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 2721 by tim, Thu Apr 20 03:42:21 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}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:ensemble}Ensemble and Phase Space}
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  
# 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 < \[
426 < L(x) = \int_0^\infty  {x(t)e^{ - pt} dt}
427 < \]
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   \[
845 < L(x + y) = L(x) + L(y)
845 > [X,Y] = XY - YX .
846   \]
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
851 > [X,Y]/4 + h^2 [Y,X]/4 \\ & & \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 & & \mbox{} +
853 > \ldots )
854 > \end{eqnarray*}
855 > Since \[ [X,Y] + [Y,X] = 0\] and \[ [X,X] = 0\], the dominant local
856 > error of Spring splitting is proportional to $h^3$. The same
857 > procedure can be applied to general splitting,  of the form
858 > \begin{equation}
859 > \varphi _{b_m h}^2  \circ \varphi _{a_m h}^1  \circ \varphi _{b_{m -
860 > 1} h}^2  \circ  \ldots  \circ \varphi _{a_1 h}^1 .
861 > \end{equation}
862 > Careful choice of coefficient $a_1 ,\ldot , b_m$ will lead to higher
863 > order method. Yoshida proposed an elegant way to compose higher
864 > order methods based on symmetric splitting. Given a symmetric second
865 > order base method $ \varphi _h^{(2)} $, a fourth-order symmetric
866 > method can be constructed by composing,
867   \[
868 < L(ax) = aL(x)
868 > \varphi _h^{(4)}  = \varphi _{\alpha h}^{(2)}  \circ \varphi _{\beta
869 > h}^{(2)}  \circ \varphi _{\alpha h}^{(2)}
870   \]
871 <
871 > where $ \alpha  =  - \frac{{2^{1/3} }}{{2 - 2^{1/3} }}$ and $ \beta
872 > = \frac{{2^{1/3} }}{{2 - 2^{1/3} }}$. Moreover, a symmetric
873 > integrator $ \varphi _h^{(2n + 2)}$ can be composed by
874 > \begin{equation}
875 > \varphi _h^{(2n + 2)}  = \varphi _{\alpha h}^{(2n)}  \circ \varphi
876 > _{\beta h}^{(2n)}  \circ \varphi _{\alpha h}^{(2n)}
877 > \end{equation}
878 > , if the weights are chosen as
879   \[
880 < L(\dot x) = pL(x) - px(0)
880 > \alpha  =  - \frac{{2^{1/(2n + 1)} }}{{2 - 2^{1/(2n + 1)} }},\beta =
881 > \frac{{2^{1/(2n + 1)} }}{{2 - 2^{1/(2n + 1)} }} .
882   \]
883  
884 + \section{\label{introSection:molecularDynamics}Molecular Dynamics}
885 +
886 + As one of the principal tools of molecular modeling, Molecular
887 + dynamics has proven to be a powerful tool for studying the functions
888 + of biological systems, providing structural, thermodynamic and
889 + dynamical information. The basic idea of molecular dynamics is that
890 + macroscopic properties are related to microscopic behavior and
891 + microscopic behavior can be calculated from the trajectories in
892 + simulations. For instance, instantaneous temperature of an
893 + Hamiltonian system of $N$ particle can be measured by
894   \[
895 < L(\ddot x) = p^2 L(x) - px(0) - \dot x(0)
895 > T(t) = \sum\limits_{i = 1}^N {\frac{{m_i v_i^2 }}{{fk_B }}}
896   \]
897 + where $m_i$ and $v_i$ are the mass and velocity of $i$th particle
898 + respectively, $f$ is the number of degrees of freedom, and $k_B$ is
899 + the boltzman constant.
900  
901 + A typical molecular dynamics run consists of three essential steps:
902 + \begin{enumerate}
903 +  \item Initialization
904 +    \begin{enumerate}
905 +    \item Preliminary preparation
906 +    \item Minimization
907 +    \item Heating
908 +    \item Equilibration
909 +    \end{enumerate}
910 +  \item Production
911 +  \item Analysis
912 + \end{enumerate}
913 + These three individual steps will be covered in the following
914 + sections. Sec.~\ref{introSec:initialSystemSettings} deals with the
915 + initialization of a simulation. Sec.~\ref{introSec:production} will
916 + discusses issues in production run, including the force evaluation
917 + and the numerical integration schemes of the equations of motion .
918 + Sec.~\ref{introSection:Analysis} provides the theoretical tools for
919 + trajectory analysis.
920 +
921 + \subsection{\label{introSec:initialSystemSettings}Initialization}
922 +
923 + \subsubsection{Preliminary preparation}
924 +
925 + When selecting the starting structure of a molecule for molecular
926 + simulation, one may retrieve its Cartesian coordinates from public
927 + databases, such as RCSB Protein Data Bank \textit{etc}. Although
928 + thousands of crystal structures of molecules are discovered every
929 + year, many more remain unknown due to the difficulties of
930 + purification and crystallization. Even for the molecule with known
931 + structure, some important information is missing. For example, the
932 + missing hydrogen atom which acts as donor in hydrogen bonding must
933 + be added. Moreover, in order to include electrostatic interaction,
934 + one may need to specify the partial charges for individual atoms.
935 + Under some circumstances, we may even need to prepare the system in
936 + a special setup. For instance, when studying transport phenomenon in
937 + membrane system, we may prepare the lipids in bilayer structure
938 + instead of placing lipids randomly in solvent, since we are not
939 + interested in self-aggregation and it takes a long time to happen.
940 +
941 + \subsubsection{Minimization}
942 +
943 + It is quite possible that some of molecules in the system from
944 + preliminary preparation may be overlapped with each other. This
945 + close proximity leads to high potential energy which consequently
946 + jeopardizes any molecular dynamics simulations. To remove these
947 + steric overlaps, one typically performs energy minimization to find
948 + a more reasonable conformation. Several energy minimization methods
949 + have been developed to exploit the energy surface and to locate the
950 + local minimum. While converging slowly near the minimum, steepest
951 + descent method is extremely robust when systems are far from
952 + harmonic. Thus, it is often used to refine structure from
953 + crystallographic data. Relied on the gradient or hessian, advanced
954 + methods like conjugate gradient and Newton-Raphson converge rapidly
955 + to a local minimum, while become unstable if the energy surface is
956 + far from quadratic. Another factor must be taken into account, when
957 + choosing energy minimization method, is the size of the system.
958 + Steepest descent and conjugate gradient can deal with models of any
959 + size. Because of the limit of computation power to calculate hessian
960 + matrix and insufficient storage capacity to store them, most
961 + Newton-Raphson methods can not be used with very large models.
962 +
963 + \subsubsection{Heating}
964 +
965 + Typically, Heating is performed by assigning random velocities
966 + according to a Gaussian distribution for a temperature. Beginning at
967 + a lower temperature and gradually increasing the temperature by
968 + assigning greater random velocities, we end up with setting the
969 + temperature of the system to a final temperature at which the
970 + simulation will be conducted. In heating phase, we should also keep
971 + the system from drifting or rotating as a whole. Equivalently, the
972 + net linear momentum and angular momentum of the system should be
973 + shifted to zero.
974 +
975 + \subsubsection{Equilibration}
976 +
977 + The purpose of equilibration is to allow the system to evolve
978 + spontaneously for a period of time and reach equilibrium. The
979 + procedure is continued until various statistical properties, such as
980 + temperature, pressure, energy, volume and other structural
981 + properties \textit{etc}, become independent of time. Strictly
982 + speaking, minimization and heating are not necessary, provided the
983 + equilibration process is long enough. However, these steps can serve
984 + as a means to arrive at an equilibrated structure in an effective
985 + way.
986 +
987 + \subsection{\label{introSection:production}Production}
988 +
989 + \subsubsection{\label{introSec:forceCalculation}The Force Calculation}
990 +
991 + \subsubsection{\label{introSection:integrationSchemes} Integration
992 + Schemes}
993 +
994 + \subsection{\label{introSection:Analysis} Analysis}
995 +
996 + Recently, advanced visualization technique are widely applied to
997 + monitor the motions of molecules. Although the dynamics of the
998 + system can be described qualitatively from animation, quantitative
999 + trajectory analysis are more appreciable. According to the
1000 + principles of Statistical Mechanics,
1001 + Sec.~\ref{introSection:statisticalMechanics}, one can compute
1002 + thermodynamics properties, analyze fluctuations of structural
1003 + parameters, and investigate time-dependent processes of the molecule
1004 + from the trajectories.
