){\rm{ }} = {\rm{ }}(\Pi  - \Pi ^T ){\rm{ }}(J^{ - 1} \Pi  + \Pi J^{
1332   - 1} ) + \sum\limits_i {[Q^T F_i (r,Q)X_i^T  - X_i F_i (r,Q)^T Q]} -
1333   (\Lambda  - \Lambda ^T ) . \label{introEquation:skewMatrixPI}
# Line 1166 | Line 1335 | unique property eliminate the requirement of iteration
1335   Since $\Lambda$ is symmetric, the last term of Equation
1336   \ref{introEquation:skewMatrixPI} is zero, which implies the Lagrange
1337   multiplier $\Lambda$ is absent from the equations of motion. This
1338 < unique property eliminate the requirement of iterations which can
1339 < not be avoided in other methods\cite{}.
1338 > unique property eliminates the requirement of iterations which can
1339 > not be avoided in other methods\cite{Kol1997, Omelyan1998}.
1340  
1341 < Applying hat-map isomorphism, we obtain the equation of motion for
1342 < angular momentum on body frame
1341 > Applying the hat-map isomorphism, we obtain the equation of motion
1342 > for angular momentum on body frame
1343   \begin{equation}
1344   \dot \pi  = \pi  \times I^{ - 1} \pi  + \sum\limits_i {\left( {Q^T
1345   F_i (r,Q)} \right) \times X_i }.
# Line 1185 | Line 1354 | If there is not external forces exerted on the rigid b
1354   \subsection{\label{introSection:SymplecticFreeRB}Symplectic
1355   Lie-Poisson Integrator for Free Rigid Body}
1356  
1357 < If there is not external forces exerted on the rigid body, the only
1358 < contribution to the rotational is from the kinetic potential (the
1359 < first term of \ref{ introEquation:bodyAngularMotion}). The free
1360 < rigid body is an example of Lie-Poisson system with Hamiltonian
1357 > If there are no external forces exerted on the rigid body, the only
1358 > contribution to the rotational motion is from the kinetic energy
1359 > (the first term of \ref{introEquation:bodyAngularMotion}). The free
1360 > rigid body is an example of a Lie-Poisson system with Hamiltonian
1361   function
1362   \begin{equation}
1363   T^r (\pi ) = T_1 ^r (\pi _1 ) + T_2^r (\pi _2 ) + T_3^r (\pi _3 )
# Line 1236 | Line 1405 | tR_1 }$, we can use Cayley transformation,
1405   \end{array}} \right),\theta _1  = \frac{{\pi _1 }}{{I_1 }}\Delta t.
1406   \]
1407   To reduce the cost of computing expensive functions in $e^{\Delta
1408 < tR_1 }$, we can use Cayley transformation,
1408 > tR_1 }$, we can use Cayley transformation to obtain a single-aixs
1409 > propagator,
1410   \[
1411   e^{\Delta tR_1 }  \approx (1 - \Delta tR_1 )^{ - 1} (1 + \Delta tR_1
1412   )
1413   \]
1414   The flow maps for $T_2^r$ and $T_3^r$ can be found in the same
1415 < manner.
1416 <
1247 < In order to construct a second-order symplectic method, we split the
1248 < angular kinetic Hamiltonian function can into five terms
1415 > manner. In order to construct a second-order symplectic method, we
1416 > split the angular kinetic Hamiltonian function can into five terms
1417   \[
1418   T^r (\pi ) = \frac{1}{2}T_1 ^r (\pi _1 ) + \frac{1}{2}T_2^r (\pi _2
1419   ) + T_3^r (\pi _3 ) + \frac{1}{2}T_2^r (\pi _2 ) + \frac{1}{2}T_1 ^r
1420 < (\pi _1 )
1421 < \].
1422 < Concatenating flows corresponding to these five terms, we can obtain
1423 < an symplectic integrator,
1420 > (\pi _1 ).
1421 > \]
1422 > By concatenating the propagators corresponding to these five terms,
1423 > we can obtain an symplectic integrator,
1424   \[
1425   \varphi _{\Delta t,T^r }  = \varphi _{\Delta t/2,\pi _1 }  \circ
1426   \varphi _{\Delta t/2,\pi _2 }  \circ \varphi _{\Delta t,\pi _3 }
# Line 1279 | Line 1447 | Lie-Poisson integrator is found to be extremely effici
1447   \]
1448   Thus $ [\nabla _\pi  F(\pi )]^T J(\pi ) =  - S'(\frac{{\parallel \pi
1449   \parallel ^2 }}{2})\pi  \times \pi  = 0 $. This explicit
1450 < Lie-Poisson integrator is found to be extremely efficient and stable
1451 < which can be explained by the fact the small angle approximation is
1452 < used and the norm of the angular momentum is conserved.
1450 > Lie-Poisson integrator is found to be both extremely efficient and
1451 > stable. These properties can be explained by the fact the small
1452 > angle approximation is used and the norm of the angular momentum is
1453 > conserved.
