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 2697 by tim, Fri Apr 7 05:03:54 2006 UTC vs.
Revision 2779 by tim, Fri May 26 18:25:41 2006 UTC

# Line 27 | Line 27 | $F_ij$ be the force that particle $i$ exerts on partic
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 $i$.
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
33   \begin{equation}
34 < F_ij = -F_ji
34 > F_{ij} = -F_{ji}
35   \label{introEquation:newtonThirdLaw}
36   \end{equation}
37  
# Line 117 | Line 117 | for a holonomic system of $f$ degrees of freedom, the
117   \subsubsection{\label{introSection:equationOfMotionLagrangian}The
118   Equations of Motion in Lagrangian Mechanics}
119  
120 < for a holonomic system of $f$ degrees of freedom, the equations of
120 > For a holonomic system of $f$ degrees of freedom, the equations of
121   motion in the Lagrangian form is
122   \begin{equation}
123   \frac{d}{{dt}}\frac{{\partial L}}{{\partial \dot q_i }} -
# Line 212 | Line 212 | q_i }}} \right) = 0}
212   }}\dot p_i } \right)}  = \sum\limits_i {\left( {\frac{{\partial
213   H}}{{\partial q_i }}\frac{{\partial H}}{{\partial p_i }} -
214   \frac{{\partial H}}{{\partial p_i }}\frac{{\partial H}}{{\partial
215 < q_i }}} \right) = 0}
216 < \label{introEquation:conserveHalmitonian}
215 > q_i }}} \right) = 0} \label{introEquation:conserveHalmitonian}
216   \end{equation}
217  
219 When studying Hamiltonian system, it is more convenient to use
220 notation
221 \begin{equation}
222 r = r(q,p)^T
223 \end{equation}
224 and to introduce a $2n \times 2n$ canonical structure matrix $J$,
225 \begin{equation}
226 J = \left( {\begin{array}{*{20}c}
227   0 & I  \\
228   { - I} & 0  \\
229 \end{array}} \right)
230 \label{introEquation:canonicalMatrix}
231 \end{equation}
232 where $I$ is a $n \times n$ identity matrix and $J$ is a
233 skew-symmetric matrix ($ J^T  =  - J $). Thus, Hamiltonian system
234 can be rewritten as,
235 \begin{equation}
236 \frac{d}{{dt}}r = J\nabla _r H(r)
237 \label{introEquation:compactHamiltonian}
238 \end{equation}
239
218   \section{\label{introSection:statisticalMechanics}Statistical
219   Mechanics}
220  
221   The thermodynamic behaviors and properties of Molecular Dynamics
222   simulation are governed by the principle of Statistical Mechanics.
223   The following section will give a brief introduction to some of the
224 < Statistical Mechanics concepts presented in this dissertation.
224 > Statistical Mechanics concepts and theorem presented in this
225 > dissertation.
226  
227 < \subsection{\label{introSection:ensemble}Ensemble and Phase Space}
227 > \subsection{\label{introSection:ensemble}Phase Space and Ensemble}
228 >
229 > Mathematically, phase space is the space which represents all
230 > possible states. Each possible state of the system corresponds to
231 > one unique point in the phase space. For mechanical systems, the
232 > phase space usually consists of all possible values of position and
233 > momentum variables. Consider a dynamic system in a cartesian space,
234 > where each of the $6f$ coordinates and momenta is assigned to one of
235 > $6f$ mutually orthogonal axes, the phase space of this system is a
236 > $6f$ dimensional space. A point, $x = (q_1 , \ldots ,q_f ,p_1 ,
237 > \ldots ,p_f )$, with a unique set of values of $6f$ coordinates and
238 > momenta is a phase space vector.
239 >
240 > A microscopic state or microstate of a classical system is
241 > specification of the complete phase space vector of a system at any
242 > instant in time. An ensemble is defined as a collection of systems
243 > sharing one or more macroscopic characteristics but each being in a
244 > unique microstate. The complete ensemble is specified by giving all
245 > systems or microstates consistent with the common macroscopic
246 > characteristics of the ensemble. Although the state of each
247 > individual system in the ensemble could be precisely described at
248 > any instance in time by a suitable phase space vector, when using
249 > ensembles for statistical purposes, there is no need to maintain
250 > distinctions between individual systems, since the numbers of
251 > systems at any time in the different states which correspond to
252 > different regions of the phase space are more interesting. Moreover,
253 > in the point of view of statistical mechanics, one would prefer to
254 > use ensembles containing a large enough population of separate
255 > members so that the numbers of systems in such different states can
256 > be regarded as changing continuously as we traverse different
257 > regions of the phase space. The condition of an ensemble at any time
258 > can be regarded as appropriately specified by the density $\rho$
259 > with which representative points are distributed over the phase
260 > space. The density of distribution for an ensemble with $f$ degrees
261 > of freedom is defined as,
262 > \begin{equation}
263 > \rho  = \rho (q_1 , \ldots ,q_f ,p_1 , \ldots ,p_f ,t).
264 > \label{introEquation:densityDistribution}
265 > \end{equation}
266 > Governed by the principles of mechanics, the phase points change
267 > their value which would change the density at any time at phase
268 > space. Hence, the density of distribution is also to be taken as a
269 > function of the time.
270 >
271 > The number of systems $\delta N$ at time $t$ can be determined by,
272 > \begin{equation}
273 > \delta N = \rho (q,p,t)dq_1  \ldots dq_f dp_1  \ldots dp_f.
274 > \label{introEquation:deltaN}
275 > \end{equation}
276 > Assuming a large enough population of systems are exploited, we can
277 > sufficiently approximate $\delta N$ without introducing
278 > discontinuity when we go from one region in the phase space to
279 > another. By integrating over the whole phase space,
280 > \begin{equation}
281 > N = \int { \ldots \int {\rho (q,p,t)dq_1 } ...dq_f dp_1 } ...dp_f
282 > \label{introEquation:totalNumberSystem}
283 > \end{equation}
284 > gives us an expression for the total number of the systems. Hence,
285 > the probability per unit in the phase space can be obtained by,
286 > \begin{equation}
287 > \frac{{\rho (q,p,t)}}{N} = \frac{{\rho (q,p,t)}}{{\int { \ldots \int
288 > {\rho (q,p,t)dq_1 } ...dq_f dp_1 } ...dp_f }}.
289 > \label{introEquation:unitProbability}
290 > \end{equation}
291 > With the help of Equation(\ref{introEquation:unitProbability}) and
292 > the knowledge of the system, it is possible to calculate the average
293 > value of any desired quantity which depends on the coordinates and
294 > momenta of the system. Even when the dynamics of the real system is
295 > complex, or stochastic, or even discontinuous, the average
296 > properties of the ensemble of possibilities as a whole may still
297 > remain well defined. For a classical system in thermal equilibrium
298 > with its environment, the ensemble average of a mechanical quantity,
299 > $\langle A(q , p) \rangle_t$, takes the form of an integral over the
300 > phase space of the system,
301 > \begin{equation}
302 > \langle  A(q , p) \rangle_t = \frac{{\int { \ldots \int {A(q,p)\rho
303 > (q,p,t)dq_1 } ...dq_f dp_1 } ...dp_f }}{{\int { \ldots \int {\rho
304 > (q,p,t)dq_1 } ...dq_f dp_1 } ...dp_f }}
305 > \label{introEquation:ensembelAverage}
306 > \end{equation}
307 >
308 > There are several different types of ensembles with different
309 > statistical characteristics. As a function of macroscopic
310 > parameters, such as temperature \textit{etc}, partition function can
311 > be used to describe the statistical properties of a system in
312 > thermodynamic equilibrium.
313 >
314 > As an ensemble of systems, each of which is known to be thermally
315 > isolated and conserve energy, Microcanonical ensemble(NVE) has a
316 > partition function like,
317 > \begin{equation}
318 > \Omega (N,V,E) = e^{\beta TS} \label{introEquation:NVEPartition}.
319 > \end{equation}
320 > A canonical ensemble(NVT)is an ensemble of systems, each of which
321 > can share its energy with a large heat reservoir. The distribution
322 > of the total energy amongst the possible dynamical states is given
323 > by the partition function,
324 > \begin{equation}
325 > \Omega (N,V,T) = e^{ - \beta A}
326 > \label{introEquation:NVTPartition}
327 > \end{equation}
328 > Here, $A$ is the Helmholtz free energy which is defined as $ A = U -
329 > TS$. Since most experiment are carried out under constant pressure
330 > condition, isothermal-isobaric ensemble(NPT) play a very important
331 > role in molecular simulation. The isothermal-isobaric ensemble allow
332 > the system to exchange energy with a heat bath of temperature $T$
333 > and to change the volume as well. Its partition function is given as
334 > \begin{equation}
335 > \Delta (N,P,T) =  - e^{\beta G}.
336 > \label{introEquation:NPTPartition}
337 > \end{equation}
338 > Here, $G = U - TS + PV$, and $G$ is called Gibbs free energy.
339  
340 + \subsection{\label{introSection:liouville}Liouville's theorem}
341 +
342 + The Liouville's theorem is the foundation on which statistical
343 + mechanics rests. It describes the time evolution of phase space
344 + distribution function. In order to calculate the rate of change of
345 + $\rho$, we begin from Equation(\ref{introEquation:deltaN}). If we
346 + consider the two faces perpendicular to the $q_1$ axis, which are
347 + located at $q_1$ and $q_1 + \delta q_1$, the number of phase points
348 + leaving the opposite face is given by the expression,
349 + \begin{equation}
350 + \left( {\rho  + \frac{{\partial \rho }}{{\partial q_1 }}\delta q_1 }
351 + \right)\left( {\dot q_1  + \frac{{\partial \dot q_1 }}{{\partial q_1
352 + }}\delta q_1 } \right)\delta q_2  \ldots \delta q_f \delta p_1
353 + \ldots \delta p_f .
354 + \end{equation}
355 + Summing all over the phase space, we obtain
356 + \begin{equation}
357 + \frac{{d(\delta N)}}{{dt}} =  - \sum\limits_{i = 1}^f {\left[ {\rho
358 + \left( {\frac{{\partial \dot q_i }}{{\partial q_i }} +
359 + \frac{{\partial \dot p_i }}{{\partial p_i }}} \right) + \left(
360 + {\frac{{\partial \rho }}{{\partial q_i }}\dot q_i  + \frac{{\partial
361 + \rho }}{{\partial p_i }}\dot p_i } \right)} \right]} \delta q_1
362 + \ldots \delta q_f \delta p_1  \ldots \delta p_f .
363 + \end{equation}
364 + Differentiating the equations of motion in Hamiltonian formalism
365 + (\ref{introEquation:motionHamiltonianCoordinate},
366 + \ref{introEquation:motionHamiltonianMomentum}), we can show,
367 + \begin{equation}
368 + \sum\limits_i {\left( {\frac{{\partial \dot q_i }}{{\partial q_i }}
369 + + \frac{{\partial \dot p_i }}{{\partial p_i }}} \right)}  = 0 ,
370 + \end{equation}
371 + which cancels the first terms of the right hand side. Furthermore,
372 + divining $ \delta q_1  \ldots \delta q_f \delta p_1  \ldots \delta
373 + p_f $ in both sides, we can write out Liouville's theorem in a
374 + simple form,
375 + \begin{equation}
376 + \frac{{\partial \rho }}{{\partial t}} + \sum\limits_{i = 1}^f
377 + {\left( {\frac{{\partial \rho }}{{\partial q_i }}\dot q_i  +
378 + \frac{{\partial \rho }}{{\partial p_i }}\dot p_i } \right)}  = 0 .
379 + \label{introEquation:liouvilleTheorem}
380 + \end{equation}
381 +
382 + Liouville's theorem states that the distribution function is
383 + constant along any trajectory in phase space. In classical
384 + statistical mechanics, since the number of particles in the system
385 + is huge, we may be able to believe the system is stationary,
386 + \begin{equation}
387 + \frac{{\partial \rho }}{{\partial t}} = 0.
388 + \label{introEquation:stationary}
389 + \end{equation}
390 + In such stationary system, the density of distribution $\rho$ can be
391 + connected to the Hamiltonian $H$ through Maxwell-Boltzmann
392 + distribution,
393 + \begin{equation}
394 + \rho  \propto e^{ - \beta H}
395 + \label{introEquation:densityAndHamiltonian}
396 + \end{equation}
397 +
398 + \subsubsection{\label{introSection:phaseSpaceConservation}Conservation of Phase Space}
399 + Lets consider a region in the phase space,
400 + \begin{equation}
401 + \delta v = \int { \ldots \int {dq_1 } ...dq_f dp_1 } ..dp_f .
402 + \end{equation}
403 + If this region is small enough, the density $\rho$ can be regarded
404 + as uniform over the whole phase space. Thus, the number of phase
405 + points inside this region is given by,
406 + \begin{equation}
407 + \delta N = \rho \delta v = \rho \int { \ldots \int {dq_1 } ...dq_f
408 + dp_1 } ..dp_f.
409 + \end{equation}
410 +
411 + \begin{equation}
412 + \frac{{d(\delta N)}}{{dt}} = \frac{{d\rho }}{{dt}}\delta v + \rho
413 + \frac{d}{{dt}}(\delta v) = 0.
414 + \end{equation}
415 + With the help of stationary assumption
416 + (\ref{introEquation:stationary}), we obtain the principle of the
417 + \emph{conservation of extension in phase space},
418 + \begin{equation}
419 + \frac{d}{{dt}}(\delta v) = \frac{d}{{dt}}\int { \ldots \int {dq_1 }
420 + ...dq_f dp_1 } ..dp_f  = 0.
421 + \label{introEquation:volumePreserving}
422 + \end{equation}
423 +
424 + \subsubsection{\label{introSection:liouvilleInOtherForms}Liouville's Theorem in Other Forms}
425 +
426 + Liouville's theorem can be expresses in a variety of different forms
427 + which are convenient within different contexts. For any two function
428 + $F$ and $G$ of the coordinates and momenta of a system, the Poisson
429 + bracket ${F, G}$ is defined as
430 + \begin{equation}
431 + \left\{ {F,G} \right\} = \sum\limits_i {\left( {\frac{{\partial
432 + F}}{{\partial q_i }}\frac{{\partial G}}{{\partial p_i }} -
433 + \frac{{\partial F}}{{\partial p_i }}\frac{{\partial G}}{{\partial
434 + q_i }}} \right)}.