1005 +
1006 + \subsubsection{\label{introSection:thermodynamicsProperties}Thermodynamics Properties}
1007 +
1008 + \subsubsection{\label{introSection:structuralProperties}Structural Properties}
1009 +
1010 + Structural Properties of a simple fluid can be described by a set of
1011 + distribution functions. Among these functions,\emph{pair
1012 + distribution function}, also known as \emph{radial distribution
1013 + function}, are of most fundamental importance to liquid-state
1014 + theory. Pair distribution function can be gathered by Fourier
1015 + transforming raw data from a series of neutron diffraction
1016 + experiments and integrating over the surface factor \cite{Powles73}.
1017 + The experiment result can serve as a criterion to justify the
1018 + correctness of the theory. Moreover, various equilibrium
1019 + thermodynamic and structural properties can also be expressed in
1020 + terms of radial distribution function \cite{allen87:csl}.
1021 +
1022 + A pair distribution functions $g(r)$ gives the probability that a
1023 + particle $i$ will be located at a distance $r$ from a another
1024 + particle $j$ in the system
1025   \[
1026 < L\left( {\int_0^t {g(t - \tau )h(\tau )d\tau } } \right) = G(p)H(p)
1026 > g(r) = \frac{V}{{N^2 }}\left\langle {\sum\limits_i {\sum\limits_{j
1027 > \ne i} {\delta (r - r_{ij} )} } } \right\rangle.
1028   \]
1029 + Note that the delta function can be replaced by a histogram in
1030 + computer simulation. Figure
1031 + \ref{introFigure:pairDistributionFunction} shows a typical pair
1032 + distribution function for the liquid argon system. The occurrence of
1033 + several peaks in the plot of $g(r)$ suggests that it is more likely
1034 + to find particles at certain radial values than at others. This is a
1035 + result of the attractive interaction at such distances. Because of
1036 + the strong repulsive forces at short distance, the probability of
1037 + locating particles at distances less than about 2.5{\AA} from each
1038 + other is essentially zero.
1039  
1040 < Some relatively important transformation,
1041 < \[
1042 < L(\cos at) = \frac{p}{{p^2  + a^2 }}
1040 > %\begin{figure}
1041 > %\centering
1042 > %\includegraphics[width=\linewidth]{pdf.eps}
1043 > %\caption[Pair distribution function for the liquid argon
1044 > %]{Pair distribution function for the liquid argon}
1045 > %\label{introFigure:pairDistributionFunction}
1046 > %\end{figure}
1047 >
1048 > \subsubsection{\label{introSection:timeDependentProperties}Time-dependent
1049 > Properties}
1050 >
1051 > Time-dependent properties are usually calculated using \emph{time
1052 > correlation function}, which correlates random variables $A$ and $B$
1053 > at two different time
1054 > \begin{equation}
1055 > C_{AB} (t) = \left\langle {A(t)B(0)} \right\rangle.
1056 > \label{introEquation:timeCorrelationFunction}
1057 > \end{equation}
1058 > If $A$ and $B$ refer to same variable, this kind of correlation
1059 > function is called \emph{auto correlation function}. One example of
1060 > auto correlation function is velocity auto-correlation function
1061 > which is directly related to transport properties of molecular
1062 > liquids. Another example is the calculation of the IR spectrum
1063 > through a Fourier transform of the dipole autocorrelation function.
1064 >
1065 > \section{\label{introSection:rigidBody}Dynamics of Rigid Bodies}
1066 >
1067 > Rigid bodies are frequently involved in the modeling of different
1068 > areas, from engineering, physics, to chemistry. For example,
1069 > missiles and vehicle are usually modeled by rigid bodies.  The
1070 > movement of the objects in 3D gaming engine or other physics
1071 > simulator is governed by the rigid body dynamics. In molecular
1072 > simulation, rigid body is used to simplify the model in
1073 > protein-protein docking study{\cite{Gray03}}.
1074 >
1075 > It is very important to develop stable and efficient methods to
1076 > integrate the equations of motion of orientational degrees of
1077 > freedom. Euler angles are the nature choice to describe the
1078 > rotational degrees of freedom. However, due to its singularity, the
1079 > numerical integration of corresponding equations of motion is very
1080 > inefficient and inaccurate. Although an alternative integrator using
1081 > different sets of Euler angles can overcome this difficulty\cite{},
1082 > the computational penalty and the lost of angular momentum
1083 > conservation still remain. A singularity free representation
1084 > utilizing quaternions was developed by Evans in 1977. Unfortunately,
1085 > this approach suffer from the nonseparable Hamiltonian resulted from
1086 > quaternion representation, which prevents the symplectic algorithm
1087 > to be utilized. Another different approach is to apply holonomic
1088 > constraints to the atoms belonging to the rigid body. Each atom
1089 > moves independently under the normal forces deriving from potential
1090 > energy and constraint forces which are used to guarantee the
1091 > rigidness. However, due to their iterative nature, SHAKE and Rattle
1092 > algorithm converge very slowly when the number of constraint
1093 > increases.
1094 >
1095 > The break through in geometric literature suggests that, in order to
1096 > develop a long-term integration scheme, one should preserve the
1097 > symplectic structure of the flow. Introducing conjugate momentum to
1098 > rotation matrix $Q$ and re-formulating Hamiltonian's equation, a
1099 > symplectic integrator, RSHAKE, was proposed to evolve the
1100 > Hamiltonian system in a constraint manifold by iteratively
1101 > satisfying the orthogonality constraint $Q_T Q = 1$. An alternative
1102 > method using quaternion representation was developed by Omelyan.
1103 > However, both of these methods are iterative and inefficient. In
1104 > this section, we will present a symplectic Lie-Poisson integrator
1105 > for rigid body developed by Dullweber and his
1106 > coworkers\cite{Dullweber1997} in depth.
1107 >
1108 > \subsection{\label{introSection:constrainedHamiltonianRB}Constrained Hamiltonian for Rigid Body}
1109 > The motion of the rigid body is Hamiltonian with the Hamiltonian
1110 > function
1111 > \begin{equation}
1112 > H = \frac{1}{2}(p^T m^{ - 1} p) + \frac{1}{2}tr(PJ^{ - 1} P) +
1113 > V(q,Q) + \frac{1}{2}tr[(QQ^T  - 1)\Lambda ].
1114 > \label{introEquation:RBHamiltonian}
1115 > \end{equation}
1116 > Here, $q$ and $Q$  are the position and rotation matrix for the
1117 > rigid-body, $p$ and $P$  are conjugate momenta to $q$  and $Q$ , and
1118 > $J$, a diagonal matrix, is defined by
1119 > \[
1120 > I_{ii}^{ - 1}  = \frac{1}{2}\sum\limits_{i \ne j} {J_{jj}^{ - 1} }
1121   \]
1122 + where $I_{ii}$ is the diagonal element of the inertia tensor. This
1123 + constrained Hamiltonian equation subjects to a holonomic constraint,
1124 + \begin{equation}
1125 + Q^T Q = 1$, \label{introEquation:orthogonalConstraint}
1126 + \end{equation}
1127 + which is used to ensure rotation matrix's orthogonality.
1128 + Differentiating \ref{introEquation:orthogonalConstraint} and using
1129 + Equation \ref{introEquation:RBMotionMomentum}, one may obtain,
1130 + \begin{equation}
1131 + Q^T PJ^{ - 1}  + J^{ - 1} P^T Q = 0 . \\
1132 + \label{introEquation:RBFirstOrderConstraint}
1133 + \end{equation}
1134  
1135 + Using Equation (\ref{introEquation:motionHamiltonianCoordinate},
1136 + \ref{introEquation:motionHamiltonianMomentum}), one can write down
1137 + the equations of motion,
1138   \[
1139 < L(\sin at) = \frac{a}{{p^2  + a^2 }}
1139 > \begin{array}{c}
1140 > \frac{{dq}}{{dt}} = \frac{p}{m} \label{introEquation:RBMotionPosition}\\
1141 > \frac{{dp}}{{dt}} =  - \nabla _q V(q,Q) \label{introEquation:RBMotionMomentum}\\
1142 > \frac{{dQ}}{{dt}} = PJ^{ - 1}  \label{introEquation:RBMotionRotation}\\
1143 > \frac{{dP}}{{dt}} =  - \nabla _Q V(q,Q) - 2Q\Lambda . \label{introEquation:RBMotionP}\\
1144 > \end{array}
1145   \]
1146  
1147 + In general, there are two ways to satisfy the holonomic constraints.