1454  
1455   \subsection{\label{introSection:RBHamiltonianSplitting} Hamiltonian
1456   Splitting for Rigid Body}
# Line 1293 | Line 1462 | kinetic energy are listed in the below table,
1462   \]
1463   The equations of motion corresponding to potential energy and
1464   kinetic energy are listed in the below table,
1465 + \begin{table}
1466 + \caption{Equations of motion due to Potential and Kinetic Energies}
1467   \begin{center}
1468   \begin{tabular}{|l|l|}
1469    \hline
# Line 1305 | Line 1476 | A second-order symplectic method is now obtained by th
1476    \hline
1477   \end{tabular}
1478   \end{center}
1479 + \end{table}
1480   A second-order symplectic method is now obtained by the composition
1481 < of the flow maps,
1481 > of the position and velocity propagators,
1482   \[
1483   \varphi _{\Delta t}  = \varphi _{\Delta t/2,V}  \circ \varphi
1484   _{\Delta t,T}  \circ \varphi _{\Delta t/2,V}.
1485   \]
1486   Moreover, $\varphi _{\Delta t/2,V}$ can be divided into two
1487 < sub-flows which corresponding to force and torque respectively,
1487 > sub-propagators which corresponding to force and torque
1488 > respectively,
1489   \[
1490   \varphi _{\Delta t/2,V}  = \varphi _{\Delta t/2,F}  \circ \varphi
1491   _{\Delta t/2,\tau }.
1492   \]
1493   Since the associated operators of $\varphi _{\Delta t/2,F} $ and
1494 < $\circ \varphi _{\Delta t/2,\tau }$ are commuted, the composition
1495 < order inside $\varphi _{\Delta t/2,V}$ does not matter.
1496 <
1497 < Furthermore, kinetic potential can be separated to translational
1325 < kinetic term, $T^t (p)$, and rotational kinetic term, $T^r (\pi )$,
1494 > $\circ \varphi _{\Delta t/2,\tau }$ commute, the composition order
1495 > inside $\varphi _{\Delta t/2,V}$ does not matter. Furthermore, the
1496 > kinetic energy can be separated to translational kinetic term, $T^t
1497 > (p)$, and rotational kinetic term, $T^r (\pi )$,
1498   \begin{equation}
1499   T(p,\pi ) =T^t (p) + T^r (\pi ).
1500   \end{equation}
1501   where $ T^t (p) = \frac{1}{2}p^T m^{ - 1} p $ and $T^r (\pi )$ is
1502   defined by \ref{introEquation:rotationalKineticRB}. Therefore, the
1503 < corresponding flow maps are given by
1503 > corresponding propagators are given by
1504   \[
1505   \varphi _{\Delta t,T}  = \varphi _{\Delta t,T^t }  \circ \varphi
1506   _{\Delta t,T^r }.
1507   \]
1508 < Finally, we obtain the overall symplectic flow maps for free moving
1509 < rigid body
1508 > Finally, we obtain the overall symplectic propagators for freely
1509 > moving rigid bodies
1510   \begin{equation}
1511   \begin{array}{c}
1512   \varphi _{\Delta t}  = \varphi _{\Delta t/2,F}  \circ \varphi _{\Delta t/2,\tau }  \\
# Line 1348 | Line 1520 | the theory of Langevin dynamics simulation. A brief de
1520   As an alternative to newtonian dynamics, Langevin dynamics, which
1521   mimics a simple heat bath with stochastic and dissipative forces,
1522   has been applied in a variety of studies. This section will review
1523 < the theory of Langevin dynamics simulation. A brief derivation of
1524 < generalized Langevin equation will be given first. Follow that, we
1525 < will discuss the physical meaning of the terms appearing in the
1526 < equation as well as the calculation of friction tensor from
1527 < hydrodynamics theory.
1523 > the theory of Langevin dynamics. A brief derivation of generalized
1524 > Langevin equation will be given first. Following that, we will
1525 > discuss the physical meaning of the terms appearing in the equation
1526 > as well as the calculation of friction tensor from hydrodynamics
1527 > theory.
1528  
1529   \subsection{\label{introSection:generalizedLangevinDynamics}Derivation of Generalized Langevin Equation}
1530  
1531 < Harmonic bath model, in which an effective set of harmonic
1531 > A harmonic bath model, in which an effective set of harmonic
1532   oscillators are used to mimic the effect of a linearly responding
1533   environment, has been widely used in quantum chemistry and
1534   statistical mechanics. One of the successful applications of
1535 < Harmonic bath model is the derivation of Deriving Generalized
1536 < Langevin Dynamics. Lets consider a system, in which the degree of
1535 > Harmonic bath model is the derivation of the Generalized Langevin
1536 > Dynamics (GLE). Lets consider a system, in which the degree of
1537   freedom $x$ is assumed to couple to the bath linearly, giving a
1538   Hamiltonian of the form
1539   \begin{equation}
1540   H = \frac{{p^2 }}{{2m}} + U(x) + H_B  + \Delta U(x,x_1 , \ldots x_N)
1541   \label{introEquation:bathGLE}.