435 + \label{introEquation:poissonBracket}
436 + \end{equation}
437 + Substituting equations of motion in Hamiltonian formalism(
438 + \ref{introEquation:motionHamiltonianCoordinate} ,
439 + \ref{introEquation:motionHamiltonianMomentum} ) into
440 + (\ref{introEquation:liouvilleTheorem}), we can rewrite Liouville's
441 + theorem using Poisson bracket notion,
442 + \begin{equation}
443 + \left( {\frac{{\partial \rho }}{{\partial t}}} \right) =  - \left\{
444 + {\rho ,H} \right\}.
445 + \label{introEquation:liouvilleTheromInPoissin}
446 + \end{equation}
447 + Moreover, the Liouville operator is defined as
448 + \begin{equation}
449 + iL = \sum\limits_{i = 1}^f {\left( {\frac{{\partial H}}{{\partial
450 + p_i }}\frac{\partial }{{\partial q_i }} - \frac{{\partial
451 + H}}{{\partial q_i }}\frac{\partial }{{\partial p_i }}} \right)}
452 + \label{introEquation:liouvilleOperator}
453 + \end{equation}
454 + In terms of Liouville operator, Liouville's equation can also be
455 + expressed as
456 + \begin{equation}
457 + \left( {\frac{{\partial \rho }}{{\partial t}}} \right) =  - iL\rho
458 + \label{introEquation:liouvilleTheoremInOperator}
459 + \end{equation}
460 +
461   \subsection{\label{introSection:ergodic}The Ergodic Hypothesis}
462  
463   Various thermodynamic properties can be calculated from Molecular
# Line 261 | Line 472 | statistical ensemble are identical \cite{Frenkel1996,
472   ensemble average. It states that time average and average over the
473   statistical ensemble are identical \cite{Frenkel1996, leach01:mm}.
474   \begin{equation}
475 < \langle A \rangle_t = \mathop {\lim }\limits_{t \to \infty }
476 < \frac{1}{t}\int\limits_0^t {A(p(t),q(t))dt = \int\limits_\Gamma
477 < {A(p(t),q(t))} } \rho (p(t), q(t)) dpdq
475 > \langle A(q , p) \rangle_t = \mathop {\lim }\limits_{t \to \infty }
476 > \frac{1}{t}\int\limits_0^t {A(q(t),p(t))dt = \int\limits_\Gamma
477 > {A(q(t),p(t))} } \rho (q(t), p(t)) dqdp
478   \end{equation}
479 < where $\langle A \rangle_t$ is an equilibrium value of a physical
480 < quantity and $\rho (p(t), q(t))$ is the equilibrium distribution
481 < function. If an observation is averaged over a sufficiently long
482 < time (longer than relaxation time), all accessible microstates in
483 < phase space are assumed to be equally probed, giving a properly
484 < weighted statistical average. This allows the researcher freedom of
485 < choice when deciding how best to measure a given observable. In case
486 < an ensemble averaged approach sounds most reasonable, the Monte
487 < Carlo techniques\cite{metropolis:1949} can be utilized. Or if the
488 < system lends itself to a time averaging approach, the Molecular
489 < Dynamics techniques in Sec.~\ref{introSection:molecularDynamics}
490 < will be the best choice\cite{Frenkel1996}.
479 > where $\langle  A(q , p) \rangle_t$ is an equilibrium value of a
480 > physical quantity and $\rho (p(t), q(t))$ is the equilibrium
481 > distribution function. If an observation is averaged over a
482 > sufficiently long time (longer than relaxation time), all accessible
483 > microstates in phase space are assumed to be equally probed, giving
484 > a properly weighted statistical average. This allows the researcher
485 > freedom of choice when deciding how best to measure a given
486 > observable. In case an ensemble averaged approach sounds most
487 > reasonable, the Monte Carlo techniques\cite{metropolis:1949} can be
488 > utilized. Or if the system lends itself to a time averaging
489 > approach, the Molecular Dynamics techniques in
490 > Sec.~\ref{introSection:molecularDynamics} will be the best
491 > choice\cite{Frenkel1996}.
492  
493   \section{\label{introSection:geometricIntegratos}Geometric Integrators}
494   A variety of numerical integrators were proposed to simulate the
# Line 324 | Line 536 | classical mechanics. According to Liouville's theorem,
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. According to Liouville's theorem, the
328 < Hamiltonian \emph{flow} or \emph{symplectomorphism} generated by the
329 < Hamiltonian vector filed preserves the volume form on the phase
330 < space, which is the basis of classical statistical mechanics.
539 > classical mechanics.
540  
541 < \subsection{\label{introSection:exactFlow}The Exact Flow of ODE}
541 > \subsection{\label{introSection:ODE}Ordinary Differential Equations}
542  
543 < \subsection{\label{introSection:hamiltonianSplitting}Hamiltonian Splitting}
543 > For a ordinary differential system defined as
544 > \begin{equation}
545 > \dot x = f(x)
546 > \end{equation}
547 > where $x = x(q,p)^T$, this system is canonical Hamiltonian, if
548 > \begin{equation}
549 > f(r) = J\nabla _x H(r).
550 > \end{equation}
551 > $H = H (q, p)$ is Hamiltonian function and $J$ is the skew-symmetric
552 > matrix
553 > \begin{equation}
554 > J = \left( {\begin{array}{*{20}c}
555 >   0 & I  \\
556 >   { - I} & 0  \\
557 > \end{array}} \right)
558 > \label{introEquation:canonicalMatrix}
559 > \end{equation}
560 > where $I$ is an identity matrix. Using this notation, Hamiltonian
561 > system can be rewritten as,
562 > \begin{equation}
563 > \frac{d}{{dt}}x = J\nabla _x H(x)
564 > \label{introEquation:compactHamiltonian}
565 > \end{equation}In this case, $f$ is
566 > called a \emph{Hamiltonian vector field}.
567  
568 < \section{\label{introSection:molecularDynamics}Molecular Dynamics}
568 > Another generalization of Hamiltonian dynamics is Poisson Dynamics,
569 > \begin{equation}
570 > \dot x = J(x)\nabla _x H \label{introEquation:poissonHamiltonian}
571 > \end{equation}
572 > The most obvious change being that matrix $J$ now depends on $x$.
573  
574 < As a special discipline of molecular modeling, Molecular dynamics
339 < has proven to be a powerful tool for studying the functions of
340 < biological systems, providing structural, thermodynamic and
341 < dynamical information.
574 > \subsection{\label{introSection:exactFlow}Exact Flow}
575  
576 < \subsection{\label{introSec:mdInit}Initialization}
576 > Let $x(t)$ be the exact solution of the ODE system,
577 > \begin{equation}
578 > \frac{{dx}}{{dt}} = f(x) \label{introEquation:ODE}
579 > \end{equation}
580 > The exact flow(solution) $\varphi_\tau$ is defined by
581 > \[
582 > x(t+\tau) =\varphi_\tau(x(t))
583 > \]
584 > where $\tau$ is a fixed time step and $\varphi$ is a map from phase
585 > space to itself. The flow has the continuous group property,
586 > \begin{equation}
587 > \varphi _{\tau _1 }  \circ \varphi _{\tau _2 }  = \varphi _{\tau _1
588 > + \tau _2 } .
589 > \end{equation}
590 > In particular,
591 > \begin{equation}
592 > \varphi _\tau   \circ \varphi _{ - \tau }  = I
593 > \end{equation}
594 > Therefore, the exact flow is self-adjoint,
595 > \begin{equation}
596 > \varphi _\tau   = \varphi _{ - \tau }^{ - 1}.
597 > \end{equation}
598 > The exact flow can also be written in terms of the of an operator,
599 > \begin{equation}
600 > \varphi _\tau  (x) = e^{\tau \sum\limits_i {f_i (x)\frac{\partial
601 > }{{\partial x_i }}} } (x) \equiv \exp (\tau f)(x).
602 > \label{introEquation:exponentialOperator}
603 > \end{equation}
604  
605 < \subsection{\label{introSection:mdIntegration} Integration of the Equations of Motion}
605 > In most cases, it is not easy to find the exact flow $\varphi_\tau$.
606 > Instead, we use a approximate map, $\psi_\tau$, which is usually
607 > called integrator. The order of an integrator $\psi_\tau$ is $p$, if
608 > the Taylor series of $\psi_\tau$ agree to order $p$,
609 > \begin{equation}
610 > \psi_tau(x) = x + \tau f(x) + O(\tau^{p+1})
611 > \end{equation}
612  
613 < \section{\label{introSection:rigidBody}Dynamics of Rigid Bodies}
613 > \subsection{\label{introSection:geometricProperties}Geometric Properties}
614  
615 < A rigid body is a body in which the distance between any two given
616 < points of a rigid body remains constant regardless of external
617 < forces exerted on it. A rigid body therefore conserves its shape
352 < during its motion.
615 > The hidden geometric properties of ODE and its flow play important
616 > roles in numerical studies. Many of them can be found in systems
617 > which occur naturally in applications.
618  
619 < Applications of dynamics of rigid bodies.
620 <
356 < \subsection{\label{introSection:lieAlgebra}Lie Algebra}
357 <
358 < \subsection{\label{introSection:DLMMotionEquation}The Euler Equations of Rigid Body Motion}
359 <
360 < \subsection{\label{introSection:otherRBMotionEquation}Other Formulations for Rigid Body Motion}
361 <
362 < %\subsection{\label{introSection:poissonBrackets}Poisson Brackets}
363 <
364 < \section{\label{introSection:correlationFunctions}Correlation Functions}
365 <
366 < \section{\label{introSection:langevinDynamics}Langevin Dynamics}
367 <
368 < \subsection{\label{introSection:LDIntroduction}Introduction and application of Langevin Dynamics}
369 <
370 < \subsection{\label{introSection:generalizedLangevinDynamics}Generalized Langevin Dynamics}
371 <
619 > Let $\varphi$ be the flow of Hamiltonian vector field, $\varphi$ is
620 > a \emph{symplectic} flow if it satisfies,
621   \begin{equation}
622 < H = \frac{{p^2 }}{{2m}} + U(x) + H_B  + \Delta U(x,x_1 , \ldots x_N)
374 < \label{introEquation:bathGLE}
622 > {\varphi '}^T J \varphi ' = J.
623   \end{equation}
624 < where $H_B$ is harmonic bath Hamiltonian,
624 > According to Liouville's theorem, the symplectic volume is invariant
625 > under a Hamiltonian flow, which is the basis for classical
626 > statistical mechanics. Furthermore, the flow of a Hamiltonian vector
627 > field on a symplectic manifold can be shown to be a
628 > symplectomorphism. As to the Poisson system,
629 > \begin{equation}
630 > {\varphi '}^T J \varphi ' = J \circ \varphi
631 > \end{equation}
632 > is the property must be preserved by the integrator.
633 >
634 > It is possible to construct a \emph{volume-preserving} flow for a
635 > source free($ \nabla \cdot f = 0 $) ODE, if the flow satisfies $
636 > \det d\varphi  = 1$. One can show easily that a symplectic flow will
637 > be volume-preserving.
638 >
639 > Changing the variables $y = h(x)$ in a ODE\ref{introEquation:ODE}
640 > will result in a new system,
641   \[
642 < H_B  =\sum\limits_{\alpha  = 1}^N {\left\{ {\frac{{p_\alpha ^2
379 < }}{{2m_\alpha  }} + \frac{1}{2}m_\alpha  w_\alpha ^2 } \right\}}
642 > \dot y = \tilde f(y) = ((dh \cdot f)h^{ - 1} )(y).
643   \]
644 < and $\Delta U$ is bilinear system-bath coupling,
644 > The vector filed $f$ has reversing symmetry $h$ if $f = - \tilde f$.
645 > In other words, the flow of this vector field is reversible if and
646 > only if $ h \circ \varphi ^{ - 1}  = \varphi  \circ h $.
647 >
648 > A \emph{first integral}, or conserved quantity of a general
649 > differential function is a function $ G:R^{2d}  \to R^d $ which is
650 > constant for all solutions of the ODE $\frac{{dx}}{{dt}} = f(x)$ ,
651   \[
652 < \Delta U =  - \sum\limits_{\alpha  = 1}^N {g_\alpha  x_\alpha  x}
652 > \frac{{dG(x(t))}}{{dt}} = 0.
653   \]
654 < Completing the square,
654 > Using chain rule, one may obtain,
655   \[
656 < H_B  + \Delta U = \sum\limits_{\alpha  = 1}^N {\left\{
388 < {\frac{{p_\alpha ^2 }}{{2m_\alpha  }} + \frac{1}{2}m_\alpha
389 < w_\alpha ^2 \left( {x_\alpha   - \frac{{g_\alpha  }}{{m_\alpha
390 < w_\alpha ^2 }}x} \right)^2 } \right\}}  - \sum\limits_{\alpha  =
391 < 1}^N {\frac{{g_\alpha ^2 }}{{2m_\alpha  w_\alpha ^2 }}} x^2
656 > \sum\limits_i {\frac{{dG}}{{dx_i }}} f_i (x) = f \bullet \nabla G,
657   \]
658 < and putting it back into Eq.~\ref{introEquation:bathGLE},
658 > which is the condition for conserving \emph{first integral}. For a
659 > canonical Hamiltonian system, the time evolution of an arbitrary
660 > smooth function $G$ is given by,
661 > \begin{equation}
662 > \begin{array}{c}
663 > \frac{{dG(x(t))}}{{dt}} = [\nabla _x G(x(t))]^T \dot x(t) \\
664 >  = [\nabla _x G(x(t))]^T J\nabla _x H(x(t)). \\
665 > \end{array}
666 > \label{introEquation:firstIntegral1}
667 > \end{equation}
668 > Using poisson bracket notion, Equation
669 > \ref{introEquation:firstIntegral1} can be rewritten as
670   \[
671 < H = \frac{{p^2 }}{{2m}} + W(x) + \sum\limits_{\alpha  = 1}^N
396 < {\left\{ {\frac{{p_\alpha ^2 }}{{2m_\alpha  }} + \frac{1}{2}m_\alpha
397 < w_\alpha ^2 \left( {x_\alpha   - \frac{{g_\alpha  }}{{m_\alpha
398 < w_\alpha ^2 }}x} \right)^2 } \right\}}
671 > \frac{d}{{dt}}G(x(t)) = \left\{ {G,H} \right\}(x(t)).