1148 + We can use constraint force provided by lagrange multiplier on the
1149 + normal manifold to keep the motion on constraint space. Or we can
1150 + simply evolve the system in constraint manifold. The two method are
1151 + proved to be equivalent. The holonomic constraint and equations of
1152 + motions define a constraint manifold for rigid body
1153   \[
1154 < L(1) = \frac{1}{p}
1154 > M = \left\{ {(Q,P):Q^T Q = 1,Q^T PJ^{ - 1}  + J^{ - 1} P^T Q = 0}
1155 > \right\}.
1156   \]
1157  
1158 < First, the bath coordinates,
1158 > Unfortunately, this constraint manifold is not the cotangent bundle
1159 > $T_{\star}SO(3)$. However, it turns out that under symplectic
1160 > transformation, the cotangent space and the phase space are
1161 > diffeomorphic. Introducing
1162   \[
1163 < p^2 L(x_\alpha  ) - px_\alpha  (0) - \dot x_\alpha  (0) =  - \omega
465 < _\alpha ^2 L(x_\alpha  ) + \frac{{g_\alpha  }}{{\omega _\alpha
466 < }}L(x)
1163 > \tilde Q = Q,\tilde P = \frac{1}{2}\left( {P - QP^T Q} \right),
1164   \]
1165 + the mechanical system subject to a holonomic constraint manifold $M$
1166 + can be re-formulated as a Hamiltonian system on the cotangent space
1167   \[
1168 < L(x_\alpha  ) = \frac{{\frac{{g_\alpha  }}{{\omega _\alpha  }}L(x) +
1169 < px_\alpha  (0) + \dot x_\alpha  (0)}}{{p^2  + \omega _\alpha ^2 }}
1168 > T^* SO(3) = \left\{ {(\tilde Q,\tilde P):\tilde Q^T \tilde Q =
1169 > 1,\tilde Q^T \tilde PJ^{ - 1}  + J^{ - 1} P^T \tilde Q = 0} \right\}
1170   \]
472 Then, the system coordinates,
473 \begin{align}
474 mL(\ddot x) &=  - \frac{1}{p}\frac{{\partial W(x)}}{{\partial x}} -
475 \sum\limits_{\alpha  = 1}^N {\left\{ {\frac{{\frac{{g_\alpha
476 }}{{\omega _\alpha  }}L(x) + px_\alpha  (0) + \dot x_\alpha
477 (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,
1171  
1172 + For a body fixed vector $X_i$ with respect to the center of mass of
1173 + the rigid body, its corresponding lab fixed vector $X_0^{lab}$  is
1174 + given as
1175 + \begin{equation}
1176 + X_i^{lab} = Q X_i + q.
1177 + \end{equation}
1178 + Therefore, potential energy $V(q,Q)$ is defined by
1179 + \[
1180 + V(q,Q) = V(Q X_0 + q).
1181 + \]
1182 + Hence, the force and torque are given by
1183 + \[
1184 + \nabla _q V(q,Q) = F(q,Q) = \sum\limits_i {F_i (q,Q)},
1185 + \]
1186 + and
1187 + \[
1188 + \nabla _Q V(q,Q) = F(q,Q)X_i^t
1189 + \]
1190 + respectively.
1191 +
1192 + As a common choice to describe the rotation dynamics of the rigid
1193 + body, angular momentum on body frame $\Pi  = Q^t P$ is introduced to
1194 + rewrite the equations of motion,
1195 + \begin{equation}
1196 + \begin{array}{l}
1197 + \mathop \Pi \limits^ \bullet   = J^{ - 1} \Pi ^T \Pi  + Q^T \sum\limits_i {F_i (q,Q)X_i^T }  - \Lambda  \\
1198 + \mathop Q\limits^{{\rm{   }} \bullet }  = Q\Pi {\rm{ }}J^{ - 1}  \\
1199 + \end{array}
1200 + \label{introEqaution:RBMotionPI}
1201 + \end{equation}
1202 + , as well as holonomic constraints,
1203 + \[
1204 + \begin{array}{l}
1205 + \Pi J^{ - 1}  + J^{ - 1} \Pi ^t  = 0 \\
1206 + Q^T Q = 1 \\
1207 + \end{array}
1208 + \]
1209 +
1210 + For a vector $v(v_1 ,v_2 ,v_3 ) \in R^3$ and a matrix $\hat v \in
1211 + so(3)^ \star$, the hat-map isomorphism,
1212 + \begin{equation}
1213 + v(v_1 ,v_2 ,v_3 ) \Leftrightarrow \hat v = \left(
1214 + {\begin{array}{*{20}c}
1215 +   0 & { - v_3 } & {v_2 }  \\
1216 +   {v_3 } & 0 & { - v_1 }  \\
1217 +   { - v_2 } & {v_1 } & 0  \\
1218 + \end{array}} \right),
1219 + \label{introEquation:hatmapIsomorphism}
1220 + \end{equation}
1221 + will let us associate the matrix products with traditional vector
1222 + operations
1223 + \[
1224 + \hat vu = v \times u
1225 + \]
1226 +
1227 + Using \ref{introEqaution:RBMotionPI}, one can construct a skew
1228 + matrix,
1229 + \begin{equation}
1230 + (\mathop \Pi \limits^ \bullet   - \mathop \Pi \limits^ \bullet  ^T
1231 + ){\rm{ }} = {\rm{ }}(\Pi  - \Pi ^T ){\rm{ }}(J^{ - 1} \Pi  + \Pi J^{
1232 + - 1} ) + \sum\limits_i {[Q^T F_i (r,Q)X_i^T  - X_i F_i (r,Q)^T Q]} -
1233 + (\Lambda  - \Lambda ^T ) . \label{introEquation:skewMatrixPI}
1234 + \end{equation}
1235 + Since $\Lambda$ is symmetric, the last term of Equation
1236 + \ref{introEquation:skewMatrixPI} is zero, which implies the Lagrange
1237 + multiplier $\Lambda$ is absent from the equations of motion. This
1238 + unique property eliminate the requirement of iterations which can
1239 + not be avoided in other methods\cite{}.
1240 +
1241 + Applying hat-map isomorphism, we obtain the equation of motion for
1242 + angular momentum on body frame
1243 + \begin{equation}
1244 + \dot \pi  = \pi  \times I^{ - 1} \pi  + \sum\limits_i {\left( {Q^T
1245 + F_i (r,Q)} \right) \times X_i }.
1246 + \label{introEquation:bodyAngularMotion}
1247 + \end{equation}
1248 + In the same manner, the equation of motion for rotation matrix is
1249 + given by
1250 + \[
1251 + \dot Q = Qskew(I^{ - 1} \pi )
1252 + \]
1253 +
1254 + \subsection{\label{introSection:SymplecticFreeRB}Symplectic
1255 + Lie-Poisson Integrator for Free Rigid Body}
1256 +
1257 + If there is not external forces exerted on the rigid body, the only
1258 + contribution to the rotational is from the kinetic potential (the
1259 + first term of \ref{ introEquation:bodyAngularMotion}). The free
1260 + rigid body is an example of Lie-Poisson system with Hamiltonian
1261 + function
1262 + \begin{equation}
1263 + T^r (\pi ) = T_1 ^r (\pi _1 ) + T_2^r (\pi _2 ) + T_3^r (\pi _3 )
1264 + \label{introEquation:rotationalKineticRB}
1265 + \end{equation}
1266 + where $T_i^r (\pi _i ) = \frac{{\pi _i ^2 }}{{2I_i }}$ and
1267 + Lie-Poisson structure matrix,
1268 + \begin{equation}
1269 + J(\pi ) = \left( {\begin{array}{*{20}c}
1270 +   0 & {\pi _3 } & { - \pi _2 }  \\
1271 +   { - \pi _3 } & 0 & {\pi _1 }  \\
1272 +   {\pi _2 } & { - \pi _1 } & 0  \\
1273 + \end{array}} \right)
1274 + \end{equation}
1275 + Thus, the dynamics of free rigid body is governed by
1276 + \begin{equation}
1277 + \frac{d}{{dt}}\pi  = J(\pi )\nabla _\pi  T^r (\pi )
1278 + \end{equation}
1279 +
1280 + One may notice that each $T_i^r$ in Equation
1281 + \ref{introEquation:rotationalKineticRB} can be solved exactly. For
1282 + instance, the equations of motion due to $T_1^r$ are given by
1283 + \begin{equation}
1284 + \frac{d}{{dt}}\pi  = R_1 \pi ,\frac{d}{{dt}}Q = QR_1
1285 + \label{introEqaution:RBMotionSingleTerm}
1286 + \end{equation}
1287 + where
1288 + \[ R_1  = \left( {\begin{array}{*{20}c}
1289 +   0 & 0 & 0  \\
1290 +   0 & 0 & {\pi _1 }  \\
1291 +   0 & { - \pi _1 } & 0  \\
1292 + \end{array}} \right).