1542   \end{equation}
1543 < Here $p$ is a momentum conjugate to $q$, $m$ is the mass associated
1544 < with this degree of freedom, $H_B$ is harmonic bath Hamiltonian,
1543 > Here $p$ is a momentum conjugate to $x$, $m$ is the mass associated
1544 > with this degree of freedom, $H_B$ is a harmonic bath Hamiltonian,
1545   \[
1546   H_B  = \sum\limits_{\alpha  = 1}^N {\left\{ {\frac{{p_\alpha ^2
1547   }}{{2m_\alpha  }} + \frac{1}{2}m_\alpha  \omega _\alpha ^2 }
# Line 1377 | Line 1549 | the harmonic bath masses, and $\Delta U$ is bilinear s
1549   \]
1550   where the index $\alpha$ runs over all the bath degrees of freedom,
1551   $\omega _\alpha$ are the harmonic bath frequencies, $m_\alpha$ are
1552 < the harmonic bath masses, and $\Delta U$ is bilinear system-bath
1552 > the harmonic bath masses, and $\Delta U$ is a bilinear system-bath
1553   coupling,
1554   \[
1555   \Delta U =  - \sum\limits_{\alpha  = 1}^N {g_\alpha  x_\alpha  x}
1556   \]
1557 < where $g_\alpha$ are the coupling constants between the bath and the
1558 < coordinate $x$. Introducing
1557 > where $g_\alpha$ are the coupling constants between the bath
1558 > coordinates ($x_ \alpha$) and the system coordinate ($x$).
1559 > Introducing
1560   \[
1561   W(x) = U(x) - \sum\limits_{\alpha  = 1}^N {\frac{{g_\alpha ^2
1562   }}{{2m_\alpha  w_\alpha ^2 }}} x^2
# Line 1398 | Line 1571 | Generalized Langevin Dynamics by Hamilton's equations
1571   \]
1572   Since the first two terms of the new Hamiltonian depend only on the
1573   system coordinates, we can get the equations of motion for
1574 < Generalized Langevin Dynamics by Hamilton's equations
1402 < \ref{introEquation:motionHamiltonianCoordinate,
1403 < introEquation:motionHamiltonianMomentum},
1574 > Generalized Langevin Dynamics by Hamilton's equations,
1575   \begin{equation}
1576   m\ddot x =  - \frac{{\partial W(x)}}{{\partial x}} -
1577   \sum\limits_{\alpha  = 1}^N {g_\alpha  \left( {x_\alpha   -
# Line 1417 | Line 1588 | differential equations, Laplace transform is the appro
1588   In order to derive an equation for $x$, the dynamics of the bath
1589   variables $x_\alpha$ must be solved exactly first. As an integral
1590   transform which is particularly useful in solving linear ordinary
1591 < differential equations, Laplace transform is the appropriate tool to
1592 < solve this problem. The basic idea is to transform the difficult
1591 > differential equations,the Laplace transform is the appropriate tool
1592 > to solve this problem. The basic idea is to transform the difficult
1593   differential equations into simple algebra problems which can be
1594 < solved easily. Then applying inverse Laplace transform, also known
1595 < as the Bromwich integral, we can retrieve the solutions of the
1594 > solved easily. Then, by applying the inverse Laplace transform, also
1595 > known as the Bromwich integral, we can retrieve the solutions of the
1596   original problems.
1597  
1598   Let $f(t)$ be a function defined on $ [0,\infty ) $. The Laplace
# Line 1431 | Line 1602 | Operator. Below are some important properties of Lapla
1602   \]
1603   where  $p$ is real and  $L$ is called the Laplace Transform
1604   Operator. Below are some important properties of Laplace transform
1434 \begin{equation}
1435 \begin{array}{c}
1436 L(x + y) = L(x) + L(y) \\
1437 L(ax) = aL(x) \\
1438 L(\dot x) = pL(x) - px(0) \\
1439 L(\ddot x) = p^2 L(x) - px(0) - \dot x(0) \\
1440 L\left( {\int_0^t {g(t - \tau )h(\tau )d\tau } } \right) = G(p)H(p) \\
1441 \end{array}
1442 \end{equation}
1605  
1606 < Applying Laplace transform to the bath coordinates, we obtain
1607 < \[
1608 < \begin{array}{c}
1609 < p^2 L(x_\alpha  ) - px_\alpha  (0) - \dot x_\alpha  (0) =  - \omega _\alpha ^2 L(x_\alpha  ) + \frac{{g_\alpha  }}{{\omega _\alpha  }}L(x) \\
1610 < L(x_\alpha  ) = \frac{{\frac{{g_\alpha  }}{{\omega _\alpha  }}L(x) + px_\alpha  (0) + \dot x_\alpha  (0)}}{{p^2  + \omega _\alpha ^2 }} \\