672   \]
673 < where
673 > Therefore, the sufficient condition for $G$ to be the \emph{first
674 > integral} of a Hamiltonian system is
675   \[
676 < W(x) = U(x) - \sum\limits_{\alpha  = 1}^N {\frac{{g_\alpha ^2
403 < }}{{2m_\alpha  w_\alpha ^2 }}} x^2
676 > \left\{ {G,H} \right\} = 0.
677   \]
678 < Since the first two terms of the new Hamiltonian depend only on the
679 < system coordinates, we can get the equations of motion for
680 < Generalized Langevin Dynamics by Hamilton's equations
681 < \ref{introEquation:motionHamiltonianCoordinate,
682 < introEquation:motionHamiltonianMomentum},
678 > As well known, the Hamiltonian (or energy) H of a Hamiltonian system
679 > is a \emph{first integral}, which is due to the fact $\{ H,H\}  =
680 > 0$.
681 >
682 >
683 > When designing any numerical methods, one should always try to
684 > preserve the structural properties of the original ODE and its flow.
685 >
686 > \subsection{\label{introSection:constructionSymplectic}Construction of Symplectic Methods}
687 > A lot of well established and very effective numerical methods have
688 > been successful precisely because of their symplecticities even
689 > though this fact was not recognized when they were first
690 > constructed. The most famous example is leapfrog methods in
691 > molecular dynamics. In general, symplectic integrators can be
692 > constructed using one of four different methods.
693 > \begin{enumerate}
694 > \item Generating functions
695 > \item Variational methods
696 > \item Runge-Kutta methods
697 > \item Splitting methods
698 > \end{enumerate}
699 >
700 > Generating function tends to lead to methods which are cumbersome
701 > and difficult to use. In dissipative systems, variational methods
702 > can capture the decay of energy accurately. Since their
703 > geometrically unstable nature against non-Hamiltonian perturbations,
704 > ordinary implicit Runge-Kutta methods are not suitable for
705 > Hamiltonian system. Recently, various high-order explicit
706 > Runge--Kutta methods have been developed to overcome this
707 > instability. However, due to computational penalty involved in
708 > implementing the Runge-Kutta methods, they do not attract too much
709 > attention from Molecular Dynamics community. Instead, splitting have
710 > been widely accepted since they exploit natural decompositions of
711 > the system\cite{Tuckerman92}.
712 >
713 > \subsubsection{\label{introSection:splittingMethod}Splitting Method}
714 >
715 > The main idea behind splitting methods is to decompose the discrete
716 > $\varphi_h$ as a composition of simpler flows,
717 > \begin{equation}
718 > \varphi _h  = \varphi _{h_1 }  \circ \varphi _{h_2 }  \ldots  \circ
719 > \varphi _{h_n }
720 > \label{introEquation:FlowDecomposition}
721 > \end{equation}
722 > where each of the sub-flow is chosen such that each represent a
723 > simpler integration of the system.
724 >
725 > Suppose that a Hamiltonian system takes the form,
726 > \[
727 > H = H_1 + H_2.
728 > \]
729 > Here, $H_1$ and $H_2$ may represent different physical processes of
730 > the system. For instance, they may relate to kinetic and potential
731 > energy respectively, which is a natural decomposition of the
732 > problem. If $H_1$ and $H_2$ can be integrated using exact flows
733 > $\varphi_1(t)$ and $\varphi_2(t)$, respectively, a simple first
734 > order is then given by the Lie-Trotter formula
735 > \begin{equation}
736 > \varphi _h  = \varphi _{1,h}  \circ \varphi _{2,h},
737 > \label{introEquation:firstOrderSplitting}
738 > \end{equation}
739 > where $\varphi _h$ is the result of applying the corresponding
740 > continuous $\varphi _i$ over a time $h$. By definition, as
741 > $\varphi_i(t)$ is the exact solution of a Hamiltonian system, it
742 > must follow that each operator $\varphi_i(t)$ is a symplectic map.
743 > It is easy to show that any composition of symplectic flows yields a
744 > symplectic map,
745 > \begin{equation}
746 > (\varphi '\phi ')^T J\varphi '\phi ' = \phi '^T \varphi '^T J\varphi
747 > '\phi ' = \phi '^T J\phi ' = J,
748 > \label{introEquation:SymplecticFlowComposition}
749 > \end{equation}
750 > where $\phi$ and $\psi$ both are symplectic maps. Thus operator
751 > splitting in this context automatically generates a symplectic map.
752 >
753 > The Lie-Trotter splitting(\ref{introEquation:firstOrderSplitting})
754 > introduces local errors proportional to $h^2$, while Strang
755 > splitting gives a second-order decomposition,
756 > \begin{equation}
757 > \varphi _h  = \varphi _{1,h/2}  \circ \varphi _{2,h}  \circ \varphi
758 > _{1,h/2} , \label{introEquation:secondOrderSplitting}
759 > \end{equation}
760 > which has a local error proportional to $h^3$. Sprang splitting's
761 > popularity in molecular simulation community attribute to its
762 > symmetric property,
763 > \begin{equation}
764 > \varphi _h^{ - 1} = \varphi _{ - h}.
765 > \label{introEquation:timeReversible}
766 > \end{equation}
767 >
768 > \subsubsection{\label{introSection:exampleSplittingMethod}Example of Splitting Method}
769 > The classical equation for a system consisting of interacting
770 > particles can be written in Hamiltonian form,
771 > \[
772 > H = T + V
773 > \]
774 > where $T$ is the kinetic energy and $V$ is the potential energy.
775 > Setting $H_1 = T, H_2 = V$ and applying Strang splitting, one
776 > obtains the following:
777   \begin{align}
778 < \dot p &=  - \frac{{\partial H}}{{\partial x}}
779 <       &= m\ddot x
780 <       &= - \frac{{\partial W(x)}}{{\partial x}} - \sum\limits_{\alpha  = 1}^N {g_\alpha  \left( {x_\alpha   - \frac{{g_\alpha  }}{{m_\alpha  w_\alpha ^2 }}x} \right)}
781 < \label{introEq:Lp5}
778 > q(\Delta t) &= q(0) + \dot{q}(0)\Delta t +
779 >    \frac{F[q(0)]}{m}\frac{\Delta t^2}{2}, %
780 > \label{introEquation:Lp10a} \\%
781 > %
782 > \dot{q}(\Delta t) &= \dot{q}(0) + \frac{\Delta t}{2m}
783 >    \biggl [F[q(0)] + F[q(\Delta t)] \biggr]. %
784 > \label{introEquation:Lp10b}
785   \end{align}
786 < , and
786 > where $F(t)$ is the force at time $t$. This integration scheme is
787 > known as \emph{velocity verlet} which is
788 > symplectic(\ref{introEquation:SymplecticFlowComposition}),
789 > time-reversible(\ref{introEquation:timeReversible}) and
790 > volume-preserving (\ref{introEquation:volumePreserving}). These
791 > geometric properties attribute to its long-time stability and its
792 > popularity in the community. However, the most commonly used
793 > velocity verlet integration scheme is written as below,
794   \begin{align}
795 < \dot p_\alpha   &=  - \frac{{\partial H}}{{\partial x_\alpha  }}
796 <                &= m\ddot x_\alpha
797 <                &= \- m_\alpha  w_\alpha ^2 \left( {x_\alpha   - \frac{{g_\alpha}}{{m_\alpha  w_\alpha ^2 }}x} \right)
795 > \dot{q}\biggl (\frac{\Delta t}{2}\biggr ) &=
796 >    \dot{q}(0) + \frac{\Delta t}{2m}\, F[q(0)], \label{introEquation:Lp9a}\\%
797 > %
798 > q(\Delta t) &= q(0) + \Delta t\, \dot{q}\biggl (\frac{\Delta t}{2}\biggr ),%
799 >    \label{introEquation:Lp9b}\\%
800 > %
801 > \dot{q}(\Delta t) &= \dot{q}\biggl (\frac{\Delta t}{2}\biggr ) +
802 >    \frac{\Delta t}{2m}\, F[q(0)]. \label{introEquation:Lp9c}
803   \end{align}
804 + From the preceding splitting, one can see that the integration of
805 + the equations of motion would follow:
806 + \begin{enumerate}
807 + \item calculate the velocities at the half step, $\frac{\Delta t}{2}$, from the forces calculated at the initial position.
808  
809 < \subsection{\label{introSection:laplaceTransform}The Laplace Transform}
809 > \item Use the half step velocities to move positions one whole step, $\Delta t$.
810  
811 + \item Evaluate the forces at the new positions, $\mathbf{r}(\Delta t)$, and use the new forces to complete the velocity move.
812 +
813 + \item Repeat from step 1 with the new position, velocities, and forces assuming the roles of the initial values.
814 + \end{enumerate}
815 +
816 + Simply switching the order of splitting and composing, a new
817 + integrator, the \emph{position verlet} integrator, can be generated,
818 + \begin{align}
819 + \dot q(\Delta t) &= \dot q(0) + \Delta tF(q(0))\left[ {q(0) +
820 + \frac{{\Delta t}}{{2m}}\dot q(0)} \right], %
821 + \label{introEquation:positionVerlet1} \\%
822 + %
823 + q(\Delta t) &= q(0) + \frac{{\Delta t}}{2}\left[ {\dot q(0) + \dot
824 + q(\Delta t)} \right]. %
825 + \label{introEquation:positionVerlet2}
826 + \end{align}
827 +
828 + \subsubsection{\label{introSection:errorAnalysis}Error Analysis and Higher Order Methods}
829 +
830 + Baker-Campbell-Hausdorff formula can be used to determine the local
831 + error of splitting method in terms of commutator of the
832 + operators(\ref{introEquation:exponentialOperator}) associated with
833 + the sub-flow. For operators $hX$ and $hY$ which are associate to
834 + $\varphi_1(t)$ and $\varphi_2(t)$ respectively , we have
835 + \begin{equation}
836 + \exp (hX + hY) = \exp (hZ)
837 + \end{equation}
838 + where
839 + \begin{equation}
840 + hZ = hX + hY + \frac{{h^2 }}{2}[X,Y] + \frac{{h^3 }}{2}\left(
841 + {[X,[X,Y]] + [Y,[Y,X]]} \right) +  \ldots .
842 + \end{equation}
843 + Here, $[X,Y]$ is the commutators of operator $X$ and $Y$ given by
844   \[
845 < L(x) = \int_0^\infty  {x(t)e^{ - pt} dt}
845 > [X,Y] = XY - YX .
846   \]
847 <
847 > Applying Baker-Campbell-Hausdorff formula to Sprang splitting, we
848 > can obtain
849 > \begin{eqnarray*}
850 > \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 \\
851 >                                   &   & \mbox{} + h^2 [X,X]/8 + h^2 [Y,Y]/8 \\
852 >                                   &   & \mbox{} + h^3 [Y,[Y,X]]/12 - h^3[X,[X,Y]]/24 + \ldots )
853 > \end{eqnarray*}
854 > Since \[ [X,Y] + [Y,X] = 0\] and \[ [X,X] = 0\], the dominant local
855 > error of Spring splitting is proportional to $h^3$. The same
856 > procedure can be applied to general splitting,  of the form
857 > \begin{equation}
858 > \varphi _{b_m h}^2  \circ \varphi _{a_m h}^1  \circ \varphi _{b_{m -
859 > 1} h}^2  \circ  \ldots  \circ \varphi _{a_1 h}^1 .
860 > \end{equation}
861 > Careful choice of coefficient $a_1 \ldot b_m$ will lead to higher
862 > order method. Yoshida proposed an elegant way to compose higher
863 > order methods based on symmetric splitting. Given a symmetric second
864 > order base method $ \varphi _h^{(2)} $, a fourth-order symmetric
865 > method can be constructed by composing,
866   \[
867 < L(x + y) = L(x) + L(y)
867 > \varphi _h^{(4)}  = \varphi _{\alpha h}^{(2)}  \circ \varphi _{\beta
868 > h}^{(2)}  \circ \varphi _{\alpha h}^{(2)}
869   \]
870 <
870 > where $ \alpha  =  - \frac{{2^{1/3} }}{{2 - 2^{1/3} }}$ and $ \beta
871 > = \frac{{2^{1/3} }}{{2 - 2^{1/3} }}$. Moreover, a symmetric
872 > integrator $ \varphi _h^{(2n + 2)}$ can be composed by
873 > \begin{equation}
874 > \varphi _h^{(2n + 2)}  = \varphi _{\alpha h}^{(2n)}  \circ \varphi
875 > _{\beta h}^{(2n)}  \circ \varphi _{\alpha h}^{(2n)}
876 > \end{equation}
877 > , if the weights are chosen as
878   \[
879 < L(ax) = aL(x)
879 > \alpha  =  - \frac{{2^{1/(2n + 1)} }}{{2 - 2^{1/(2n + 1)} }},\beta =
880 > \frac{{2^{1/(2n + 1)} }}{{2 - 2^{1/(2n + 1)} }} .
881   \]
882  
883 + \section{\label{introSection:molecularDynamics}Molecular Dynamics}
884 +
885 + As one of the principal tools of molecular modeling, Molecular
886 + dynamics has proven to be a powerful tool for studying the functions
887 + of biological systems, providing structural, thermodynamic and
888 + dynamical information. The basic idea of molecular dynamics is that
889 + macroscopic properties are related to microscopic behavior and
890 + microscopic behavior can be calculated from the trajectories in
891 + simulations. For instance, instantaneous temperature of an
892 + Hamiltonian system of $N$ particle can be measured by
893   \[
894 < L(\dot x) = pL(x) - px(0)
894 > T = \sum\limits_{i = 1}^N {\frac{{m_i v_i^2 }}{{fk_B }}}
895   \]
896 + where $m_i$ and $v_i$ are the mass and velocity of $i$th particle
897 + respectively, $f$ is the number of degrees of freedom, and $k_B$ is
898 + the boltzman constant.