1293 + \]
1294 + The solutions of Equation \ref{introEqaution:RBMotionSingleTerm} is
1295 + \[
1296 + \pi (\Delta t) = e^{\Delta tR_1 } \pi (0),Q(\Delta t) =
1297 + Q(0)e^{\Delta tR_1 }
1298 + \]
1299 + with
1300 + \[
1301 + e^{\Delta tR_1 }  = \left( {\begin{array}{*{20}c}
1302 +   0 & 0 & 0  \\
1303 +   0 & {\cos \theta _1 } & {\sin \theta _1 }  \\
1304 +   0 & { - \sin \theta _1 } & {\cos \theta _1 }  \\
1305 + \end{array}} \right),\theta _1  = \frac{{\pi _1 }}{{I_1 }}\Delta t.
1306 + \]
1307 + To reduce the cost of computing expensive functions in $e^{\Delta
1308 + tR_1 }$, we can use Cayley transformation,
1309 + \[
1310 + e^{\Delta tR_1 }  \approx (1 - \Delta tR_1 )^{ - 1} (1 + \Delta tR_1
1311 + )
1312 + \]
1313 + The flow maps for $T_2^r$ and $T_3^r$ can be found in the same
1314 + manner.
1315 +
1316 + In order to construct a second-order symplectic method, we split the
1317 + angular kinetic Hamiltonian function can into five terms
1318 + \[
1319 + T^r (\pi ) = \frac{1}{2}T_1 ^r (\pi _1 ) + \frac{1}{2}T_2^r (\pi _2
1320 + ) + T_3^r (\pi _3 ) + \frac{1}{2}T_2^r (\pi _2 ) + \frac{1}{2}T_1 ^r
1321 + (\pi _1 )
1322 + \].
1323 + Concatenating flows corresponding to these five terms, we can obtain
1324 + an symplectic integrator,
1325 + \[
1326 + \varphi _{\Delta t,T^r }  = \varphi _{\Delta t/2,\pi _1 }  \circ
1327 + \varphi _{\Delta t/2,\pi _2 }  \circ \varphi _{\Delta t,\pi _3 }
1328 + \circ \varphi _{\Delta t/2,\pi _2 }  \circ \varphi _{\Delta t/2,\pi
1329 + _1 }.
1330 + \]
1331 +
1332 + The non-canonical Lie-Poisson bracket ${F, G}$ of two function
1333 + $F(\pi )$ and $G(\pi )$ is defined by
1334 + \[
1335 + \{ F,G\} (\pi ) = [\nabla _\pi  F(\pi )]^T J(\pi )\nabla _\pi  G(\pi
1336 + )
1337 + \]
1338 + If the Poisson bracket of a function $F$ with an arbitrary smooth
1339 + function $G$ is zero, $F$ is a \emph{Casimir}, which is the
1340 + conserved quantity in Poisson system. We can easily verify that the
1341 + norm of the angular momentum, $\parallel \pi
1342 + \parallel$, is a \emph{Casimir}. Let$ F(\pi ) = S(\frac{{\parallel
1343 + \pi \parallel ^2 }}{2})$ for an arbitrary function $ S:R \to R$ ,
1344 + then by the chain rule
1345 + \[
1346 + \nabla _\pi  F(\pi ) = S'(\frac{{\parallel \pi \parallel ^2
1347 + }}{2})\pi
1348 + \]
1349 + Thus $ [\nabla _\pi  F(\pi )]^T J(\pi ) =  - S'(\frac{{\parallel \pi
1350 + \parallel ^2 }}{2})\pi  \times \pi  = 0 $. This explicit
1351 + Lie-Poisson integrator is found to be extremely efficient and stable
1352 + which can be explained by the fact the small angle approximation is
1353 + used and the norm of the angular momentum is conserved.
1354 +
1355 + \subsection{\label{introSection:RBHamiltonianSplitting} Hamiltonian
1356 + Splitting for Rigid Body}
1357 +
1358 + The Hamiltonian of rigid body can be separated in terms of kinetic
1359 + energy and potential energy,
1360 + \[
1361 + H = T(p,\pi ) + V(q,Q)
1362 + \]
1363 + The equations of motion corresponding to potential energy and
1364 + kinetic energy are listed in the below table,
1365 + \begin{center}
1366 + \begin{tabular}{|l|l|}
1367 +  \hline
1368 +  % after \\: \hline or \cline{col1-col2} \cline{col3-col4} ...
1369 +  Potential & Kinetic \\
1370 +  $\frac{{dq}}{{dt}} = \frac{p}{m}$ & $\frac{d}{{dt}}q = p$ \\
1371 +  $\frac{d}{{dt}}p =  - \frac{{\partial V}}{{\partial q}}$ & $ \frac{d}{{dt}}p = 0$ \\
1372 +  $\frac{d}{{dt}}Q = 0$ & $ \frac{d}{{dt}}Q = Qskew(I^{ - 1} j)$ \\
1373 +  $ \frac{d}{{dt}}\pi  = \sum\limits_i {\left( {Q^T F_i (r,Q)} \right) \times X_i }$ & $\frac{d}{{dt}}\pi  = \pi  \times I^{ - 1} \pi$\\
1374 +  \hline
1375 + \end{tabular}
1376 + \end{center}
1377 + A second-order symplectic method is now obtained by the composition
1378 + of the flow maps,
1379 + \[
1380 + \varphi _{\Delta t}  = \varphi _{\Delta t/2,V}  \circ \varphi
1381 + _{\Delta t,T}  \circ \varphi _{\Delta t/2,V}.
1382 + \]
1383 + Moreover, $\varphi _{\Delta t/2,V}$ can be divided into two
1384 + sub-flows which corresponding to force and torque respectively,
1385 + \[
1386 + \varphi _{\Delta t/2,V}  = \varphi _{\Delta t/2,F}  \circ \varphi
1387 + _{\Delta t/2,\tau }.
1388 + \]
1389 + Since the associated operators of $\varphi _{\Delta t/2,F} $ and
1390 + $\circ \varphi _{\Delta t/2,\tau }$ are commuted, the composition
1391 + order inside $\varphi _{\Delta t/2,V}$ does not matter.
1392 +
1393 + Furthermore, kinetic potential can be separated to translational
1394 + kinetic term, $T^t (p)$, and rotational kinetic term, $T^r (\pi )$,
1395 + \begin{equation}
1396 + T(p,\pi ) =T^t (p) + T^r (\pi ).
1397 + \end{equation}
1398 + where $ T^t (p) = \frac{1}{2}p^T m^{ - 1} p $ and $T^r (\pi )$ is
1399 + defined by \ref{introEquation:rotationalKineticRB}. Therefore, the
1400 + corresponding flow maps are given by
1401 + \[
1402 + \varphi _{\Delta t,T}  = \varphi _{\Delta t,T^t }  \circ \varphi
1403 + _{\Delta t,T^r }.