1611 < \end{array}
1612 < \]
1606 > \begin{eqnarray*}
1607 > L(x + y)  & = & L(x) + L(y) \\
1608 > L(ax)     & = & aL(x) \\
1609 > L(\dot x) & = & pL(x) - px(0) \\
1610 > L(\ddot x)& = & p^2 L(x) - px(0) - \dot x(0) \\
1611 > L\left( {\int_0^t {g(t - \tau )h(\tau )d\tau } } \right)& = & G(p)H(p) \\
1612 > \end{eqnarray*}
1613 >
1614 >
1615 > Applying the Laplace transform to the bath coordinates, we obtain
1616 > \begin{eqnarray*}
1617 > 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) \\
1618 > L(x_\alpha  ) & = & \frac{{\frac{{g_\alpha  }}{{\omega _\alpha  }}L(x) + px_\alpha  (0) + \dot x_\alpha  (0)}}{{p^2  + \omega _\alpha ^2 }} \\
1619 > \end{eqnarray*}
1620 >
1621   By the same way, the system coordinates become
1622 < \[
1623 < \begin{array}{c}
1624 < mL(\ddot x) =  - \frac{1}{p}\frac{{\partial W(x)}}{{\partial x}} \\
1625 <  - \sum\limits_{\alpha  = 1}^N {\left\{ { - \frac{{g_\alpha ^2 }}{{m_\alpha  \omega _\alpha ^2 }}\frac{p}{{p^2  + \omega _\alpha ^2 }}pL(x) - \frac{p}{{p^2  + \omega _\alpha ^2 }}g_\alpha  x_\alpha  (0) - \frac{1}{{p^2  + \omega _\alpha ^2 }}g_\alpha  \dot x_\alpha  (0)} \right\}}  \\
1456 < \end{array}
1457 < \]
1622 > \begin{eqnarray*}
1623 > mL(\ddot x) & = & - \frac{1}{p}\frac{{\partial W(x)}}{{\partial x}} \\
1624 >  & & \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\}}  \\
1625 > \end{eqnarray*}
1626  
1627   With the help of some relatively important inverse Laplace
1628   transformations:
# Line 1466 | Line 1634 | transformations:
1634   \end{array}
1635   \]
1636   , we obtain
1637 < \begin{align}
1638 < m\ddot x &=  - \frac{{\partial W(x)}}{{\partial x}} -
1637 > \begin{eqnarray*}
1638 > m\ddot x & =  & - \frac{{\partial W(x)}}{{\partial x}} -
1639   \sum\limits_{\alpha  = 1}^N {\left\{ {\left( { - \frac{{g_\alpha ^2
1640   }}{{m_\alpha  \omega _\alpha ^2 }}} \right)\int_0^t {\cos (\omega
1641 < _\alpha  t)\dot x(t - \tau )d\tau  - \left[ {g_\alpha  x_\alpha  (0)
1642 < - \frac{{g_\alpha  }}{{m_\alpha  \omega _\alpha  }}} \right]\cos
1643 < (\omega _\alpha  t) - \frac{{g_\alpha  \dot x_\alpha  (0)}}{{\omega
1644 < _\alpha  }}\sin (\omega _\alpha  t)} } \right\}}
1645 < %
1646 < &= - \frac{{\partial W(x)}}{{\partial x}} - \int_0^t
1641 > _\alpha  t)\dot x(t - \tau )d\tau } } \right\}}  \\
1642 > & & + \sum\limits_{\alpha  = 1}^N {\left\{ {\left[ {g_\alpha
1643 > x_\alpha (0) - \frac{{g_\alpha  }}{{m_\alpha  \omega _\alpha  }}}
1644 > \right]\cos (\omega _\alpha  t) + \frac{{g_\alpha  \dot x_\alpha
1645 > (0)}}{{\omega _\alpha  }}\sin (\omega _\alpha  t)} \right\}}
1646 > \end{eqnarray*}
1647 > \begin{eqnarray*}
1648 > m\ddot x & = & - \frac{{\partial W(x)}}{{\partial x}} - \int_0^t
1649   {\sum\limits_{\alpha  = 1}^N {\left( { - \frac{{g_\alpha ^2
1650   }}{{m_\alpha  \omega _\alpha ^2 }}} \right)\cos (\omega _\alpha
1651 < t)\dot x(t - \tau )d} \tau }  + \sum\limits_{\alpha  = 1}^N {\left\{
1652 < {\left[ {g_\alpha  x_\alpha  (0) - \frac{{g_\alpha  }}{{m_\alpha
1653 < \omega _\alpha  }}} \right]\cos (\omega _\alpha  t) +
1654 < \frac{{g_\alpha  \dot x_\alpha  (0)}}{{\omega _\alpha  }}\sin
1655 < (\omega _\alpha  t)} \right\}}
1656 < \end{align}
1487 <
1651 > t)\dot x(t - \tau )d} \tau }  \\
1652 > & & + \sum\limits_{\alpha  = 1}^N {\left\{ {\left[ {g_\alpha
1653 > x_\alpha (0) - \frac{{g_\alpha }}{{m_\alpha \omega _\alpha  }}}
1654 > \right]\cos (\omega _\alpha  t) + \frac{{g_\alpha  \dot x_\alpha
1655 > (0)}}{{\omega _\alpha  }}\sin (\omega _\alpha  t)} \right\}}
1656 > \end{eqnarray*}
1657   Introducing a \emph{dynamic friction kernel}
1658   \begin{equation}
1659   \xi (t) = \sum\limits_{\alpha  = 1}^N {\left( { - \frac{{g_\alpha ^2
# Line 1507 | Line 1676 | which is known as the \emph{generalized Langevin equat
1676   \end{equation}
1677   which is known as the \emph{generalized Langevin equation}.