899  
900 + A typical molecular dynamics run consists of three essential steps:
901 + \begin{enumerate}
902 +  \item Initialization
903 +    \begin{enumerate}
904 +    \item Preliminary preparation
905 +    \item Minimization
906 +    \item Heating
907 +    \item Equilibration
908 +    \end{enumerate}
909 +  \item Production
910 +  \item Analysis
911 + \end{enumerate}
912 + These three individual steps will be covered in the following
913 + sections. Sec.~\ref{introSec:initialSystemSettings} deals with the
914 + initialization of a simulation. Sec.~\ref{introSec:production} will
915 + discusses issues in production run. Sec.~\ref{introSection:Analysis}
916 + provides the theoretical tools for trajectory analysis.
917 +
918 + \subsection{\label{introSec:initialSystemSettings}Initialization}
919 +
920 + \subsubsection{Preliminary preparation}
921 +
922 + When selecting the starting structure of a molecule for molecular
923 + simulation, one may retrieve its Cartesian coordinates from public
924 + databases, such as RCSB Protein Data Bank \textit{etc}. Although
925 + thousands of crystal structures of molecules are discovered every
926 + year, many more remain unknown due to the difficulties of
927 + purification and crystallization. Even for the molecule with known
928 + structure, some important information is missing. For example, the
929 + missing hydrogen atom which acts as donor in hydrogen bonding must
930 + be added. Moreover, in order to include electrostatic interaction,
931 + one may need to specify the partial charges for individual atoms.
932 + Under some circumstances, we may even need to prepare the system in
933 + a special setup. For instance, when studying transport phenomenon in
934 + membrane system, we may prepare the lipids in bilayer structure
935 + instead of placing lipids randomly in solvent, since we are not
936 + interested in self-aggregation and it takes a long time to happen.
937 +
938 + \subsubsection{Minimization}
939 +
940 + It is quite possible that some of molecules in the system from
941 + preliminary preparation may be overlapped with each other. This
942 + close proximity leads to high potential energy which consequently
943 + jeopardizes any molecular dynamics simulations. To remove these
944 + steric overlaps, one typically performs energy minimization to find
945 + a more reasonable conformation. Several energy minimization methods
946 + have been developed to exploit the energy surface and to locate the
947 + local minimum. While converging slowly near the minimum, steepest
948 + descent method is extremely robust when systems are far from
949 + harmonic. Thus, it is often used to refine structure from
950 + crystallographic data. Relied on the gradient or hessian, advanced
951 + methods like conjugate gradient and Newton-Raphson converge rapidly
952 + to a local minimum, while become unstable if the energy surface is
953 + far from quadratic. Another factor must be taken into account, when
954 + choosing energy minimization method, is the size of the system.
955 + Steepest descent and conjugate gradient can deal with models of any
956 + size. Because of the limit of computation power to calculate hessian
957 + matrix and insufficient storage capacity to store them, most
958 + Newton-Raphson methods can not be used with very large models.
959 +
960 + \subsubsection{Heating}
961 +
962 + Typically, Heating is performed by assigning random velocities
963 + according to a Gaussian distribution for a temperature. Beginning at
964 + a lower temperature and gradually increasing the temperature by
965 + assigning greater random velocities, we end up with setting the
966 + temperature of the system to a final temperature at which the
967 + simulation will be conducted. In heating phase, we should also keep
968 + the system from drifting or rotating as a whole. Equivalently, the
969 + net linear momentum and angular momentum of the system should be
970 + shifted to zero.
971 +
972 + \subsubsection{Equilibration}
973 +
974 + The purpose of equilibration is to allow the system to evolve
975 + spontaneously for a period of time and reach equilibrium. The
976 + procedure is continued until various statistical properties, such as
977 + temperature, pressure, energy, volume and other structural
978 + properties \textit{etc}, become independent of time. Strictly
979 + speaking, minimization and heating are not necessary, provided the
980 + equilibration process is long enough. However, these steps can serve
981 + as a means to arrive at an equilibrated structure in an effective
982 + way.
983 +
984 + \subsection{\label{introSection:production}Production}
985 +
986 + Production run is the most important steps of the simulation, in
987 + which the equilibrated structure is used as a starting point and the
988 + motions of the molecules are collected for later analysis. In order
989 + to capture the macroscopic properties of the system, the molecular
990 + dynamics simulation must be performed in correct and efficient way.
991 +
992 + The most expensive part of a molecular dynamics simulation is the
993 + calculation of non-bonded forces, such as van der Waals force and
994 + Coulombic forces \textit{etc}. For a system of $N$ particles, the
995 + complexity of the algorithm for pair-wise interactions is $O(N^2 )$,
996 + which making large simulations prohibitive in the absence of any
997 + computation saving techniques.
998 +
999 + A natural approach to avoid system size issue is to represent the
1000 + bulk behavior by a finite number of the particles. However, this
1001 + approach will suffer from the surface effect. To offset this,
1002 + \textit{Periodic boundary condition} is developed to simulate bulk
1003 + properties with a relatively small number of particles. In this
1004 + method, the simulation box is replicated throughout space to form an
1005 + infinite lattice. During the simulation, when a particle moves in
1006 + the primary cell, its image in other cells move in exactly the same
1007 + direction with exactly the same orientation. Thus, as a particle
1008 + leaves the primary cell, one of its images will enter through the
1009 + opposite face.
1010 + %\begin{figure}
1011 + %\centering
1012 + %\includegraphics[width=\linewidth]{pbcFig.eps}
1013 + %\caption[An illustration of periodic boundary conditions]{A 2-D
1014 + %illustration of periodic boundary conditions. As one particle leaves
1015 + %the right of the simulation box, an image of it enters the left.}
1016 + %\label{introFig:pbc}
1017 + %\end{figure}
1018 +
1019 + %cutoff and minimum image convention
1020 + Another important technique to improve the efficiency of force
1021 + evaluation is to apply cutoff where particles farther than a
1022 + predetermined distance, are not included in the calculation
1023 + \cite{Frenkel1996}. The use of a cutoff radius will cause a
1024 + discontinuity in the potential energy curve. Fortunately, one can
1025 + shift the potential to ensure the potential curve go smoothly to
1026 + zero at the cutoff radius. Cutoff strategy works pretty well for
1027 + Lennard-Jones interaction because of its short range nature.
1028 + However, simply truncating the electrostatic interaction with the
1029 + use of cutoff has been shown to lead to severe artifacts in
1030 + simulations. Ewald summation, in which the slowly conditionally
1031 + convergent Coulomb potential is transformed into direct and
1032 + reciprocal sums with rapid and absolute convergence, has proved to
1033 + minimize the periodicity artifacts in liquid simulations. Taking the
1034 + advantages of the fast Fourier transform (FFT) for calculating
1035 + discrete Fourier transforms, the particle mesh-based methods are
1036 + accelerated from $O(N^{3/2})$ to $O(N logN)$. An alternative
1037 + approach is \emph{fast multipole method}, which treats Coulombic
1038 + interaction exactly at short range, and approximate the potential at
1039 + long range through multipolar expansion. In spite of their wide
1040 + acceptances at the molecular simulation community, these two methods
1041 + are hard to be implemented correctly and efficiently. Instead, we
1042 + use a damped and charge-neutralized Coulomb potential method
1043 + developed by Wolf and his coworkers. The shifted Coulomb potential
1044 + for particle $i$ and particle $j$ at distance $r_{rj}$ is given by:
1045 + \begin{equation}
1046 + V(r_{ij})= \frac{q_i q_j \textrm{erfc}(\alpha
1047 + r_{ij})}{r_{ij}}-\lim_{r_{ij}\rightarrow
1048 + R_\textrm{c}}\left\{\frac{q_iq_j \textrm{erfc}(\alpha
1049 + r_{ij})}{r_{ij}}\right\}. \label{introEquation:shiftedCoulomb}
1050 + \end{equation}
1051 + where $\alpha$ is the convergence parameter. Due to the lack of
1052 + inherent periodicity and rapid convergence,this method is extremely
1053 + efficient and easy to implement.
1054 + %\begin{figure}
1055 + %\centering
1056 + %\includegraphics[width=\linewidth]{pbcFig.eps}
1057 + %\caption[An illustration of shifted Coulomb potential]{An illustration of shifted Coulomb potential.}
1058 + %\label{introFigure:shiftedCoulomb}
1059 + %\end{figure}
1060 +
1061 + %multiple time step
1062 +
1063 + \subsection{\label{introSection:Analysis} Analysis}
1064 +
1065 + Recently, advanced visualization technique are widely applied to
1066 + monitor the motions of molecules. Although the dynamics of the
1067 + system can be described qualitatively from animation, quantitative
1068 + trajectory analysis are more appreciable. According to the
1069 + principles of Statistical Mechanics,
1070 + Sec.~\ref{introSection:statisticalMechanics}, one can compute
1071 + thermodynamics properties, analyze fluctuations of structural
1072 + parameters, and investigate time-dependent processes of the molecule
1073 + from the trajectories.
1074 +
1075 + \subsubsection{\label{introSection:thermodynamicsProperties}Thermodynamics Properties}
1076 +
1077 + Thermodynamics properties, which can be expressed in terms of some
1078 + function of the coordinates and momenta of all particles in the
1079 + system, can be directly computed from molecular dynamics. The usual
1080 + way to measure the pressure is based on virial theorem of Clausius
1081 + which states that the virial is equal to $-3Nk_BT$. For a system
1082 + with forces between particles, the total virial, $W$, contains the
1083 + contribution from external pressure and interaction between the
1084 + particles:
1085   \[
1086 < L(\ddot x) = p^2 L(x) - px(0) - \dot x(0)
1086 > W =  - 3PV + \left\langle {\sum\limits_{i < j} {r{}_{ij} \cdot
1087 > f_{ij} } } \right\rangle
1088   \]
1089 + where $f_{ij}$ is the force between particle $i$ and $j$ at a
1090 + distance $r_{ij}$. Thus, the expression for the pressure is given
1091 + by:
1092 + \begin{equation}
1093 + P = \frac{{Nk_B T}}{V} - \frac{1}{{3V}}\left\langle {\sum\limits_{i
1094 + < j} {r{}_{ij} \cdot f_{ij} } } \right\rangle
1095 + \end{equation}
1096  
1097 + \subsubsection{\label{introSection:structuralProperties}Structural Properties}
1098 +
1099 + Structural Properties of a simple fluid can be described by a set of
1100 + distribution functions. Among these functions,\emph{pair
1101 + distribution function}, also known as \emph{radial distribution
1102 + function}, is of most fundamental importance to liquid-state theory.
1103 + Pair distribution function can be gathered by Fourier transforming
1104 + raw data from a series of neutron diffraction experiments and
1105 + integrating over the surface factor \cite{Powles73}. The experiment
1106 + result can serve as a criterion to justify the correctness of the
1107 + theory. Moreover, various equilibrium thermodynamic and structural
1108 + properties can also be expressed in terms of radial distribution
1109 + function \cite{allen87:csl}.
1110 +
1111 + A pair distribution functions $g(r)$ gives the probability that a
1112 + particle $i$ will be located at a distance $r$ from a another
1113 + particle $j$ in the system
1114   \[
1115 < L\left( {\int_0^t {g(t - \tau )h(\tau )d\tau } } \right) = G(p)H(p)
1115 > g(r) = \frac{V}{{N^2 }}\left\langle {\sum\limits_i {\sum\limits_{j
1116 > \ne i} {\delta (r - r_{ij} )} } } \right\rangle.
1117   \]
1118 + Note that the delta function can be replaced by a histogram in
1119 + computer simulation. Figure
1120 + \ref{introFigure:pairDistributionFunction} shows a typical pair
1121 + distribution function for the liquid argon system. The occurrence of
1122 + several peaks in the plot of $g(r)$ suggests that it is more likely
1123 + to find particles at certain radial values than at others. This is a
1124 + result of the attractive interaction at such distances. Because of
1125 + the strong repulsive forces at short distance, the probability of
1126 + locating particles at distances less than about 2.5{\AA} from each
1127 + other is essentially zero.
1128  
1129 < Some relatively important transformation,
1130 < \[
1131 < L(\cos at) = \frac{p}{{p^2  + a^2 }}
1129 > %\begin{figure}
1130 > %\centering
1131 > %\includegraphics[width=\linewidth]{pdf.eps}
1132 > %\caption[Pair distribution function for the liquid argon
1133 > %]{Pair distribution function for the liquid argon}
1134 > %\label{introFigure:pairDistributionFunction}
1135 > %\end{figure}
1136 >
1137 > \subsubsection{\label{introSection:timeDependentProperties}Time-dependent
1138 > Properties}
1139 >
1140 > Time-dependent properties are usually calculated using \emph{time
1141 > correlation function}, which correlates random variables $A$ and $B$
1142 > at two different time
1143 > \begin{equation}
1144 > C_{AB} (t) = \left\langle {A(t)B(0)} \right\rangle.
1145 > \label{introEquation:timeCorrelationFunction}
1146 > \end{equation}
1147 > If $A$ and $B$ refer to same variable, this kind of correlation
1148 > function is called \emph{auto correlation function}. One example of
1149 > auto correlation function is velocity auto-correlation function
1150 > which is directly related to transport properties of molecular
1151 > liquids:
1152 > \[
1153 > D = \frac{1}{3}\int\limits_0^\infty  {\left\langle {v(t) \cdot v(0)}
1154 > \right\rangle } dt
1155   \]
1156 + where $D$ is diffusion constant. Unlike velocity autocorrelation
1157 + function which is averaging over time origins and over all the
1158 + atoms, dipole autocorrelation are calculated for the entire system.