1404 + \]
1405 + Finally, we obtain the overall symplectic flow maps for free moving
1406 + rigid body
1407 + \begin{equation}
1408 + \begin{array}{c}
1409 + \varphi _{\Delta t}  = \varphi _{\Delta t/2,F}  \circ \varphi _{\Delta t/2,\tau }  \\
1410 +  \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 }  \\
1411 +  \circ \varphi _{\Delta t/2,\tau }  \circ \varphi _{\Delta t/2,F}  .\\
1412 + \end{array}
1413 + \label{introEquation:overallRBFlowMaps}
1414 + \end{equation}
1415 +
1416 + \section{\label{introSection:langevinDynamics}Langevin Dynamics}
1417 + As an alternative to newtonian dynamics, Langevin dynamics, which
1418 + mimics a simple heat bath with stochastic and dissipative forces,
1419 + has been applied in a variety of studies. This section will review
1420 + the theory of Langevin dynamics simulation. A brief derivation of
1421 + generalized Langevin equation will be given first. Follow that, we
1422 + will discuss the physical meaning of the terms appearing in the
1423 + equation as well as the calculation of friction tensor from
1424 + hydrodynamics theory.
1425 +
1426 + \subsection{\label{introSection:generalizedLangevinDynamics}Derivation of Generalized Langevin Equation}
1427 +
1428 + Harmonic bath model, in which an effective set of harmonic
1429 + oscillators are used to mimic the effect of a linearly responding
1430 + environment, has been widely used in quantum chemistry and
1431 + statistical mechanics. One of the successful applications of
1432 + Harmonic bath model is the derivation of Deriving Generalized
1433 + Langevin Dynamics. Lets consider a system, in which the degree of
1434 + freedom $x$ is assumed to couple to the bath linearly, giving a
1435 + Hamiltonian of the form
1436 + \begin{equation}
1437 + H = \frac{{p^2 }}{{2m}} + U(x) + H_B  + \Delta U(x,x_1 , \ldots x_N)
1438 + \label{introEquation:bathGLE}.
1439 + \end{equation}
1440 + Here $p$ is a momentum conjugate to $q$, $m$ is the mass associated
1441 + with this degree of freedom, $H_B$ is harmonic bath Hamiltonian,
1442 + \[
1443 + H_B  = \sum\limits_{\alpha  = 1}^N {\left\{ {\frac{{p_\alpha ^2
1444 + }}{{2m_\alpha  }} + \frac{1}{2}m_\alpha  \omega _\alpha ^2 }
1445 + \right\}}
1446 + \]
1447 + where the index $\alpha$ runs over all the bath degrees of freedom,
1448 + $\omega _\alpha$ are the harmonic bath frequencies, $m_\alpha$ are
1449 + the harmonic bath masses, and $\Delta U$ is bilinear system-bath
1450 + coupling,
1451 + \[
1452 + \Delta U =  - \sum\limits_{\alpha  = 1}^N {g_\alpha  x_\alpha  x}
1453 + \]
1454 + where $g_\alpha$ are the coupling constants between the bath and the
1455 + coordinate $x$. Introducing
1456 + \[
1457 + W(x) = U(x) - \sum\limits_{\alpha  = 1}^N {\frac{{g_\alpha ^2
1458 + }}{{2m_\alpha  w_\alpha ^2 }}} x^2
1459 + \] and combining the last two terms in Equation
1460 + \ref{introEquation:bathGLE}, we may rewrite the Harmonic bath
1461 + Hamiltonian as
1462 + \[
1463 + H = \frac{{p^2 }}{{2m}} + W(x) + \sum\limits_{\alpha  = 1}^N
1464 + {\left\{ {\frac{{p_\alpha ^2 }}{{2m_\alpha  }} + \frac{1}{2}m_\alpha
1465 + w_\alpha ^2 \left( {x_\alpha   - \frac{{g_\alpha  }}{{m_\alpha
1466 + w_\alpha ^2 }}x} \right)^2 } \right\}}
1467 + \]
1468 + Since the first two terms of the new Hamiltonian depend only on the
1469 + system coordinates, we can get the equations of motion for
1470 + Generalized Langevin Dynamics by Hamilton's equations
1471 + \ref{introEquation:motionHamiltonianCoordinate,
1472 + introEquation:motionHamiltonianMomentum},
1473 + \begin{equation}
1474 + m\ddot x =  - \frac{{\partial W(x)}}{{\partial x}} -
1475 + \sum\limits_{\alpha  = 1}^N {g_\alpha  \left( {x_\alpha   -
1476 + \frac{{g_\alpha  }}{{m_\alpha  w_\alpha ^2 }}x} \right)},
1477 + \label{introEquation:coorMotionGLE}
1478 + \end{equation}
1479 + and
1480 + \begin{equation}
1481 + m\ddot x_\alpha   =  - m_\alpha  w_\alpha ^2 \left( {x_\alpha   -
1482 + \frac{{g_\alpha  }}{{m_\alpha  w_\alpha ^2 }}x} \right).
1483 + \label{introEquation:bathMotionGLE}
1484 + \end{equation}
1485 +
1486 + In order to derive an equation for $x$, the dynamics of the bath
1487 + variables $x_\alpha$ must be solved exactly first. As an integral
1488 + transform which is particularly useful in solving linear ordinary
1489 + differential equations, Laplace transform is the appropriate tool to
1490 + solve this problem. The basic idea is to transform the difficult
1491 + differential equations into simple algebra problems which can be
1492 + solved easily. Then applying inverse Laplace transform, also known
1493 + as the Bromwich integral, we can retrieve the solutions of the
1494 + original problems.
1495 +
1496 + Let $f(t)$ be a function defined on $ [0,\infty ) $. The Laplace
1497 + transform of f(t) is a new function defined as
1498 + \[
1499 + L(f(t)) \equiv F(p) = \int_0^\infty  {f(t)e^{ - pt} dt}
1500 + \]
1501 + where  $p$ is real and  $L$ is called the Laplace Transform
1502 + Operator. Below are some important properties of Laplace transform
1503 + \begin{equation}
1504 + \begin{array}{c}
1505 + L(x + y) = L(x) + L(y) \\
1506 + L(ax) = aL(x) \\
1507 + L(\dot x) = pL(x) - px(0) \\
1508 + L(\ddot x) = p^2 L(x) - px(0) - \dot x(0) \\
1509 + L\left( {\int_0^t {g(t - \tau )h(\tau )d\tau } } \right) = G(p)H(p) \\
1510 + \end{array}
1511 + \end{equation}
1512 +
1513 + Applying Laplace transform to the bath coordinates, we obtain
1514 + \[
1515 + \begin{array}{c}
1516 + p^2 L(x_\alpha  ) - px_\alpha  (0) - \dot x_\alpha  (0) =  - \omega _\alpha ^2 L(x_\alpha  ) + \frac{{g_\alpha  }}{{\omega _\alpha  }}L(x) \\
1517 + L(x_\alpha  ) = \frac{{\frac{{g_\alpha  }}{{\omega _\alpha  }}L(x) + px_\alpha  (0) + \dot x_\alpha  (0)}}{{p^2  + \omega _\alpha ^2 }} \\
1518 + \end{array}
1519 + \]
1520 + By the same way, the system coordinates become
1521 + \[
1522 + \begin{array}{c}
1523 + mL(\ddot x) =  - \frac{1}{p}\frac{{\partial W(x)}}{{\partial x}} \\
1524 +  - \sum\limits_{\alpha  = 1}^N {\left\{ { - \frac{{g_\alpha ^2 }}{{m_\alpha  \omega _\alpha ^2 }}\frac{p}{{p^2  + \omega _\alpha ^2 }}pL(x) - \frac{p}{{p^2  + \omega _\alpha ^2 }}g_\alpha  x_\alpha  (0) - \frac{1}{{p^2  + \omega _\alpha ^2 }}g_\alpha  \dot x_\alpha  (0)} \right\}}  \\
1525 + \end{array}
1526 + \]
1527 +
1528 + With the help of some relatively important inverse Laplace
1529 + transformations:
1530 + \[
1531 + \begin{array}{c}
1532 + L(\cos at) = \frac{p}{{p^2  + a^2 }} \\
1533 + L(\sin at) = \frac{a}{{p^2  + a^2 }} \\
1534 + L(1) = \frac{1}{p} \\
1535 + \end{array}
1536 + \]
1537 + , we obtain
1538   \begin{align}
1539   m\ddot x &=  - \frac{{\partial W(x)}}{{\partial x}} -
1540   \sum\limits_{\alpha  = 1}^N {\left\{ {\left( { - \frac{{g_\alpha ^2
# Line 503 | Line 1554 | t)\dot x(t - \tau )d} \tau }  + \sum\limits_{\alpha  =
1554   (\omega _\alpha  t)} \right\}}
1555   \end{align}
1556  
1557 + Introducing a \emph{dynamic friction kernel}
1558   \begin{equation}
1559 + \xi (t) = \sum\limits_{\alpha  = 1}^N {\left( { - \frac{{g_\alpha ^2
1560 + }}{{m_\alpha  \omega _\alpha ^2 }}} \right)\cos (\omega _\alpha  t)}
1561 + \label{introEquation:dynamicFrictionKernelDefinition}
1562 + \end{equation}
1563 + and \emph{a random force}
1564 + \begin{equation}
1565 + R(t) = \sum\limits_{\alpha  = 1}^N {\left( {g_\alpha  x_\alpha  (0)
1566 + - \frac{{g_\alpha ^2 }}{{m_\alpha  \omega _\alpha ^2 }}x(0)}
1567 + \right)\cos (\omega _\alpha  t)}  + \frac{{\dot x_\alpha
1568 + (0)}}{{\omega _\alpha  }}\sin (\omega _\alpha  t),
1569 + \label{introEquation:randomForceDefinition}
1570 + \end{equation}
1571 + the equation of motion can be rewritten as
1572 + \begin{equation}
1573   m\ddot x =  - \frac{{\partial W}}{{\partial x}} - \int_0^t {\xi
1574   (t)\dot x(t - \tau )d\tau }  + R(t)
1575   \label{introEuqation:GeneralizedLangevinDynamics}
1576   \end{equation}
1577 < %where $ {\xi (t)}$ is friction kernel, $R(t)$ is random force and
1578 < %$W$ is the potential of mean force. $W(x) =  - kT\ln p(x)$
1577 > which is known as the \emph{generalized Langevin equation}.