1678  
1679 < \subsubsection{\label{introSection:randomForceDynamicFrictionKernel}Random Force and Dynamic Friction Kernel}
1679 > \subsubsection{\label{introSection:randomForceDynamicFrictionKernel}\textbf{Random Force and Dynamic Friction Kernel}}
1680  
1681   One may notice that $R(t)$ depends only on initial conditions, which
1682   implies it is completely deterministic within the context of a
# Line 1520 | Line 1689 | as the model, which is gaussian distribution in genera
1689   \end{array}
1690   \]
1691   This property is what we expect from a truly random process. As long
1692 < as the model, which is gaussian distribution in general, chosen for
1693 < $R(t)$ is a truly random process, the stochastic nature of the GLE
1525 < still remains.
1692 > as the model chosen for $R(t)$ was a gaussian distribution in
1693 > general, the stochastic nature of the GLE still remains.
1694  
1695   %dynamic friction kernel
1696   The convolution integral
# Line 1543 | Line 1711 | which can be used to describe dynamic caging effect. T
1711   m\ddot x =  - \frac{\partial }{{\partial x}}\left( {W(x) +
1712   \frac{1}{2}\xi _0 (x - x_0 )^2 } \right) + R(t),
1713   \]
1714 < which can be used to describe dynamic caging effect. The other
1715 < extreme is the bath that responds infinitely quickly to motions in
1716 < the system. Thus, $\xi (t)$ can be taken as a $delta$ function in
1717 < time:
1714 > which can be used to describe the effect of dynamic caging in
1715 > viscous solvents. The other extreme is the bath that responds
1716 > infinitely quickly to motions in the system. Thus, $\xi (t)$ can be
1717 > taken as a $delta$ function in time:
1718   \[
1719   \xi (t) = 2\xi _0 \delta (t)
1720   \]
# Line 1562 | Line 1730 | or be determined by Stokes' law for regular shaped par
1730   \end{equation}
1731   which is known as the Langevin equation. The static friction
1732   coefficient $\xi _0$ can either be calculated from spectral density
1733 < or be determined by Stokes' law for regular shaped particles.A
1733 > or be determined by Stokes' law for regular shaped particles. A
1734   briefly review on calculating friction tensor for arbitrary shaped
1735   particles is given in Sec.~\ref{introSection:frictionTensor}.
1736  
1737 < \subsubsection{\label{introSection:secondFluctuationDissipation}The Second Fluctuation Dissipation Theorem}
1737 > \subsubsection{\label{introSection:secondFluctuationDissipation}\textbf{The Second Fluctuation Dissipation Theorem}}
1738  
1739   Defining a new set of coordinates,
1740   \[
# Line 1578 | Line 1746 | And since the $q$ coordinates are harmonic oscillators
1746   R(t) = \sum\limits_{\alpha  = 1}^N {g_\alpha  q_\alpha  (t)}.
1747   \]
1748   And since the $q$ coordinates are harmonic oscillators,
1749 < \[
1750 < \begin{array}{c}
1751 < \left\langle {q_\alpha ^2 } \right\rangle  = \frac{{kT}}{{m_\alpha  \omega _\alpha ^2 }} \\
1752 < \left\langle {q_\alpha  (t)q_\alpha  (0)} \right\rangle  = \left\langle {q_\alpha ^2 (0)} \right\rangle \cos (\omega _\alpha  t) \\
1753 < \left\langle {q_\alpha  (t)q_\beta  (0)} \right\rangle  = \delta _{\alpha \beta } \left\langle {q_\alpha  (t)q_\alpha  (0)} \right\rangle  \\
1754 < \left\langle {R(t)R(0)} \right\rangle  = \sum\limits_\alpha  {\sum\limits_\beta  {g_\alpha  g_\beta  \left\langle {q_\alpha  (t)q_\beta  (0)} \right\rangle } }  \\
1755 <  = \sum\limits_\alpha  {g_\alpha ^2 \left\langle {q_\alpha ^2 (0)} \right\rangle \cos (\omega _\alpha  t)}  \\
1756 <  = kT\xi (t) \\
1757 < \end{array}
1758 < \]
1749 >
1750 > \begin{eqnarray*}
1751 > \left\langle {q_\alpha ^2 } \right\rangle  & = & \frac{{kT}}{{m_\alpha  \omega _\alpha ^2 }} \\
1752 > \left\langle {q_\alpha  (t)q_\alpha  (0)} \right\rangle & = & \left\langle {q_\alpha ^2 (0)} \right\rangle \cos (\omega _\alpha  t) \\
1753 > \left\langle {q_\alpha  (t)q_\beta  (0)} \right\rangle & = &\delta _{\alpha \beta } \left\langle {q_\alpha  (t)q_\alpha  (0)} \right\rangle  \\
1754 > \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 } }  \\
1755 >  & = &\sum\limits_\alpha  {g_\alpha ^2 \left\langle {q_\alpha ^2 (0)} \right\rangle \cos (\omega _\alpha  t)}  \\
1756 >  & = &kT\xi (t) \\
1757 > \end{eqnarray*}
1758 >
1759   Thus, we recover the \emph{second fluctuation dissipation theorem}
1760   \begin{equation}
1761   \xi (t) = \left\langle {R(t)R(0)} \right\rangle
# Line 1595 | Line 1763 | can model the random force and friction kernel.
1763   \end{equation}
1764   In effect, it acts as a constraint on the possible ways in which one
1765   can model the random force and friction kernel.