1159 + The dipole autocorrelation function is given by:
1160 + \[
1161 + c_{dipole}  = \left\langle {u_{tot} (t) \cdot u_{tot} (t)}
1162 + \right\rangle
1163 + \]
1164 + Here $u_{tot}$ is the net dipole of the entire system and is given
1165 + by
1166 + \[
1167 + u_{tot} (t) = \sum\limits_i {u_i (t)}
1168 + \]
1169 + In principle, many time correlation functions can be related with
1170 + Fourier transforms of the infrared, Raman, and inelastic neutron
1171 + scattering spectra of molecular liquids. In practice, one can
1172 + extract the IR spectrum from the intensity of dipole fluctuation at
1173 + each frequency using the following relationship:
1174 + \[
1175 + \hat c_{dipole} (v) = \int_{ - \infty }^\infty  {c_{dipole} (t)e^{ -
1176 + i2\pi vt} dt}
1177 + \]
1178  
1179 + \section{\label{introSection:rigidBody}Dynamics of Rigid Bodies}
1180 +
1181 + Rigid bodies are frequently involved in the modeling of different
1182 + areas, from engineering, physics, to chemistry. For example,
1183 + missiles and vehicle are usually modeled by rigid bodies.  The
1184 + movement of the objects in 3D gaming engine or other physics
1185 + simulator is governed by the rigid body dynamics. In molecular
1186 + simulation, rigid body is used to simplify the model in
1187 + protein-protein docking study{\cite{Gray03}}.
1188 +
1189 + It is very important to develop stable and efficient methods to
1190 + integrate the equations of motion of orientational degrees of
1191 + freedom. Euler angles are the nature choice to describe the
1192 + rotational degrees of freedom. However, due to its singularity, the
1193 + numerical integration of corresponding equations of motion is very
1194 + inefficient and inaccurate. Although an alternative integrator using
1195 + different sets of Euler angles can overcome this difficulty\cite{},
1196 + the computational penalty and the lost of angular momentum
1197 + conservation still remain. A singularity free representation
1198 + utilizing quaternions was developed by Evans in 1977. Unfortunately,
1199 + this approach suffer from the nonseparable Hamiltonian resulted from
1200 + quaternion representation, which prevents the symplectic algorithm
1201 + to be utilized. Another different approach is to apply holonomic
1202 + constraints to the atoms belonging to the rigid body. Each atom
1203 + moves independently under the normal forces deriving from potential
1204 + energy and constraint forces which are used to guarantee the
1205 + rigidness. However, due to their iterative nature, SHAKE and Rattle
1206 + algorithm converge very slowly when the number of constraint
1207 + increases.
1208 +
1209 + The break through in geometric literature suggests that, in order to
1210 + develop a long-term integration scheme, one should preserve the
1211 + symplectic structure of the flow. Introducing conjugate momentum to
1212 + rotation matrix $Q$ and re-formulating Hamiltonian's equation, a
1213 + symplectic integrator, RSHAKE, was proposed to evolve the
1214 + Hamiltonian system in a constraint manifold by iteratively
1215 + satisfying the orthogonality constraint $Q_T Q = 1$. An alternative
1216 + method using quaternion representation was developed by Omelyan.
1217 + However, both of these methods are iterative and inefficient. In
1218 + this section, we will present a symplectic Lie-Poisson integrator
1219 + for rigid body developed by Dullweber and his
1220 + coworkers\cite{Dullweber1997} in depth.
1221 +
1222 + \subsection{\label{introSection:constrainedHamiltonianRB}Constrained Hamiltonian for Rigid Body}
1223 + The motion of the rigid body is Hamiltonian with the Hamiltonian
1224 + function
1225 + \begin{equation}
1226 + H = \frac{1}{2}(p^T m^{ - 1} p) + \frac{1}{2}tr(PJ^{ - 1} P) +
1227 + V(q,Q) + \frac{1}{2}tr[(QQ^T  - 1)\Lambda ].
1228 + \label{introEquation:RBHamiltonian}
1229 + \end{equation}
1230 + Here, $q$ and $Q$  are the position and rotation matrix for the
1231 + rigid-body, $p$ and $P$  are conjugate momenta to $q$  and $Q$ , and
1232 + $J$, a diagonal matrix, is defined by
1233   \[
1234 < L(\sin at) = \frac{a}{{p^2  + a^2 }}
1234 > I_{ii}^{ - 1}  = \frac{1}{2}\sum\limits_{i \ne j} {J_{jj}^{ - 1} }
1235   \]
1236 + where $I_{ii}$ is the diagonal element of the inertia tensor. This
1237 + constrained Hamiltonian equation subjects to a holonomic constraint,
1238 + \begin{equation}
1239 + Q^T Q = 1, \label{introEquation:orthogonalConstraint}
1240 + \end{equation}
1241 + which is used to ensure rotation matrix's orthogonality.
1242 + Differentiating \ref{introEquation:orthogonalConstraint} and using
1243 + Equation \ref{introEquation:RBMotionMomentum}, one may obtain,
1244 + \begin{equation}
1245 + Q^T PJ^{ - 1}  + J^{ - 1} P^T Q = 0 . \\
1246 + \label{introEquation:RBFirstOrderConstraint}
1247 + \end{equation}
1248  
1249 + Using Equation (\ref{introEquation:motionHamiltonianCoordinate},
1250 + \ref{introEquation:motionHamiltonianMomentum}), one can write down
1251 + the equations of motion,
1252   \[
1253 < L(1) = \frac{1}{p}
1253 > \begin{array}{c}
1254 > \frac{{dq}}{{dt}} = \frac{p}{m} \label{introEquation:RBMotionPosition}\\
1255 > \frac{{dp}}{{dt}} =  - \nabla _q V(q,Q) \label{introEquation:RBMotionMomentum}\\
1256 > \frac{{dQ}}{{dt}} = PJ^{ - 1}  \label{introEquation:RBMotionRotation}\\
1257 > \frac{{dP}}{{dt}} =  - \nabla _Q V(q,Q) - 2Q\Lambda . \label{introEquation:RBMotionP}\\
1258 > \end{array}
1259   \]
1260  
1261 < First, the bath coordinates,
1261 > In general, there are two ways to satisfy the holonomic constraints.
1262 > We can use constraint force provided by lagrange multiplier on the
1263 > normal manifold to keep the motion on constraint space. Or we can
1264 > simply evolve the system in constraint manifold. These two methods
1265 > are proved to be equivalent. The holonomic constraint and equations
1266 > of motions define a constraint manifold for rigid body
1267   \[
1268 < p^2 L(x_\alpha  ) - px_\alpha  (0) - \dot x_\alpha  (0) =  - \omega
1269 < _\alpha ^2 L(x_\alpha  ) + \frac{{g_\alpha  }}{{\omega _\alpha
466 < }}L(x)
1268 > M = \left\{ {(Q,P):Q^T Q = 1,Q^T PJ^{ - 1}  + J^{ - 1} P^T Q = 0}
1269 > \right\}.
1270   \]
1271 +
1272 + Unfortunately, this constraint manifold is not the cotangent bundle
1273 + $T_{\star}SO(3)$. However, it turns out that under symplectic
1274 + transformation, the cotangent space and the phase space are
1275 + diffeomorphic. Introducing
1276   \[
1277 < L(x_\alpha  ) = \frac{{\frac{{g_\alpha  }}{{\omega _\alpha  }}L(x) +
470 < px_\alpha  (0) + \dot x_\alpha  (0)}}{{p^2  + \omega _\alpha ^2 }}
1277 > \tilde Q = Q,\tilde P = \frac{1}{2}\left( {P - QP^T Q} \right),
1278   \]
1279 < Then, the system coordinates,
1280 < \begin{align}
1281 < mL(\ddot x) &=  - \frac{1}{p}\frac{{\partial W(x)}}{{\partial x}} -
1282 < \sum\limits_{\alpha  = 1}^N {\left\{ {\frac{{\frac{{g_\alpha
1283 < }}{{\omega _\alpha  }}L(x) + px_\alpha  (0) + \dot x_\alpha
1284 < (0)}}{{p^2  + \omega _\alpha ^2 }} - \frac{{g_\alpha ^2 }}{{m_\alpha
478 < }}\omega _\alpha ^2 L(x)} \right\}}
479 < %
480 < &= - \frac{1}{p}\frac{{\partial W(x)}}{{\partial x}} -
481 < \sum\limits_{\alpha  = 1}^N {\left\{ { - \frac{{g_\alpha ^2 }}{{m_\alpha  \omega _\alpha ^2 }}\frac{p}{{p^2  + \omega _\alpha ^2 }}pL(x)
482 < - \frac{p}{{p^2  + \omega _\alpha ^2 }}g_\alpha  x_\alpha  (0)
483 < - \frac{1}{{p^2  + \omega _\alpha ^2 }}g_\alpha  \dot x_\alpha  (0)} \right\}}
484 < \end{align}
485 < Then, the inverse transform,
1279 > the mechanical system subject to a holonomic constraint manifold $M$
1280 > can be re-formulated as a Hamiltonian system on the cotangent space
1281 > \[
1282 > T^* SO(3) = \left\{ {(\tilde Q,\tilde P):\tilde Q^T \tilde Q =
1283 > 1,\tilde Q^T \tilde PJ^{ - 1}  + J^{ - 1} P^T \tilde Q = 0} \right\}
1284 > \]
1285  
1286 + For a body fixed vector $X_i$ with respect to the center of mass of
1287 + the rigid body, its corresponding lab fixed vector $X_0^{lab}$  is
1288 + given as
1289 + \begin{equation}
1290 + X_i^{lab} = Q X_i + q.
1291 + \end{equation}
1292 + Therefore, potential energy $V(q,Q)$ is defined by
1293 + \[
1294 + V(q,Q) = V(Q X_0 + q).
1295 + \]
1296 + Hence, the force and torque are given by
1297 + \[
1298 + \nabla _q V(q,Q) = F(q,Q) = \sum\limits_i {F_i (q,Q)},
1299 + \]
1300 + and
1301 + \[
1302 + \nabla _Q V(q,Q) = F(q,Q)X_i^t
1303 + \]
1304 + respectively.
1305 +
1306 + As a common choice to describe the rotation dynamics of the rigid
1307 + body, angular momentum on body frame $\Pi  = Q^t P$ is introduced to
1308 + rewrite the equations of motion,
1309 + \begin{equation}
1310 + \begin{array}{l}
1311 + \mathop \Pi \limits^ \bullet   = J^{ - 1} \Pi ^T \Pi  + Q^T \sum\limits_i {F_i (q,Q)X_i^T }  - \Lambda  \\
1312 + \mathop Q\limits^{{\rm{   }} \bullet }  = Q\Pi {\rm{ }}J^{ - 1}  \\
1313 + \end{array}
1314 + \label{introEqaution:RBMotionPI}
1315 + \end{equation}
1316 + , as well as holonomic constraints,
1317 + \[
1318 + \begin{array}{l}
1319 + \Pi J^{ - 1}  + J^{ - 1} \Pi ^t  = 0 \\
1320 + Q^T Q = 1 \\
1321 + \end{array}
1322 + \]
1323 +
1324 + For a vector $v(v_1 ,v_2 ,v_3 ) \in R^3$ and a matrix $\hat v \in
1325 + so(3)^ \star$, the hat-map isomorphism,
1326 + \begin{equation}
1327 + v(v_1 ,v_2 ,v_3 ) \Leftrightarrow \hat v = \left(
1328 + {\begin{array}{*{20}c}
1329 +   0 & { - v_3 } & {v_2 }  \\
1330 +   {v_3 } & 0 & { - v_1 }  \\
1331 +   { - v_2 } & {v_1 } & 0  \\
1332 + \end{array}} \right),
1333 + \label{introEquation:hatmapIsomorphism}
1334 + \end{equation}
1335 + will let us associate the matrix products with traditional vector
1336 + operations
1337 + \[
1338 + \hat vu = v \times u
1339 + \]
1340 +
1341 + Using \ref{introEqaution:RBMotionPI}, one can construct a skew
1342 + matrix,
1343 + \begin{equation}
1344 + (\mathop \Pi \limits^ \bullet   - \mathop \Pi \limits^ \bullet  ^T
1345 + ){\rm{ }} = {\rm{ }}(\Pi  - \Pi ^T ){\rm{ }}(J^{ - 1} \Pi  + \Pi J^{
1346 + - 1} ) + \sum\limits_i {[Q^T F_i (r,Q)X_i^T  - X_i F_i (r,Q)^T Q]} -
1347 + (\Lambda  - \Lambda ^T ) . \label{introEquation:skewMatrixPI}
1348 + \end{equation}
1349 + Since $\Lambda$ is symmetric, the last term of Equation
1350 + \ref{introEquation:skewMatrixPI} is zero, which implies the Lagrange
1351 + multiplier $\Lambda$ is absent from the equations of motion. This
1352 + unique property eliminate the requirement of iterations which can
1353 + not be avoided in other methods\cite{}.
1354 +
1355 + Applying hat-map isomorphism, we obtain the equation of motion for
1356 + angular momentum on body frame
1357 + \begin{equation}
1358 + \dot \pi  = \pi  \times I^{ - 1} \pi  + \sum\limits_i {\left( {Q^T
1359 + F_i (r,Q)} \right) \times X_i }.