1578 >
1579 > \subsubsection{\label{introSection:randomForceDynamicFrictionKernel}Random Force and Dynamic Friction Kernel}
1580 >
1581 > One may notice that $R(t)$ depends only on initial conditions, which
1582 > implies it is completely deterministic within the context of a
1583 > harmonic bath. However, it is easy to verify that $R(t)$ is totally
1584 > uncorrelated to $x$ and $\dot x$,
1585   \[
1586 < \xi (t) = \sum\limits_{\alpha  = 1}^N {\left( { - \frac{{g_\alpha ^2
1587 < }}{{m_\alpha  \omega _\alpha ^2 }}} \right)\cos (\omega _\alpha  t)}
1586 > \begin{array}{l}
1587 > \left\langle {x(t)R(t)} \right\rangle  = 0, \\
1588 > \left\langle {\dot x(t)R(t)} \right\rangle  = 0. \\
1589 > \end{array}
1590   \]
1591 < For an infinite harmonic bath, we can use the spectral density and
1592 < an integral over frequencies.
1591 > This property is what we expect from a truly random process. As long
1592 > as the model, which is gaussian distribution in general, chosen for
1593 > $R(t)$ is a truly random process, the stochastic nature of the GLE
1594 > still remains.
1595  
1596 + %dynamic friction kernel
1597 + The convolution integral
1598   \[
1599 < R(t) = \sum\limits_{\alpha  = 1}^N {\left( {g_\alpha  x_\alpha  (0)
522 < - \frac{{g_\alpha ^2 }}{{m_\alpha  \omega _\alpha ^2 }}x(0)}
523 < \right)\cos (\omega _\alpha  t)}  + \frac{{\dot x_\alpha
524 < (0)}}{{\omega _\alpha  }}\sin (\omega _\alpha  t)
1599 > \int_0^t {\xi (t)\dot x(t - \tau )d\tau }
1600   \]
1601 < The random forces depend only on initial conditions.
1601 > depends on the entire history of the evolution of $x$, which implies
1602 > that the bath retains memory of previous motions. In other words,
1603 > the bath requires a finite time to respond to change in the motion
1604 > of the system. For a sluggish bath which responds slowly to changes
1605 > in the system coordinate, we may regard $\xi(t)$ as a constant
1606 > $\xi(t) = \Xi_0$. Hence, the convolution integral becomes
1607 > \[
1608 > \int_0^t {\xi (t)\dot x(t - \tau )d\tau }  = \xi _0 (x(t) - x(0))
1609 > \]
1610 > and Equation \ref{introEuqation:GeneralizedLangevinDynamics} becomes
1611 > \[
1612 > m\ddot x =  - \frac{\partial }{{\partial x}}\left( {W(x) +
1613 > \frac{1}{2}\xi _0 (x - x_0 )^2 } \right) + R(t),
1614 > \]
1615 > which can be used to describe dynamic caging effect. The other
1616 > extreme is the bath that responds infinitely quickly to motions in
1617 > the system. Thus, $\xi (t)$ can be taken as a $delta$ function in
1618 > time:
1619 > \[
1620 > \xi (t) = 2\xi _0 \delta (t)
1621 > \]
1622 > Hence, the convolution integral becomes
1623 > \[
1624 > \int_0^t {\xi (t)\dot x(t - \tau )d\tau }  = 2\xi _0 \int_0^t
1625 > {\delta (t)\dot x(t - \tau )d\tau }  = \xi _0 \dot x(t),
1626 > \]
1627 > and Equation \ref{introEuqation:GeneralizedLangevinDynamics} becomes
1628 > \begin{equation}
1629 > m\ddot x =  - \frac{{\partial W(x)}}{{\partial x}} - \xi _0 \dot
1630 > x(t) + R(t) \label{introEquation:LangevinEquation}
1631 > \end{equation}
1632 > which is known as the Langevin equation. The static friction
1633 > coefficient $\xi _0$ can either be calculated from spectral density
1634 > or be determined by Stokes' law for regular shaped particles.A
1635 > briefly review on calculating friction tensor for arbitrary shaped
1636 > particles is given in Sec.~\ref{introSection:frictionTensor}.
1637  
1638   \subsubsection{\label{introSection:secondFluctuationDissipation}The Second Fluctuation Dissipation Theorem}
1639 < So we can define a new set of coordinates,
1639 >
1640 > Defining a new set of coordinates,
1641   \[
1642   q_\alpha  (t) = x_\alpha  (t) - \frac{1}{{m_\alpha  \omega _\alpha
1643   ^2 }}x(0)
1644 < \]
1645 < This makes
1644 > \],
1645 > we can rewrite $R(T)$ as
1646   \[
1647 < R(t) = \sum\limits_{\alpha  = 1}^N {g_\alpha  q_\alpha  (t)}
1647 > R(t) = \sum\limits_{\alpha  = 1}^N {g_\alpha  q_\alpha  (t)}.
1648   \]
1649   And since the $q$ coordinates are harmonic oscillators,
1650   \[
1651 < \begin{array}{l}
1651 > \begin{array}{c}
1652 > \left\langle {q_\alpha ^2 } \right\rangle  = \frac{{kT}}{{m_\alpha  \omega _\alpha ^2 }} \\
1653   \left\langle {q_\alpha  (t)q_\alpha  (0)} \right\rangle  = \left\langle {q_\alpha ^2 (0)} \right\rangle \cos (\omega _\alpha  t) \\
1654   \left\langle {q_\alpha  (t)q_\beta  (0)} \right\rangle  = \delta _{\alpha \beta } \left\langle {q_\alpha  (t)q_\alpha  (0)} \right\rangle  \\
1655 + \left\langle {R(t)R(0)} \right\rangle  = \sum\limits_\alpha  {\sum\limits_\beta  {g_\alpha  g_\beta  \left\langle {q_\alpha  (t)q_\beta  (0)} \right\rangle } }  \\
1656 +  = \sum\limits_\alpha  {g_\alpha ^2 \left\langle {q_\alpha ^2 (0)} \right\rangle \cos (\omega _\alpha  t)}  \\
1657 +  = kT\xi (t) \\
1658   \end{array}
1659   \]
1660 <
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 <
1660 > Thus, we recover the \emph{second fluctuation dissipation theorem}
1661   \begin{equation}
1662   \xi (t) = \left\langle {R(t)R(0)} \right\rangle
1663 < \label{introEquation:secondFluctuationDissipation}
1663 > \label{introEquation:secondFluctuationDissipation}.
1664   \end{equation}
1665 + In effect, it acts as a constraint on the possible ways in which one
1666 + can model the random force and friction kernel.