1598
1599 \subsection{\label{introSection:frictionTensor} Friction Tensor}
1600 Theoretically, the friction kernel can be determined using velocity
1601 autocorrelation function. However, this approach become impractical
1602 when the system become more and more complicate. Instead, various
1603 approaches based on hydrodynamics have been developed to calculate
1604 the friction coefficients. The friction effect is isotropic in
1605 Equation, \zeta can be taken as a scalar. In general, friction
1606 tensor \Xi is a $6\times 6$ matrix given by
1607 \[
1608 \Xi  = \left( {\begin{array}{*{20}c}
1609   {\Xi _{}^{tt} } & {\Xi _{}^{rt} }  \\
1610   {\Xi _{}^{tr} } & {\Xi _{}^{rr} }  \\
1611 \end{array}} \right).
1612 \]
1613 Here, $ {\Xi^{tt} }$ and $ {\Xi^{rr} }$ are translational friction
1614 tensor and rotational resistance (friction) tensor respectively,
1615 while ${\Xi^{tr} }$ is translation-rotation coupling tensor and $
1616 {\Xi^{rt} }$ is rotation-translation coupling tensor. When a
1617 particle moves in a fluid, it may experience friction force or
1618 torque along the opposite direction of the velocity or angular
1619 velocity,
1620 \[
1621 \left( \begin{array}{l}
1622 F_R  \\
1623 \tau _R  \\
1624 \end{array} \right) =  - \left( {\begin{array}{*{20}c}
1625   {\Xi ^{tt} } & {\Xi ^{rt} }  \\
1626   {\Xi ^{tr} } & {\Xi ^{rr} }  \\
1627 \end{array}} \right)\left( \begin{array}{l}
1628 v \\
1629 w \\
1630 \end{array} \right)
1631 \]
1632 where $F_r$ is the friction force and $\tau _R$ is the friction
1633 toque.
1634
1635 \subsubsection{\label{introSection:resistanceTensorRegular}The Resistance Tensor for Regular Shape}
1636
1637 For a spherical particle, the translational and rotational friction
1638 constant can be calculated from Stoke's law,
1639 \[
1640 \Xi ^{tt}  = \left( {\begin{array}{*{20}c}
1641   {6\pi \eta R} & 0 & 0  \\
1642   0 & {6\pi \eta R} & 0  \\
1643   0 & 0 & {6\pi \eta R}  \\
1644 \end{array}} \right)
1645 \]
1646 and
1647 \[
1648 \Xi ^{rr}  = \left( {\begin{array}{*{20}c}
1649   {8\pi \eta R^3 } & 0 & 0  \\
1650   0 & {8\pi \eta R^3 } & 0  \\
1651   0 & 0 & {8\pi \eta R^3 }  \\
1652 \end{array}} \right)
1653 \]
1654 where $\eta$ is the viscosity of the solvent and $R$ is the
1655 hydrodynamics radius.
1656
1657 Other non-spherical shape, such as cylinder and ellipsoid
1658 \textit{etc}, are widely used as reference for developing new
1659 hydrodynamics theory, because their properties can be calculated
1660 exactly. In 1936, Perrin extended Stokes's law to general ellipsoid,
1661 also called a triaxial ellipsoid, which is given in Cartesian
1662 coordinates by
1663 \[
1664 \frac{{x^2 }}{{a^2 }} + \frac{{y^2 }}{{b^2 }} + \frac{{z^2 }}{{c^2
1665 }} = 1
1666 \]
1667 where the semi-axes are of lengths $a$, $b$, and $c$. Unfortunately,
1668 due to the complexity of the elliptic integral, only the ellipsoid
1669 with the restriction of two axes having to be equal, \textit{i.e.}
1670 prolate($ a \ge b = c$) and oblate ($ a < b = c $), can be solved
1671 exactly. Introducing an elliptic integral parameter $S$ for prolate,
1672 \[
1673 S = \frac{2}{{\sqrt {a^2  - b^2 } }}\ln \frac{{a + \sqrt {a^2  - b^2
1674 } }}{b},
1675 \]
1676 and oblate,
1677 \[
1678 S = \frac{2}{{\sqrt {b^2  - a^2 } }}arctg\frac{{\sqrt {b^2  - a^2 }
1679 }}{a}
1680 \],
1681 one can write down the translational and rotational resistance
1682 tensors
1683 \[
1684 \begin{array}{l}
1685 \Xi _a^{tt}  = 16\pi \eta \frac{{a^2  - b^2 }}{{(2a^2  - b^2 )S - 2a}} \\
1686 \Xi _b^{tt}  = \Xi _c^{tt}  = 32\pi \eta \frac{{a^2  - b^2 }}{{(2a^2  - 3b^2 )S + 2a}} \\
1687 \end{array},
1688 \]
1689 and
1690 \[
1691 \begin{array}{l}
1692 \Xi _a^{rr}  = \frac{{32\pi }}{3}\eta \frac{{(a^2  - b^2 )b^2 }}{{2a - b^2 S}} \\
1693 \Xi _b^{rr}  = \Xi _c^{rr}  = \frac{{32\pi }}{3}\eta \frac{{(a^4  - b^4 )}}{{(2a^2  - b^2 )S - 2a}} \\
1694 \end{array}.