1360 + \label{introEquation:bodyAngularMotion}
1361 + \end{equation}
1362 + In the same manner, the equation of motion for rotation matrix is
1363 + given by
1364 + \[
1365 + \dot Q = Qskew(I^{ - 1} \pi )
1366 + \]
1367 +
1368 + \subsection{\label{introSection:SymplecticFreeRB}Symplectic
1369 + Lie-Poisson Integrator for Free Rigid Body}
1370 +
1371 + If there is not external forces exerted on the rigid body, the only
1372 + contribution to the rotational is from the kinetic potential (the
1373 + first term of \ref{ introEquation:bodyAngularMotion}). The free
1374 + rigid body is an example of Lie-Poisson system with Hamiltonian
1375 + function
1376 + \begin{equation}
1377 + T^r (\pi ) = T_1 ^r (\pi _1 ) + T_2^r (\pi _2 ) + T_3^r (\pi _3 )
1378 + \label{introEquation:rotationalKineticRB}
1379 + \end{equation}
1380 + where $T_i^r (\pi _i ) = \frac{{\pi _i ^2 }}{{2I_i }}$ and
1381 + Lie-Poisson structure matrix,
1382 + \begin{equation}
1383 + J(\pi ) = \left( {\begin{array}{*{20}c}
1384 +   0 & {\pi _3 } & { - \pi _2 }  \\
1385 +   { - \pi _3 } & 0 & {\pi _1 }  \\
1386 +   {\pi _2 } & { - \pi _1 } & 0  \\
1387 + \end{array}} \right)
1388 + \end{equation}
1389 + Thus, the dynamics of free rigid body is governed by
1390 + \begin{equation}
1391 + \frac{d}{{dt}}\pi  = J(\pi )\nabla _\pi  T^r (\pi )
1392 + \end{equation}
1393 +
1394 + One may notice that each $T_i^r$ in Equation
1395 + \ref{introEquation:rotationalKineticRB} can be solved exactly. For
1396 + instance, the equations of motion due to $T_1^r$ are given by
1397 + \begin{equation}
1398 + \frac{d}{{dt}}\pi  = R_1 \pi ,\frac{d}{{dt}}Q = QR_1
1399 + \label{introEqaution:RBMotionSingleTerm}
1400 + \end{equation}
1401 + where
1402 + \[ R_1  = \left( {\begin{array}{*{20}c}
1403 +   0 & 0 & 0  \\
1404 +   0 & 0 & {\pi _1 }  \\
1405 +   0 & { - \pi _1 } & 0  \\
1406 + \end{array}} \right).
1407 + \]
1408 + The solutions of Equation \ref{introEqaution:RBMotionSingleTerm} is
1409 + \[
1410 + \pi (\Delta t) = e^{\Delta tR_1 } \pi (0),Q(\Delta t) =
1411 + Q(0)e^{\Delta tR_1 }
1412 + \]
1413 + with
1414 + \[
1415 + e^{\Delta tR_1 }  = \left( {\begin{array}{*{20}c}
1416 +   0 & 0 & 0  \\
1417 +   0 & {\cos \theta _1 } & {\sin \theta _1 }  \\
1418 +   0 & { - \sin \theta _1 } & {\cos \theta _1 }  \\
1419 + \end{array}} \right),\theta _1  = \frac{{\pi _1 }}{{I_1 }}\Delta t.
1420 + \]
1421 + To reduce the cost of computing expensive functions in $e^{\Delta
1422 + tR_1 }$, we can use Cayley transformation,
1423 + \[
1424 + e^{\Delta tR_1 }  \approx (1 - \Delta tR_1 )^{ - 1} (1 + \Delta tR_1
1425 + )
1426 + \]
1427 + The flow maps for $T_2^r$ and $T_3^r$ can be found in the same
1428 + manner.
1429 +
1430 + In order to construct a second-order symplectic method, we split the
1431 + angular kinetic Hamiltonian function can into five terms
1432 + \[
1433 + T^r (\pi ) = \frac{1}{2}T_1 ^r (\pi _1 ) + \frac{1}{2}T_2^r (\pi _2
1434 + ) + T_3^r (\pi _3 ) + \frac{1}{2}T_2^r (\pi _2 ) + \frac{1}{2}T_1 ^r
1435 + (\pi _1 )
1436 + \].
1437 + Concatenating flows corresponding to these five terms, we can obtain
1438 + an symplectic integrator,
1439 + \[
1440 + \varphi _{\Delta t,T^r }  = \varphi _{\Delta t/2,\pi _1 }  \circ
1441 + \varphi _{\Delta t/2,\pi _2 }  \circ \varphi _{\Delta t,\pi _3 }
1442 + \circ \varphi _{\Delta t/2,\pi _2 }  \circ \varphi _{\Delta t/2,\pi
1443 + _1 }.
1444 + \]
1445 +
1446 + The non-canonical Lie-Poisson bracket ${F, G}$ of two function
1447 + $F(\pi )$ and $G(\pi )$ is defined by
1448 + \[
1449 + \{ F,G\} (\pi ) = [\nabla _\pi  F(\pi )]^T J(\pi )\nabla _\pi  G(\pi
1450 + )
1451 + \]
1452 + If the Poisson bracket of a function $F$ with an arbitrary smooth
1453 + function $G$ is zero, $F$ is a \emph{Casimir}, which is the
1454 + conserved quantity in Poisson system. We can easily verify that the
1455 + norm of the angular momentum, $\parallel \pi
1456 + \parallel$, is a \emph{Casimir}. Let$ F(\pi ) = S(\frac{{\parallel
1457 + \pi \parallel ^2 }}{2})$ for an arbitrary function $ S:R \to R$ ,
1458 + then by the chain rule
1459 + \[
1460 + \nabla _\pi  F(\pi ) = S'(\frac{{\parallel \pi \parallel ^2
1461 + }}{2})\pi
1462 + \]
1463 + Thus $ [\nabla _\pi  F(\pi )]^T J(\pi ) =  - S'(\frac{{\parallel \pi
1464 + \parallel ^2 }}{2})\pi  \times \pi  = 0 $. This explicit
1465 + Lie-Poisson integrator is found to be extremely efficient and stable
1466 + which can be explained by the fact the small angle approximation is
1467 + used and the norm of the angular momentum is conserved.
1468 +
1469 + \subsection{\label{introSection:RBHamiltonianSplitting} Hamiltonian
1470 + Splitting for Rigid Body}
1471 +
1472 + The Hamiltonian of rigid body can be separated in terms of kinetic
1473 + energy and potential energy,
1474 + \[
1475 + H = T(p,\pi ) + V(q,Q)
1476 + \]
1477 + The equations of motion corresponding to potential energy and
1478 + kinetic energy are listed in the below table,
1479 + \begin{table}
1480 + \caption{Equations of motion due to Potential and Kinetic Energies}
1481 + \begin{center}
1482 + \begin{tabular}{|l|l|}
1483 +  \hline
1484 +  % after \\: \hline or \cline{col1-col2} \cline{col3-col4} ...
1485 +  Potential & Kinetic \\
1486 +  $\frac{{dq}}{{dt}} = \frac{p}{m}$ & $\frac{d}{{dt}}q = p$ \\
1487 +  $\frac{d}{{dt}}p =  - \frac{{\partial V}}{{\partial q}}$ & $ \frac{d}{{dt}}p = 0$ \\
1488 +  $\frac{d}{{dt}}Q = 0$ & $ \frac{d}{{dt}}Q = Qskew(I^{ - 1} j)$ \\
1489 +  $ \frac{d}{{dt}}\pi  = \sum\limits_i {\left( {Q^T F_i (r,Q)} \right) \times X_i }$ & $\frac{d}{{dt}}\pi  = \pi  \times I^{ - 1} \pi$\\
1490 +  \hline
1491 + \end{tabular}
1492 + \end{center}
1493 + \end{table}
1494 + A second-order symplectic method is now obtained by the
1495 + composition of the flow maps,
1496 + \[
1497 + \varphi _{\Delta t}  = \varphi _{\Delta t/2,V}  \circ \varphi
1498 + _{\Delta t,T}  \circ \varphi _{\Delta t/2,V}.
1499 + \]
1500 + Moreover, $\varphi _{\Delta t/2,V}$ can be divided into two
1501 + sub-flows which corresponding to force and torque respectively,
1502 + \[
1503 + \varphi _{\Delta t/2,V}  = \varphi _{\Delta t/2,F}  \circ \varphi
1504 + _{\Delta t/2,\tau }.
1505 + \]
1506 + Since the associated operators of $\varphi _{\Delta t/2,F} $ and
1507 + $\circ \varphi _{\Delta t/2,\tau }$ are commuted, the composition
1508 + order inside $\varphi _{\Delta t/2,V}$ does not matter.
1509 +
1510 + Furthermore, kinetic potential can be separated to translational
1511 + kinetic term, $T^t (p)$, and rotational kinetic term, $T^r (\pi )$,
1512 + \begin{equation}
1513 + T(p,\pi ) =T^t (p) + T^r (\pi ).
1514 + \end{equation}
1515 + where $ T^t (p) = \frac{1}{2}p^T m^{ - 1} p $ and $T^r (\pi )$ is
1516 + defined by \ref{introEquation:rotationalKineticRB}. Therefore, the
1517 + corresponding flow maps are given by
1518 + \[
1519 + \varphi _{\Delta t,T}  = \varphi _{\Delta t,T^t }  \circ \varphi
1520 + _{\Delta t,T^r }.
1521 + \]
1522 + Finally, we obtain the overall symplectic flow maps for free moving
1523 + rigid body
1524 + \begin{equation}
1525 + \begin{array}{c}
1526 + \varphi _{\Delta t}  = \varphi _{\Delta t/2,F}  \circ \varphi _{\Delta t/2,\tau }  \\
1527 +  \circ \varphi _{\Delta t,T^t }  \circ \varphi _{\Delta t/2,\pi _1 }  \circ \varphi _{\Delta t/2,\pi _2 }  \circ \varphi _{\Delta t,\pi _3 }  \circ \varphi _{\Delta t/2,\pi _2 }  \circ \varphi _{\Delta t/2,\pi _1 }  \\
1528 +  \circ \varphi _{\Delta t/2,\tau }  \circ \varphi _{\Delta t/2,F}  .\\
1529 + \end{array}
1530 + \label{introEquation:overallRBFlowMaps}
1531 + \end{equation}
1532 +
1533 + \section{\label{introSection:langevinDynamics}Langevin Dynamics}
1534 + As an alternative to newtonian dynamics, Langevin dynamics, which
1535 + mimics a simple heat bath with stochastic and dissipative forces,
1536 + has been applied in a variety of studies. This section will review
1537 + the theory of Langevin dynamics simulation. A brief derivation of
1538 + generalized Langevin equation will be given first. Follow that, we
1539 + will discuss the physical meaning of the terms appearing in the
1540 + equation as well as the calculation of friction tensor from
1541 + hydrodynamics theory.
1542 +
1543 + \subsection{\label{introSection:generalizedLangevinDynamics}Derivation of Generalized Langevin Equation}
1544 +
1545 + Harmonic bath model, in which an effective set of harmonic
1546 + oscillators are used to mimic the effect of a linearly responding
1547 + environment, has been widely used in quantum chemistry and
1548 + statistical mechanics. One of the successful applications of
1549 + Harmonic bath model is the derivation of Deriving Generalized
1550 + Langevin Dynamics. Lets consider a system, in which the degree of
1551 + freedom $x$ is assumed to couple to the bath linearly, giving a
1552 + Hamiltonian of the form
1553 + \begin{equation}
1554 + H = \frac{{p^2 }}{{2m}} + U(x) + H_B  + \Delta U(x,x_1 , \ldots x_N)
1555 + \label{introEquation:bathGLE}.
1556 + \end{equation}
1557 + Here $p$ is a momentum conjugate to $q$, $m$ is the mass associated
1558 + with this degree of freedom, $H_B$ is harmonic bath Hamiltonian,
1559 + \[
1560 + H_B  = \sum\limits_{\alpha  = 1}^N {\left\{ {\frac{{p_\alpha ^2
1561 + }}{{2m_\alpha  }} + \frac{1}{2}m_\alpha  \omega _\alpha ^2 }
1562 + \right\}}
1563 + \]
1564 + where the index $\alpha$ runs over all the bath degrees of freedom,
1565 + $\omega _\alpha$ are the harmonic bath frequencies, $m_\alpha$ are
1566 + the harmonic bath masses, and $\Delta U$ is bilinear system-bath
1567 + coupling,
1568 + \[
1569 + \Delta U =  - \sum\limits_{\alpha  = 1}^N {g_\alpha  x_\alpha  x}
1570 + \]
1571 + where $g_\alpha$ are the coupling constants between the bath and the
1572 + coordinate $x$. Introducing
1573 + \[
1574 + W(x) = U(x) - \sum\limits_{\alpha  = 1}^N {\frac{{g_\alpha ^2
1575 + }}{{2m_\alpha  w_\alpha ^2 }}} x^2
1576 + \] and combining the last two terms in Equation
1577 + \ref{introEquation:bathGLE}, we may rewrite the Harmonic bath
1578 + Hamiltonian as
1579 + \[
1580 + H = \frac{{p^2 }}{{2m}} + W(x) + \sum\limits_{\alpha  = 1}^N
1581 + {\left\{ {\frac{{p_\alpha ^2 }}{{2m_\alpha  }} + \frac{1}{2}m_\alpha
1582 + w_\alpha ^2 \left( {x_\alpha   - \frac{{g_\alpha  }}{{m_\alpha
1583 + w_\alpha ^2 }}x} \right)^2 } \right\}}
1584 + \]
1585 + Since the first two terms of the new Hamiltonian depend only on the
1586 + system coordinates, we can get the equations of motion for
1587 + Generalized Langevin Dynamics by Hamilton's equations
1588 + \ref{introEquation:motionHamiltonianCoordinate,
1589 + introEquation:motionHamiltonianMomentum},
1590 + \begin{equation}
1591 + m\ddot x =  - \frac{{\partial W(x)}}{{\partial x}} -
1592 + \sum\limits_{\alpha  = 1}^N {g_\alpha  \left( {x_\alpha   -
1593 + \frac{{g_\alpha  }}{{m_\alpha  w_\alpha ^2 }}x} \right)},
1594 + \label{introEquation:coorMotionGLE}
1595 + \end{equation}
1596 + and
1597 + \begin{equation}
1598 + m\ddot x_\alpha   =  - m_\alpha  w_\alpha ^2 \left( {x_\alpha   -
1599 + \frac{{g_\alpha  }}{{m_\alpha  w_\alpha ^2 }}x} \right).
1600 + \label{introEquation:bathMotionGLE}
1601 + \end{equation}
1602 +
1603 + In order to derive an equation for $x$, the dynamics of the bath
1604 + variables $x_\alpha$ must be solved exactly first. As an integral
1605 + transform which is particularly useful in solving linear ordinary
1606 + differential equations, Laplace transform is the appropriate tool to
1607 + solve this problem. The basic idea is to transform the difficult
1608 + differential equations into simple algebra problems which can be
1609 + solved easily. Then applying inverse Laplace transform, also known
1610 + as the Bromwich integral, we can retrieve the solutions of the
1611 + original problems.