1667  
562 \section{\label{introSection:hydroynamics}Hydrodynamics}
563
1668   \subsection{\label{introSection:frictionTensor} Friction Tensor}
1669 < \subsection{\label{introSection:analyticalApproach}Analytical
1670 < Approach}
1669 > Theoretically, the friction kernel can be determined using velocity
1670 > autocorrelation function. However, this approach become impractical
1671 > when the system become more and more complicate. Instead, various
1672 > approaches based on hydrodynamics have been developed to calculate
1673 > the friction coefficients. The friction effect is isotropic in
1674 > Equation, \zeta can be taken as a scalar. In general, friction
1675 > tensor \Xi is a $6\times 6$ matrix given by
1676 > \[
1677 > \Xi  = \left( {\begin{array}{*{20}c}
1678 >   {\Xi _{}^{tt} } & {\Xi _{}^{rt} }  \\
1679 >   {\Xi _{}^{tr} } & {\Xi _{}^{rr} }  \\
1680 > \end{array}} \right).
1681 > \]
1682 > Here, $ {\Xi^{tt} }$ and $ {\Xi^{rr} }$ are translational friction
1683 > tensor and rotational resistance (friction) tensor respectively,
1684 > while ${\Xi^{tr} }$ is translation-rotation coupling tensor and $
1685 > {\Xi^{rt} }$ is rotation-translation coupling tensor. When a
1686 > particle moves in a fluid, it may experience friction force or
1687 > torque along the opposite direction of the velocity or angular
1688 > velocity,
1689 > \[
1690 > \left( \begin{array}{l}
1691 > F_R  \\
1692 > \tau _R  \\
1693 > \end{array} \right) =  - \left( {\begin{array}{*{20}c}
1694 >   {\Xi ^{tt} } & {\Xi ^{rt} }  \\
1695 >   {\Xi ^{tr} } & {\Xi ^{rr} }  \\
1696 > \end{array}} \right)\left( \begin{array}{l}
1697 > v \\
1698 > w \\
1699 > \end{array} \right)
1700 > \]
1701 > where $F_r$ is the friction force and $\tau _R$ is the friction
1702 > toque.
1703  
1704 < \subsection{\label{introSection:approximationApproach}Approximation
569 < Approach}
1704 > \subsubsection{\label{introSection:resistanceTensorRegular}The Resistance Tensor for Regular Shape}
1705  
1706 < \subsection{\label{introSection:centersRigidBody}Centers of Rigid
1707 < Body}
1706 > For a spherical particle, the translational and rotational friction
1707 > constant can be calculated from Stoke's law,
1708 > \[
1709 > \Xi ^{tt}  = \left( {\begin{array}{*{20}c}
1710 >   {6\pi \eta R} & 0 & 0  \\
1711 >   0 & {6\pi \eta R} & 0  \\
1712 >   0 & 0 & {6\pi \eta R}  \\
1713 > \end{array}} \right)
1714 > \]
1715 > and
1716 > \[
1717 > \Xi ^{rr}  = \left( {\begin{array}{*{20}c}
1718 >   {8\pi \eta R^3 } & 0 & 0  \\
1719 >   0 & {8\pi \eta R^3 } & 0  \\
1720 >   0 & 0 & {8\pi \eta R^3 }  \\
1721 > \end{array}} \right)
1722 > \]
1723 > where $\eta$ is the viscosity of the solvent and $R$ is the
1724 > hydrodynamics radius.
1725 >
1726 > Other non-spherical shape, such as cylinder and ellipsoid
1727 > \textit{etc}, are widely used as reference for developing new
1728 > hydrodynamics theory, because their properties can be calculated
1729 > exactly. In 1936, Perrin extended Stokes's law to general ellipsoid,
1730 > also called a triaxial ellipsoid, which is given in Cartesian
1731 > coordinates by
1732 > \[
1733 > \frac{{x^2 }}{{a^2 }} + \frac{{y^2 }}{{b^2 }} + \frac{{z^2 }}{{c^2
1734 > }} = 1
1735 > \]
1736 > where the semi-axes are of lengths $a$, $b$, and $c$. Unfortunately,
1737 > due to the complexity of the elliptic integral, only the ellipsoid
1738 > with the restriction of two axes having to be equal, \textit{i.e.}
1739 > prolate($ a \ge b = c$) and oblate ($ a < b = c $), can be solved
1740 > exactly. Introducing an elliptic integral parameter $S$ for prolate,
1741 > \[
1742 > S = \frac{2}{{\sqrt {a^2  - b^2 } }}\ln \frac{{a + \sqrt {a^2  - b^2
1743 > } }}{b},
1744 > \]
1745 > and oblate,
1746 > \[
1747 > S = \frac{2}{{\sqrt {b^2  - a^2 } }}arctg\frac{{\sqrt {b^2  - a^2 }
1748 > }}{a}
1749 > \],
1750 > one can write down the translational and rotational resistance
1751 > tensors
1752 > \[
1753 > \begin{array}{l}
1754 > \Xi _a^{tt}  = 16\pi \eta \frac{{a^2  - b^2 }}{{(2a^2  - b^2 )S - 2a}} \\
1755 > \Xi _b^{tt}  = \Xi _c^{tt}  = 32\pi \eta \frac{{a^2  - b^2 }}{{(2a^2  - 3b^2 )S + 2a}} \\
1756 > \end{array},
1757 > \]
1758 > and
1759 > \[
1760 > \begin{array}{l}
1761 > \Xi _a^{rr}  = \frac{{32\pi }}{3}\eta \frac{{(a^2  - b^2 )b^2 }}{{2a - b^2 S}} \\
1762 > \Xi _b^{rr}  = \Xi _c^{rr}  = \frac{{32\pi }}{3}\eta \frac{{(a^4  - b^4 )}}{{(2a^2  - b^2 )S - 2a}} \\
1763 > \end{array}.
1764 > \]
1765 >
1766 > \subsubsection{\label{introSection:resistanceTensorRegularArbitrary}The Resistance Tensor for Arbitrary Shape}
1767 >
1768 > Unlike spherical and other regular shaped molecules, there is not
1769 > analytical solution for friction tensor of any arbitrary shaped
1770 > rigid molecules. The ellipsoid of revolution model and general
1771 > triaxial ellipsoid model have been used to approximate the
1772 > hydrodynamic properties of rigid bodies. However, since the mapping
1773 > from all possible ellipsoidal space, $r$-space, to all possible
1774 > combination of rotational diffusion coefficients, $D$-space is not
1775 > unique\cite{Wegener79} as well as the intrinsic coupling between
1776 > translational and rotational motion of rigid body\cite{}, general
1777 > ellipsoid is not always suitable for modeling arbitrarily shaped
1778 > rigid molecule. A number of studies have been devoted to determine
1779 > the friction tensor for irregularly shaped rigid bodies using more
1780 > advanced method\cite{} where the molecule of interest was modeled by
1781 > combinations of spheres(beads)\cite{} and the hydrodynamics
1782 > properties of the molecule can be calculated using the hydrodynamic
1783 > interaction tensor. Let us consider a rigid assembly of $N$ beads
1784 > immersed in a continuous medium. Due to hydrodynamics interaction,
1785 > the ``net'' velocity of $i$th bead, $v'_i$ is different than its
1786 > unperturbed velocity $v_i$,
1787 > \[
1788 > v'_i  = v_i  - \sum\limits_{j \ne i} {T_{ij} F_j }
1789 > \]
1790 > where $F_i$ is the frictional force, and $T_{ij}$ is the
1791 > hydrodynamic interaction tensor. The friction force of $i$th bead is
1792 > proportional to its ``net'' velocity
1793 > \begin{equation}
1794 > F_i  = \zeta _i v_i  - \zeta _i \sum\limits_{j \ne i} {T_{ij} F_j }.
1795 > \label{introEquation:tensorExpression}
1796 > \end{equation}
1797 > This equation is the basis for deriving the hydrodynamic tensor. In
1798 > 1930, Oseen and Burgers gave a simple solution to Equation
1799 > \ref{introEquation:tensorExpression}
1800 > \begin{equation}
1801 > T_{ij}  = \frac{1}{{8\pi \eta r_{ij} }}\left( {I + \frac{{R_{ij}
1802 > R_{ij}^T }}{{R_{ij}^2 }}} \right).