1695 \]
1696
1697 \subsubsection{\label{introSection:resistanceTensorRegularArbitrary}The Resistance Tensor for Arbitrary Shape}
1698
1699 Unlike spherical and other regular shaped molecules, there is not
1700 analytical solution for friction tensor of any arbitrary shaped
1701 rigid molecules. The ellipsoid of revolution model and general
1702 triaxial ellipsoid model have been used to approximate the
1703 hydrodynamic properties of rigid bodies. However, since the mapping
1704 from all possible ellipsoidal space, $r$-space, to all possible
1705 combination of rotational diffusion coefficients, $D$-space is not
1706 unique\cite{Wegener79} as well as the intrinsic coupling between
1707 translational and rotational motion of rigid body\cite{}, general
1708 ellipsoid is not always suitable for modeling arbitrarily shaped
1709 rigid molecule. A number of studies have been devoted to determine
1710 the friction tensor for irregularly shaped rigid bodies using more
1711 advanced method\cite{} where the molecule of interest was modeled by
1712 combinations of spheres(beads)\cite{} and the hydrodynamics
1713 properties of the molecule can be calculated using the hydrodynamic
1714 interaction tensor. Let us consider a rigid assembly of $N$ beads
1715 immersed in a continuous medium. Due to hydrodynamics interaction,
1716 the ``net'' velocity of $i$th bead, $v'_i$ is different than its
1717 unperturbed velocity $v_i$,
1718 \[
1719 v'_i  = v_i  - \sum\limits_{j \ne i} {T_{ij} F_j }
1720 \]
1721 where $F_i$ is the frictional force, and $T_{ij}$ is the
1722 hydrodynamic interaction tensor. The friction force of $i$th bead is
1723 proportional to its ``net'' velocity
1724 \begin{equation}
1725 F_i  = \zeta _i v_i  - \zeta _i \sum\limits_{j \ne i} {T_{ij} F_j }.
1726 \label{introEquation:tensorExpression}
1727 \end{equation}
1728 This equation is the basis for deriving the hydrodynamic tensor. In
1729 1930, Oseen and Burgers gave a simple solution to Equation
1730 \ref{introEquation:tensorExpression}
1731 \begin{equation}
1732 T_{ij}  = \frac{1}{{8\pi \eta r_{ij} }}\left( {I + \frac{{R_{ij}
1733 R_{ij}^T }}{{R_{ij}^2 }}} \right).
1734 \label{introEquation:oseenTensor}
1735 \end{equation}
1736 Here $R_{ij}$ is the distance vector between bead $i$ and bead $j$.
1737 A second order expression for element of different size was
1738 introduced by Rotne and Prager\cite{} and improved by Garc\'{i}a de
1739 la Torre and Bloomfield,
1740 \begin{equation}
1741 T_{ij}  = \frac{1}{{8\pi \eta R_{ij} }}\left[ {\left( {I +
1742 \frac{{R_{ij} R_{ij}^T }}{{R_{ij}^2 }}} \right) + R\frac{{\sigma
1743 _i^2  + \sigma _j^2 }}{{r_{ij}^2 }}\left( {\frac{I}{3} -
1744 \frac{{R_{ij} R_{ij}^T }}{{R_{ij}^2 }}} \right)} \right].
1745 \label{introEquation:RPTensorNonOverlapped}
1746 \end{equation}
1747 Both of the Equation \ref{introEquation:oseenTensor} and Equation
1748 \ref{introEquation:RPTensorNonOverlapped} have an assumption $R_{ij}
1749 \ge \sigma _i  + \sigma _j$. An alternative expression for
1750 overlapping beads with the same radius, $\sigma$, is given by
1751 \begin{equation}
1752 T_{ij}  = \frac{1}{{6\pi \eta R_{ij} }}\left[ {\left( {1 -
1753 \frac{2}{{32}}\frac{{R_{ij} }}{\sigma }} \right)I +
1754 \frac{2}{{32}}\frac{{R_{ij} R_{ij}^T }}{{R_{ij} \sigma }}} \right]
1755 \label{introEquation:RPTensorOverlapped}
1756 \end{equation}
1757
1758 To calculate the resistance tensor at an arbitrary origin $O$, we
1759 construct a $3N \times 3N$ matrix consisting of $N \times N$
1760 $B_{ij}$ blocks
1761 \begin{equation}
1762 B = \left( {\begin{array}{*{20}c}
1763   {B_{11} } &  \ldots  & {B_{1N} }  \\
1764    \vdots  &  \ddots  &  \vdots   \\
1765   {B_{N1} } &  \cdots  & {B_{NN} }  \\
1766 \end{array}} \right),
1767 \end{equation}
1768 where $B_{ij}$ is given by
1769 \[
1770 B_{ij}  = \delta _{ij} \frac{I}{{6\pi \eta R}} + (1 - \delta _{ij}
1771 )T_{ij}
1772 \]
1773 where $\delta _{ij}$ is Kronecker delta function. Inverting matrix