1612 +
1613 + Let $f(t)$ be a function defined on $ [0,\infty ) $. The Laplace
1614 + transform of f(t) is a new function defined as
1615 + \[
1616 + L(f(t)) \equiv F(p) = \int_0^\infty  {f(t)e^{ - pt} dt}
1617 + \]
1618 + where  $p$ is real and  $L$ is called the Laplace Transform
1619 + Operator. Below are some important properties of Laplace transform
1620 + \begin{equation}
1621 + \begin{array}{c}
1622 + L(x + y) = L(x) + L(y) \\
1623 + L(ax) = aL(x) \\
1624 + L(\dot x) = pL(x) - px(0) \\
1625 + L(\ddot x) = p^2 L(x) - px(0) - \dot x(0) \\
1626 + L\left( {\int_0^t {g(t - \tau )h(\tau )d\tau } } \right) = G(p)H(p) \\
1627 + \end{array}
1628 + \end{equation}
1629 +
1630 + Applying Laplace transform to the bath coordinates, we obtain
1631 + \[
1632 + \begin{array}{c}
1633 + p^2 L(x_\alpha  ) - px_\alpha  (0) - \dot x_\alpha  (0) =  - \omega _\alpha ^2 L(x_\alpha  ) + \frac{{g_\alpha  }}{{\omega _\alpha  }}L(x) \\
1634 + L(x_\alpha  ) = \frac{{\frac{{g_\alpha  }}{{\omega _\alpha  }}L(x) + px_\alpha  (0) + \dot x_\alpha  (0)}}{{p^2  + \omega _\alpha ^2 }} \\
1635 + \end{array}
1636 + \]
1637 + By the same way, the system coordinates become
1638 + \[
1639 + \begin{array}{c}
1640 + mL(\ddot x) =  - \frac{1}{p}\frac{{\partial W(x)}}{{\partial x}} \\
1641 +  - \sum\limits_{\alpha  = 1}^N {\left\{ { - \frac{{g_\alpha ^2 }}{{m_\alpha  \omega _\alpha ^2 }}\frac{p}{{p^2  + \omega _\alpha ^2 }}pL(x) - \frac{p}{{p^2  + \omega _\alpha ^2 }}g_\alpha  x_\alpha  (0) - \frac{1}{{p^2  + \omega _\alpha ^2 }}g_\alpha  \dot x_\alpha  (0)} \right\}}  \\
1642 + \end{array}
1643 + \]
1644 +
1645 + With the help of some relatively important inverse Laplace
1646 + transformations:
1647 + \[
1648 + \begin{array}{c}
1649 + L(\cos at) = \frac{p}{{p^2  + a^2 }} \\
1650 + L(\sin at) = \frac{a}{{p^2  + a^2 }} \\
1651 + L(1) = \frac{1}{p} \\
1652 + \end{array}
1653 + \]
1654 + , we obtain
1655   \begin{align}
1656   m\ddot x &=  - \frac{{\partial W(x)}}{{\partial x}} -
1657   \sum\limits_{\alpha  = 1}^N {\left\{ {\left( { - \frac{{g_\alpha ^2
# Line 503 | Line 1671 | t)\dot x(t - \tau )d} \tau }  + \sum\limits_{\alpha  =
1671   (\omega _\alpha  t)} \right\}}
1672   \end{align}
1673  
1674 + Introducing a \emph{dynamic friction kernel}
1675   \begin{equation}
1676 + \xi (t) = \sum\limits_{\alpha  = 1}^N {\left( { - \frac{{g_\alpha ^2
1677 + }}{{m_\alpha  \omega _\alpha ^2 }}} \right)\cos (\omega _\alpha  t)}
1678 + \label{introEquation:dynamicFrictionKernelDefinition}
1679 + \end{equation}
1680 + and \emph{a random force}
1681 + \begin{equation}
1682 + R(t) = \sum\limits_{\alpha  = 1}^N {\left( {g_\alpha  x_\alpha  (0)
1683 + - \frac{{g_\alpha ^2 }}{{m_\alpha  \omega _\alpha ^2 }}x(0)}
1684 + \right)\cos (\omega _\alpha  t)}  + \frac{{\dot x_\alpha
1685 + (0)}}{{\omega _\alpha  }}\sin (\omega _\alpha  t),
1686 + \label{introEquation:randomForceDefinition}
1687 + \end{equation}
1688 + the equation of motion can be rewritten as
1689 + \begin{equation}
1690   m\ddot x =  - \frac{{\partial W}}{{\partial x}} - \int_0^t {\xi
1691   (t)\dot x(t - \tau )d\tau }  + R(t)
1692   \label{introEuqation:GeneralizedLangevinDynamics}
1693   \end{equation}
1694 < %where $ {\xi (t)}$ is friction kernel, $R(t)$ is random force and
1695 < %$W$ is the potential of mean force. $W(x) =  - kT\ln p(x)$
1694 > which is known as the \emph{generalized Langevin equation}.
1695 >
1696 > \subsubsection{\label{introSection:randomForceDynamicFrictionKernel}Random Force and Dynamic Friction Kernel}
1697 >
1698 > One may notice that $R(t)$ depends only on initial conditions, which
1699 > implies it is completely deterministic within the context of a
1700 > harmonic bath. However, it is easy to verify that $R(t)$ is totally
1701 > uncorrelated to $x$ and $\dot x$,
1702   \[
1703 < \xi (t) = \sum\limits_{\alpha  = 1}^N {\left( { - \frac{{g_\alpha ^2
1704 < }}{{m_\alpha  \omega _\alpha ^2 }}} \right)\cos (\omega _\alpha  t)}
1703 > \begin{array}{l}
1704 > \left\langle {x(t)R(t)} \right\rangle  = 0, \\
1705 > \left\langle {\dot x(t)R(t)} \right\rangle  = 0. \\
1706 > \end{array}
1707   \]
1708 < For an infinite harmonic bath, we can use the spectral density and
1709 < an integral over frequencies.
1708 > This property is what we expect from a truly random process. As long
1709 > as the model, which is gaussian distribution in general, chosen for
1710 > $R(t)$ is a truly random process, the stochastic nature of the GLE
1711 > still remains.
1712  
1713 + %dynamic friction kernel
1714 + The convolution integral
1715   \[
1716 < R(t) = \sum\limits_{\alpha  = 1}^N {\left( {g_\alpha  x_\alpha  (0)
522 < - \frac{{g_\alpha ^2 }}{{m_\alpha  \omega _\alpha ^2 }}x(0)}
523 < \right)\cos (\omega _\alpha  t)}  + \frac{{\dot x_\alpha
524 < (0)}}{{\omega _\alpha  }}\sin (\omega _\alpha  t)
1716 > \int_0^t {\xi (t)\dot x(t - \tau )d\tau }
1717   \]
1718 < The random forces depend only on initial conditions.
1718 > depends on the entire history of the evolution of $x$, which implies
1719 > that the bath retains memory of previous motions. In other words,
1720 > the bath requires a finite time to respond to change in the motion
1721 > of the system. For a sluggish bath which responds slowly to changes
1722 > in the system coordinate, we may regard $\xi(t)$ as a constant
1723 > $\xi(t) = \Xi_0$. Hence, the convolution integral becomes
1724 > \[
1725 > \int_0^t {\xi (t)\dot x(t - \tau )d\tau }  = \xi _0 (x(t) - x(0))
1726 > \]
1727 > and Equation \ref{introEuqation:GeneralizedLangevinDynamics} becomes
1728 > \[
1729 > m\ddot x =  - \frac{\partial }{{\partial x}}\left( {W(x) +
1730 > \frac{1}{2}\xi _0 (x - x_0 )^2 } \right) + R(t),
1731 > \]
1732 > which can be used to describe dynamic caging effect. The other
1733 > extreme is the bath that responds infinitely quickly to motions in
1734 > the system. Thus, $\xi (t)$ can be taken as a $delta$ function in
1735 > time:
1736 > \[
1737 > \xi (t) = 2\xi _0 \delta (t)
1738 > \]
1739 > Hence, the convolution integral becomes
1740 > \[
1741 > \int_0^t {\xi (t)\dot x(t - \tau )d\tau }  = 2\xi _0 \int_0^t
1742 > {\delta (t)\dot x(t - \tau )d\tau }  = \xi _0 \dot x(t),
1743 > \]
1744 > and Equation \ref{introEuqation:GeneralizedLangevinDynamics} becomes
1745 > \begin{equation}
1746 > m\ddot x =  - \frac{{\partial W(x)}}{{\partial x}} - \xi _0 \dot
1747 > x(t) + R(t) \label{introEquation:LangevinEquation}
1748 > \end{equation}
1749 > which is known as the Langevin equation. The static friction
1750 > coefficient $\xi _0$ can either be calculated from spectral density
1751 > or be determined by Stokes' law for regular shaped particles.A
1752 > briefly review on calculating friction tensor for arbitrary shaped
1753 > particles is given in Sec.~\ref{introSection:frictionTensor}.
1754  
1755   \subsubsection{\label{introSection:secondFluctuationDissipation}The Second Fluctuation Dissipation Theorem}
1756 < So we can define a new set of coordinates,
1756 >
1757 > Defining a new set of coordinates,
1758   \[
1759   q_\alpha  (t) = x_\alpha  (t) - \frac{1}{{m_\alpha  \omega _\alpha
1760   ^2 }}x(0)
1761 < \]
1762 < This makes
1761 > \],
1762 > we can rewrite $R(T)$ as
1763   \[
1764 < R(t) = \sum\limits_{\alpha  = 1}^N {g_\alpha  q_\alpha  (t)}
1764 > R(t) = \sum\limits_{\alpha  = 1}^N {g_\alpha  q_\alpha  (t)}.
1765   \]
1766   And since the $q$ coordinates are harmonic oscillators,
1767   \[
1768 < \begin{array}{l}
1768 > \begin{array}{c}
1769 > \left\langle {q_\alpha ^2 } \right\rangle  = \frac{{kT}}{{m_\alpha  \omega _\alpha ^2 }} \\
1770   \left\langle {q_\alpha  (t)q_\alpha  (0)} \right\rangle  = \left\langle {q_\alpha ^2 (0)} \right\rangle \cos (\omega _\alpha  t) \\
1771   \left\langle {q_\alpha  (t)q_\beta  (0)} \right\rangle  = \delta _{\alpha \beta } \left\langle {q_\alpha  (t)q_\alpha  (0)} \right\rangle  \\
1772 + \left\langle {R(t)R(0)} \right\rangle  = \sum\limits_\alpha  {\sum\limits_\beta  {g_\alpha  g_\beta  \left\langle {q_\alpha  (t)q_\beta  (0)} \right\rangle } }  \\
1773 +  = \sum\limits_\alpha  {g_\alpha ^2 \left\langle {q_\alpha ^2 (0)} \right\rangle \cos (\omega _\alpha  t)}  \\
1774 +  = kT\xi (t) \\
1775   \end{array}
1776   \]
1777 <
546 < \begin{align}
547 < \left\langle {R(t)R(0)} \right\rangle  &= \sum\limits_\alpha
548 < {\sum\limits_\beta  {g_\alpha  g_\beta  \left\langle {q_\alpha
549 < (t)q_\beta  (0)} \right\rangle } }
550 < %
551 < &= \sum\limits_\alpha  {g_\alpha ^2 \left\langle {q_\alpha ^2 (0)}
552 < \right\rangle \cos (\omega _\alpha  t)}
553 < %
554 < &= kT\xi (t)
555 < \end{align}
556 <
1777 > Thus, we recover the \emph{second fluctuation dissipation theorem}
1778   \begin{equation}
1779   \xi (t) = \left\langle {R(t)R(0)} \right\rangle
1780 < \label{introEquation:secondFluctuationDissipation}
1780 > \label{introEquation:secondFluctuationDissipation}.
1781   \end{equation}
1782 + In effect, it acts as a constraint on the possible ways in which one
1783 + can model the random force and friction kernel.
1784  
562 \section{\label{introSection:hydroynamics}Hydrodynamics}
563
1785   \subsection{\label{introSection:frictionTensor} Friction Tensor}
1786 < \subsection{\label{introSection:analyticalApproach}Analytical
1787 < Approach}
1786 > Theoretically, the friction kernel can be determined using velocity
1787 > autocorrelation function. However, this approach become impractical
1788 > when the system become more and more complicate. Instead, various
1789 > approaches based on hydrodynamics have been developed to calculate
1790 > the friction coefficients. The friction effect is isotropic in
1791 > Equation, $\zeta$ can be taken as a scalar. In general, friction
1792 > tensor $\Xi$ is a $6\times 6$ matrix given by
1793 > \[
1794 > \Xi  = \left( {\begin{array}{*{20}c}
1795 >   {\Xi _{}^{tt} } & {\Xi _{}^{rt} }  \\
1796 >   {\Xi _{}^{tr} } & {\Xi _{}^{rr} }  \\
1797 > \end{array}} \right).
1798 > \]
1799 > Here, $ {\Xi^{tt} }$ and $ {\Xi^{rr} }$ are translational friction
1800 > tensor and rotational resistance (friction) tensor respectively,
1801 > while ${\Xi^{tr} }$ is translation-rotation coupling tensor and $
1802 > {\Xi^{rt} }$ is rotation-translation coupling tensor. When a
1803 > particle moves in a fluid, it may experience friction force or
1804 > torque along the opposite direction of the velocity or angular
1805 > velocity,
1806 > \[
1807 > \left( \begin{array}{l}
1808 > F_R  \\
1809 > \tau _R  \\
1810 > \end{array} \right) =  - \left( {\begin{array}{*{20}c}
1811 >   {\Xi ^{tt} } & {\Xi ^{rt} }  \\
1812 >   {\Xi ^{tr} } & {\Xi ^{rr} }  \\
1813 > \end{array}} \right)\left( \begin{array}{l}
1814 > v \\
1815 > w \\
1816 > \end{array} \right)
1817 > \]
1818 > where $F_r$ is the friction force and $\tau _R$ is the friction
1819 > toque.