1803 > \label{introEquation:oseenTensor}
1804 > \end{equation}
1805 > Here $R_{ij}$ is the distance vector between bead $i$ and bead $j$.
1806 > A second order expression for element of different size was
1807 > introduced by Rotne and Prager\cite{} and improved by Garc\'{i}a de
1808 > la Torre and Bloomfield,
1809 > \begin{equation}
1810 > T_{ij}  = \frac{1}{{8\pi \eta R_{ij} }}\left[ {\left( {I +
1811 > \frac{{R_{ij} R_{ij}^T }}{{R_{ij}^2 }}} \right) + R\frac{{\sigma
1812 > _i^2  + \sigma _j^2 }}{{r_{ij}^2 }}\left( {\frac{I}{3} -
1813 > \frac{{R_{ij} R_{ij}^T }}{{R_{ij}^2 }}} \right)} \right].
1814 > \label{introEquation:RPTensorNonOverlapped}
1815 > \end{equation}
1816 > Both of the Equation \ref{introEquation:oseenTensor} and Equation
1817 > \ref{introEquation:RPTensorNonOverlapped} have an assumption $R_{ij}
1818 > \ge \sigma _i  + \sigma _j$. An alternative expression for
1819 > overlapping beads with the same radius, $\sigma$, is given by
1820 > \begin{equation}
1821 > T_{ij}  = \frac{1}{{6\pi \eta R_{ij} }}\left[ {\left( {1 -
1822 > \frac{2}{{32}}\frac{{R_{ij} }}{\sigma }} \right)I +
1823 > \frac{2}{{32}}\frac{{R_{ij} R_{ij}^T }}{{R_{ij} \sigma }}} \right]
1824 > \label{introEquation:RPTensorOverlapped}
1825 > \end{equation}
1826 >
1827 > To calculate the resistance tensor at an arbitrary origin $O$, we
1828 > construct a $3N \times 3N$ matrix consisting of $N \times N$
1829 > $B_{ij}$ blocks
1830 > \begin{equation}
1831 > B = \left( {\begin{array}{*{20}c}
1832 >   {B_{11} } &  \ldots  & {B_{1N} }  \\
1833 >    \vdots  &  \ddots  &  \vdots   \\
1834 >   {B_{N1} } &  \cdots  & {B_{NN} }  \\
1835 > \end{array}} \right),
1836 > \end{equation}
1837 > where $B_{ij}$ is given by
1838 > \[
1839 > B_{ij}  = \delta _{ij} \frac{I}{{6\pi \eta R}} + (1 - \delta _{ij}
1840 > )T_{ij}
1841 > \]
1842 > where $\delta _{ij}$ is Kronecker delta function. Inverting matrix
1843 > $B$, we obtain
1844 >
1845 > \[
1846 > C = B^{ - 1}  = \left( {\begin{array}{*{20}c}
1847 >   {C_{11} } &  \ldots  & {C_{1N} }  \\
1848 >    \vdots  &  \ddots  &  \vdots   \\
1849 >   {C_{N1} } &  \cdots  & {C_{NN} }  \\
1850 > \end{array}} \right)
1851 > \]
1852 > , which can be partitioned into $N \times N$ $3 \times 3$ block
1853 > $C_{ij}$. With the help of $C_{ij}$ and skew matrix $U_i$
1854 > \[
1855 > U_i  = \left( {\begin{array}{*{20}c}
1856 >   0 & { - z_i } & {y_i }  \\
1857 >   {z_i } & 0 & { - x_i }  \\
1858 >   { - y_i } & {x_i } & 0  \\
1859 > \end{array}} \right)
1860 > \]
1861 > where $x_i$, $y_i$, $z_i$ are the components of the vector joining
1862 > bead $i$ and origin $O$. Hence, the elements of resistance tensor at
1863 > arbitrary origin $O$ can be written as
1864 > \begin{equation}
1865 > \begin{array}{l}
1866 > \Xi _{}^{tt}  = \sum\limits_i {\sum\limits_j {C_{ij} } } , \\
1867 > \Xi _{}^{tr}  = \Xi _{}^{rt}  = \sum\limits_i {\sum\limits_j {U_i C_{ij} } } , \\
1868 > \Xi _{}^{rr}  =  - \sum\limits_i {\sum\limits_j {U_i C_{ij} } } U_j  \\
1869 > \end{array}
1870 > \label{introEquation:ResistanceTensorArbitraryOrigin}
1871 > \end{equation}
1872 >
1873 > The resistance tensor depends on the origin to which they refer. The
1874 > proper location for applying friction force is the center of
1875 > resistance (reaction), at which the trace of rotational resistance
1876 > tensor, $ \Xi ^{rr}$ reaches minimum. Mathematically, the center of
1877 > resistance is defined as an unique point of the rigid body at which
1878 > the translation-rotation coupling tensor are symmetric,
1879 > \begin{equation}
1880 > \Xi^{tr}  = \left( {\Xi^{tr} } \right)^T
1881 > \label{introEquation:definitionCR}
1882 > \end{equation}
1883 > Form Equation \ref{introEquation:ResistanceTensorArbitraryOrigin},
1884 > we can easily find out that the translational resistance tensor is
1885 > origin independent, while the rotational resistance tensor and
1886 > translation-rotation coupling resistance tensor depend on the
1887 > origin. Given resistance tensor at an arbitrary origin $O$, and a
1888 > vector ,$r_{OP}(x_{OP}, y_{OP}, z_{OP})$, from $O$ to $P$, we can
1889 > obtain the resistance tensor at $P$ by
1890 > \begin{equation}
1891 > \begin{array}{l}
1892 > \Xi _P^{tt}  = \Xi _O^{tt}  \\
1893 > \Xi _P^{tr}  = \Xi _P^{rt}  = \Xi _O^{tr}  - U_{OP} \Xi _O^{tt}  \\
1894 > \Xi _P^{rr}  = \Xi _O^{rr}  - U_{OP} \Xi _O^{tt} U_{OP}  + \Xi _O^{tr} U_{OP}  - U_{OP} \Xi _O^{tr} ^{^T }  \\
1895 > \end{array}
1896 > \label{introEquation:resistanceTensorTransformation}
1897 > \end{equation}
1898 > where
1899 > \[
1900 > U_{OP}  = \left( {\begin{array}{*{20}c}
1901 >   0 & { - z_{OP} } & {y_{OP} }  \\
1902 >   {z_i } & 0 & { - x_{OP} }  \\
1903 >   { - y_{OP} } & {x_{OP} } & 0  \\
1904 > \end{array}} \right)
1905 > \]
1906 > Using Equations \ref{introEquation:definitionCR} and
1907 > \ref{introEquation:resistanceTensorTransformation}, one can locate
1908 > the position of center of resistance,
1909 > \[
1910 > \left( \begin{array}{l}
1911 > x_{OR}  \\
1912 > y_{OR}  \\
1913 > z_{OR}  \\
1914 > \end{array} \right) = \left( {\begin{array}{*{20}c}
1915 >   {(\Xi _O^{rr} )_{yy}  + (\Xi _O^{rr} )_{zz} } & { - (\Xi _O^{rr} )_{xy} } & { - (\Xi _O^{rr} )_{xz} }  \\
1916 >   { - (\Xi _O^{rr} )_{xy} } & {(\Xi _O^{rr} )_{zz}  + (\Xi _O^{rr} )_{xx} } & { - (\Xi _O^{rr} )_{yz} }  \\
1917 >   { - (\Xi _O^{rr} )_{xz} } & { - (\Xi _O^{rr} )_{yz} } & {(\Xi _O^{rr} )_{xx}  + (\Xi _O^{rr} )_{yy} }  \\
1918 > \end{array}} \right)^{ - 1} \left( \begin{array}{l}
1919 > (\Xi _O^{tr} )_{yz}  - (\Xi _O^{tr} )_{zy}  \\
1920 > (\Xi _O^{tr} )_{zx}  - (\Xi _O^{tr} )_{xz}  \\
1921 > (\Xi _O^{tr} )_{xy}  - (\Xi _O^{tr} )_{yx}  \\
1922 > \end{array} \right).
1923 > \]
1924 > where $x_OR$, $y_OR$, $z_OR$ are the components of the vector
1925 > joining center of resistance $R$ and origin $O$.

Diff Legend

Removed lines
+ Added lines
< Changed lines
> Changed lines