1774 $B$, we obtain
1775
1776 \[
1777 C = B^{ - 1}  = \left( {\begin{array}{*{20}c}
1778   {C_{11} } &  \ldots  & {C_{1N} }  \\
1779    \vdots  &  \ddots  &  \vdots   \\
1780   {C_{N1} } &  \cdots  & {C_{NN} }  \\
1781 \end{array}} \right)
1782 \]
1783 , which can be partitioned into $N \times N$ $3 \times 3$ block
1784 $C_{ij}$. With the help of $C_{ij}$ and skew matrix $U_i$
1785 \[
1786 U_i  = \left( {\begin{array}{*{20}c}
1787   0 & { - z_i } & {y_i }  \\
1788   {z_i } & 0 & { - x_i }  \\
1789   { - y_i } & {x_i } & 0  \\
1790 \end{array}} \right)
1791 \]
1792 where $x_i$, $y_i$, $z_i$ are the components of the vector joining
1793 bead $i$ and origin $O$. Hence, the elements of resistance tensor at
1794 arbitrary origin $O$ can be written as
1795 \begin{equation}
1796 \begin{array}{l}
1797 \Xi _{}^{tt}  = \sum\limits_i {\sum\limits_j {C_{ij} } } , \\
1798 \Xi _{}^{tr}  = \Xi _{}^{rt}  = \sum\limits_i {\sum\limits_j {U_i C_{ij} } } , \\
1799 \Xi _{}^{rr}  =  - \sum\limits_i {\sum\limits_j {U_i C_{ij} } } U_j  \\
1800 \end{array}
1801 \label{introEquation:ResistanceTensorArbitraryOrigin}
1802 \end{equation}
1803
1804 The resistance tensor depends on the origin to which they refer. The
1805 proper location for applying friction force is the center of
1806 resistance (reaction), at which the trace of rotational resistance
1807 tensor, $ \Xi ^{rr}$ reaches minimum. Mathematically, the center of
1808 resistance is defined as an unique point of the rigid body at which
1809 the translation-rotation coupling tensor are symmetric,
1810 \begin{equation}
1811 \Xi^{tr}  = \left( {\Xi^{tr} } \right)^T
1812 \label{introEquation:definitionCR}
1813 \end{equation}
1814 Form Equation \ref{introEquation:ResistanceTensorArbitraryOrigin},
1815 we can easily find out that the translational resistance tensor is
1816 origin independent, while the rotational resistance tensor and
1817 translation-rotation coupling resistance tensor depend on the
1818 origin. Given resistance tensor at an arbitrary origin $O$, and a
1819 vector ,$r_{OP}(x_{OP}, y_{OP}, z_{OP})$, from $O$ to $P$, we can
1820 obtain the resistance tensor at $P$ by
1821 \begin{equation}
1822 \begin{array}{l}
1823 \Xi _P^{tt}  = \Xi _O^{tt}  \\
1824 \Xi _P^{tr}  = \Xi _P^{rt}  = \Xi _O^{tr}  - U_{OP} \Xi _O^{tt}  \\
1825 \Xi _P^{rr}  = \Xi _O^{rr}  - U_{OP} \Xi _O^{tt} U_{OP}  + \Xi _O^{tr} U_{OP}  - U_{OP} \Xi _O^{tr} ^{^T }  \\
1826 \end{array}
1827 \label{introEquation:resistanceTensorTransformation}
1828 \end{equation}
1829 where
1830 \[
1831 U_{OP}  = \left( {\begin{array}{*{20}c}
1832   0 & { - z_{OP} } & {y_{OP} }  \\
1833   {z_i } & 0 & { - x_{OP} }  \\
1834   { - y_{OP} } & {x_{OP} } & 0  \\
1835 \end{array}} \right)
1836 \]
1837 Using Equations \ref{introEquation:definitionCR} and
1838 \ref{introEquation:resistanceTensorTransformation}, one can locate
1839 the position of center of resistance,
1840 \[
1841 \left( \begin{array}{l}
1842 x_{OR}  \\
1843 y_{OR}  \\
1844 z_{OR}  \\
1845 \end{array} \right) = \left( {\begin{array}{*{20}c}
1846   {(\Xi _O^{rr} )_{yy}  + (\Xi _O^{rr} )_{zz} } & { - (\Xi _O^{rr} )_{xy} } & { - (\Xi _O^{rr} )_{xz} }  \\
1847   { - (\Xi _O^{rr} )_{xy} } & {(\Xi _O^{rr} )_{zz}  + (\Xi _O^{rr} )_{xx} } & { - (\Xi _O^{rr} )_{yz} }  \\
1848   { - (\Xi _O^{rr} )_{xz} } & { - (\Xi _O^{rr} )_{yz} } & {(\Xi _O^{rr} )_{xx}  + (\Xi _O^{rr} )_{yy} }  \\
1849 \end{array}} \right)^{ - 1} \left( \begin{array}{l}
1850 (\Xi _O^{tr} )_{yz}  - (\Xi _O^{tr} )_{zy}  \\
1851 (\Xi _O^{tr} )_{zx}  - (\Xi _O^{tr} )_{xz}  \\
1852 (\Xi _O^{tr} )_{xy}  - (\Xi _O^{tr} )_{yx}  \\
1853 \end{array} \right).
1854 \]
1855 where $x_OR$, $y_OR$, $z_OR$ are the components of the vector
1856 joining center of resistance $R$ and origin $O$.
1857
1858 %\section{\label{introSection:correlationFunctions}Correlation Functions}

Diff Legend

Removed lines
+ Added lines
< Changed lines
> Changed lines