1820  
1821 < \subsection{\label{introSection:approximationApproach}Approximation
569 < Approach}
1821 > \subsubsection{\label{introSection:resistanceTensorRegular}The Resistance Tensor for Regular Shape}
1822  
1823 < \subsection{\label{introSection:centersRigidBody}Centers of Rigid
1824 < Body}
1823 > For a spherical particle, the translational and rotational friction
1824 > constant can be calculated from Stoke's law,
1825 > \[
1826 > \Xi ^{tt}  = \left( {\begin{array}{*{20}c}
1827 >   {6\pi \eta R} & 0 & 0  \\
1828 >   0 & {6\pi \eta R} & 0  \\
1829 >   0 & 0 & {6\pi \eta R}  \\
1830 > \end{array}} \right)
1831 > \]
1832 > and
1833 > \[
1834 > \Xi ^{rr}  = \left( {\begin{array}{*{20}c}
1835 >   {8\pi \eta R^3 } & 0 & 0  \\
1836 >   0 & {8\pi \eta R^3 } & 0  \\
1837 >   0 & 0 & {8\pi \eta R^3 }  \\
1838 > \end{array}} \right)
1839 > \]
1840 > where $\eta$ is the viscosity of the solvent and $R$ is the
1841 > hydrodynamics radius.
1842 >
1843 > Other non-spherical shape, such as cylinder and ellipsoid
1844 > \textit{etc}, are widely used as reference for developing new
1845 > hydrodynamics theory, because their properties can be calculated
1846 > exactly. In 1936, Perrin extended Stokes's law to general ellipsoid,
1847 > also called a triaxial ellipsoid, which is given in Cartesian
1848 > coordinates by
1849 > \[
1850 > \frac{{x^2 }}{{a^2 }} + \frac{{y^2 }}{{b^2 }} + \frac{{z^2 }}{{c^2
1851 > }} = 1
1852 > \]
1853 > where the semi-axes are of lengths $a$, $b$, and $c$. Unfortunately,
1854 > due to the complexity of the elliptic integral, only the ellipsoid
1855 > with the restriction of two axes having to be equal, \textit{i.e.}
1856 > prolate($ a \ge b = c$) and oblate ($ a < b = c $), can be solved
1857 > exactly. Introducing an elliptic integral parameter $S$ for prolate,
1858 > \[
1859 > S = \frac{2}{{\sqrt {a^2  - b^2 } }}\ln \frac{{a + \sqrt {a^2  - b^2
1860 > } }}{b},
1861 > \]
1862 > and oblate,
1863 > \[
1864 > S = \frac{2}{{\sqrt {b^2  - a^2 } }}arctg\frac{{\sqrt {b^2  - a^2 }
1865 > }}{a}
1866 > \],
1867 > one can write down the translational and rotational resistance
1868 > tensors
1869 > \[
1870 > \begin{array}{l}
1871 > \Xi _a^{tt}  = 16\pi \eta \frac{{a^2  - b^2 }}{{(2a^2  - b^2 )S - 2a}} \\
1872 > \Xi _b^{tt}  = \Xi _c^{tt}  = 32\pi \eta \frac{{a^2  - b^2 }}{{(2a^2  - 3b^2 )S + 2a}} \\
1873 > \end{array},
1874 > \]
1875 > and
1876 > \[
1877 > \begin{array}{l}
1878 > \Xi _a^{rr}  = \frac{{32\pi }}{3}\eta \frac{{(a^2  - b^2 )b^2 }}{{2a - b^2 S}} \\
1879 > \Xi _b^{rr}  = \Xi _c^{rr}  = \frac{{32\pi }}{3}\eta \frac{{(a^4  - b^4 )}}{{(2a^2  - b^2 )S - 2a}} \\
1880 > \end{array}.
1881 > \]
1882 >
1883 > \subsubsection{\label{introSection:resistanceTensorRegularArbitrary}The Resistance Tensor for Arbitrary Shape}
1884 >
1885 > Unlike spherical and other regular shaped molecules, there is not
1886 > analytical solution for friction tensor of any arbitrary shaped
1887 > rigid molecules. The ellipsoid of revolution model and general
1888 > triaxial ellipsoid model have been used to approximate the
1889 > hydrodynamic properties of rigid bodies. However, since the mapping
1890 > from all possible ellipsoidal space, $r$-space, to all possible
1891 > combination of rotational diffusion coefficients, $D$-space is not
1892 > unique\cite{Wegener79} as well as the intrinsic coupling between
1893 > translational and rotational motion of rigid body\cite{}, general
1894 > ellipsoid is not always suitable for modeling arbitrarily shaped
1895 > rigid molecule. A number of studies have been devoted to determine
1896 > the friction tensor for irregularly shaped rigid bodies using more
1897 > advanced method\cite{} where the molecule of interest was modeled by
1898 > combinations of spheres(beads)\cite{} and the hydrodynamics
1899 > properties of the molecule can be calculated using the hydrodynamic
1900 > interaction tensor. Let us consider a rigid assembly of $N$ beads
1901 > immersed in a continuous medium. Due to hydrodynamics interaction,
1902 > the ``net'' velocity of $i$th bead, $v'_i$ is different than its
1903 > unperturbed velocity $v_i$,
1904 > \[
1905 > v'_i  = v_i  - \sum\limits_{j \ne i} {T_{ij} F_j }
1906 > \]
1907 > where $F_i$ is the frictional force, and $T_{ij}$ is the
1908 > hydrodynamic interaction tensor. The friction force of $i$th bead is
1909 > proportional to its ``net'' velocity
1910 > \begin{equation}
1911 > F_i  = \zeta _i v_i  - \zeta _i \sum\limits_{j \ne i} {T_{ij} F_j }.
1912 > \label{introEquation:tensorExpression}
1913 > \end{equation}
1914 > This equation is the basis for deriving the hydrodynamic tensor. In
1915 > 1930, Oseen and Burgers gave a simple solution to Equation
1916 > \ref{introEquation:tensorExpression}
1917 > \begin{equation}
1918 > T_{ij}  = \frac{1}{{8\pi \eta r_{ij} }}\left( {I + \frac{{R_{ij}
1919 > R_{ij}^T }}{{R_{ij}^2 }}} \right).
1920 > \label{introEquation:oseenTensor}
1921 > \end{equation}
1922 > Here $R_{ij}$ is the distance vector between bead $i$ and bead $j$.
1923 > A second order expression for element of different size was
1924 > introduced by Rotne and Prager\cite{} and improved by Garc\'{i}a de
1925 > la Torre and Bloomfield,
1926 > \begin{equation}
1927 > T_{ij}  = \frac{1}{{8\pi \eta R_{ij} }}\left[ {\left( {I +
1928 > \frac{{R_{ij} R_{ij}^T }}{{R_{ij}^2 }}} \right) + R\frac{{\sigma
1929 > _i^2  + \sigma _j^2 }}{{r_{ij}^2 }}\left( {\frac{I}{3} -
1930 > \frac{{R_{ij} R_{ij}^T }}{{R_{ij}^2 }}} \right)} \right].
1931 > \label{introEquation:RPTensorNonOverlapped}
1932 > \end{equation}
1933 > Both of the Equation \ref{introEquation:oseenTensor} and Equation
1934 > \ref{introEquation:RPTensorNonOverlapped} have an assumption $R_{ij}
1935 > \ge \sigma _i  + \sigma _j$. An alternative expression for
1936 > overlapping beads with the same radius, $\sigma$, is given by
1937 > \begin{equation}
1938 > T_{ij}  = \frac{1}{{6\pi \eta R_{ij} }}\left[ {\left( {1 -
1939 > \frac{2}{{32}}\frac{{R_{ij} }}{\sigma }} \right)I +
1940 > \frac{2}{{32}}\frac{{R_{ij} R_{ij}^T }}{{R_{ij} \sigma }}} \right]
1941 > \label{introEquation:RPTensorOverlapped}
1942 > \end{equation}
1943 >
1944 > To calculate the resistance tensor at an arbitrary origin $O$, we
1945 > construct a $3N \times 3N$ matrix consisting of $N \times N$
1946 > $B_{ij}$ blocks
1947 > \begin{equation}
1948 > B = \left( {\begin{array}{*{20}c}
1949 >   {B_{11} } &  \ldots  & {B_{1N} }  \\
1950 >    \vdots  &  \ddots  &  \vdots   \\
1951 >   {B_{N1} } &  \cdots  & {B_{NN} }  \\
1952 > \end{array}} \right),
1953 > \end{equation}
1954 > where $B_{ij}$ is given by
1955 > \[
1956 > B_{ij}  = \delta _{ij} \frac{I}{{6\pi \eta R}} + (1 - \delta _{ij}
1957 > )T_{ij}
1958 > \]
1959 > where $\delta _{ij}$ is Kronecker delta function. Inverting matrix
1960 > $B$, we obtain
1961 >
1962 > \[
1963 > C = B^{ - 1}  = \left( {\begin{array}{*{20}c}
1964 >   {C_{11} } &  \ldots  & {C_{1N} }  \\
1965 >    \vdots  &  \ddots  &  \vdots   \\
1966 >   {C_{N1} } &  \cdots  & {C_{NN} }  \\
1967 > \end{array}} \right)
1968 > \]
1969 > , which can be partitioned into $N \times N$ $3 \times 3$ block
1970 > $C_{ij}$. With the help of $C_{ij}$ and skew matrix $U_i$
1971 > \[
1972 > U_i  = \left( {\begin{array}{*{20}c}
1973 >   0 & { - z_i } & {y_i }  \\
1974 >   {z_i } & 0 & { - x_i }  \\
1975 >   { - y_i } & {x_i } & 0  \\
1976 > \end{array}} \right)
1977 > \]
1978 > where $x_i$, $y_i$, $z_i$ are the components of the vector joining
1979 > bead $i$ and origin $O$. Hence, the elements of resistance tensor at
1980 > arbitrary origin $O$ can be written as
1981 > \begin{equation}
1982 > \begin{array}{l}
1983 > \Xi _{}^{tt}  = \sum\limits_i {\sum\limits_j {C_{ij} } } , \\
1984 > \Xi _{}^{tr}  = \Xi _{}^{rt}  = \sum\limits_i {\sum\limits_j {U_i C_{ij} } } , \\
1985 > \Xi _{}^{rr}  =  - \sum\limits_i {\sum\limits_j {U_i C_{ij} } } U_j  \\
1986 > \end{array}
1987 > \label{introEquation:ResistanceTensorArbitraryOrigin}
1988 > \end{equation}
1989 >
1990 > The resistance tensor depends on the origin to which they refer. The
1991 > proper location for applying friction force is the center of
1992 > resistance (reaction), at which the trace of rotational resistance
1993 > tensor, $ \Xi ^{rr}$ reaches minimum. Mathematically, the center of
1994 > resistance is defined as an unique point of the rigid body at which
1995 > the translation-rotation coupling tensor are symmetric,
1996 > \begin{equation}
1997 > \Xi^{tr}  = \left( {\Xi^{tr} } \right)^T
1998 > \label{introEquation:definitionCR}
1999 > \end{equation}
2000 > Form Equation \ref{introEquation:ResistanceTensorArbitraryOrigin},
2001 > we can easily find out that the translational resistance tensor is
2002 > origin independent, while the rotational resistance tensor and
2003 > translation-rotation coupling resistance tensor depend on the
2004 > origin. Given resistance tensor at an arbitrary origin $O$, and a
2005 > vector ,$r_{OP}(x_{OP}, y_{OP}, z_{OP})$, from $O$ to $P$, we can
2006 > obtain the resistance tensor at $P$ by
2007 > \begin{equation}
2008 > \begin{array}{l}
2009 > \Xi _P^{tt}  = \Xi _O^{tt}  \\
2010 > \Xi _P^{tr}  = \Xi _P^{rt}  = \Xi _O^{tr}  - U_{OP} \Xi _O^{tt}  \\
2011 > \Xi _P^{rr}  = \Xi _O^{rr}  - U_{OP} \Xi _O^{tt} U_{OP}  + \Xi _O^{tr} U_{OP}  - U_{OP} \Xi _O^{tr} ^{^T }  \\
2012 > \end{array}
2013 > \label{introEquation:resistanceTensorTransformation}
2014 > \end{equation}
2015 > where
2016 > \[
2017 > U_{OP}  = \left( {\begin{array}{*{20}c}
2018 >   0 & { - z_{OP} } & {y_{OP} }  \\
2019 >   {z_i } & 0 & { - x_{OP} }  \\
2020 >   { - y_{OP} } & {x_{OP} } & 0  \\
2021 > \end{array}} \right)
2022 > \]
2023 > Using Equations \ref{introEquation:definitionCR} and
2024 > \ref{introEquation:resistanceTensorTransformation}, one can locate
2025 > the position of center of resistance,
2026 > \[
2027 > \left( \begin{array}{l}
2028 > x_{OR}  \\
2029 > y_{OR}  \\
2030 > z_{OR}  \\
2031 > \end{array} \right) = \left( {\begin{array}{*{20}c}
2032 >   {(\Xi _O^{rr} )_{yy}  + (\Xi _O^{rr} )_{zz} } & { - (\Xi _O^{rr} )_{xy} } & { - (\Xi _O^{rr} )_{xz} }  \\
2033 >   { - (\Xi _O^{rr} )_{xy} } & {(\Xi _O^{rr} )_{zz}  + (\Xi _O^{rr} )_{xx} } & { - (\Xi _O^{rr} )_{yz} }  \\
2034 >   { - (\Xi _O^{rr} )_{xz} } & { - (\Xi _O^{rr} )_{yz} } & {(\Xi _O^{rr} )_{xx}  + (\Xi _O^{rr} )_{yy} }  \\
2035 > \end{array}} \right)^{ - 1} \left( \begin{array}{l}
2036 > (\Xi _O^{tr} )_{yz}  - (\Xi _O^{tr} )_{zy}  \\
2037 > (\Xi _O^{tr} )_{zx}  - (\Xi _O^{tr} )_{xz}  \\
2038 > (\Xi _O^{tr} )_{xy}  - (\Xi _O^{tr} )_{yx}  \\
2039 > \end{array} \right).
2040 > \]
2041 > where $x_OR$, $y_OR$, $z_OR$ are the components of the vector
2042 > joining center of resistance $R$ and origin $O$.

Diff Legend

Removed lines
+ Added lines
< Changed lines
> Changed lines