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 2694 by tim, Wed Apr 5 21:00:19 2006 UTC vs.
Revision 2716 by tim, Mon Apr 17 20:39:26 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 63 | Line 63 | that if all forces are conservative, Energy $E = T + V
63   \end{equation}
64   If there are no external torques acting on a body, the angular
65   momentum of it is conserved. The last conservation theorem state
66 < that if all forces are conservative, Energy $E = T + V$ is
67 < conserved. All of these conserved quantities are important factors
68 < to determine the quality of numerical integration scheme for rigid
69 < body \cite{Dullweber1997}.
66 > that if all forces are conservative, Energy
67 > \begin{equation}E = T + V \label{introEquation:energyConservation}
68 > \end{equation}
69 > is conserved. All of these conserved quantities are
70 > important factors to determine the quality of numerical integration
71 > scheme for rigid body \cite{Dullweber1997}.
72  
73   \subsection{\label{introSection:lagrangian}Lagrangian Mechanics}
74  
# Line 115 | 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 200 | Line 202 | When studying Hamiltonian system, it is more convenien
202   independent variables and it only works with 1st-order differential
203   equations\cite{Marion90}.
204  
205 < When studying Hamiltonian system, it is more convenient to use
206 < notation
205 > In Newtonian Mechanics, a system described by conservative forces
206 > conserves the total energy \ref{introEquation:energyConservation}.
207 > It follows that Hamilton's equations of motion conserve the total
208 > Hamiltonian.
209   \begin{equation}
210 < r = r(q,p)^T
210 > \frac{{dH}}{{dt}} = \sum\limits_i {\left( {\frac{{\partial
211 > H}}{{\partial q_i }}\dot q_i  + \frac{{\partial H}}{{\partial p_i
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} \label{introEquation:conserveHalmitonian}
216   \end{equation}
208 and to introduce a $2n \times 2n$ canonical structure matrix $J$,
209 \begin{equation}
210 J = \left( {\begin{array}{*{20}c}
211   0 & I  \\
212   { - I} & 0  \\
213 \end{array}} \right)
214 \label{introEquation:canonicalMatrix}
215 \end{equation}
216 Thus, Hamiltonian system can be rewritten as,
217 \begin{equation}
218 \frac{d}{{dt}}r = J\nabla _r H(r)
219 \label{introEquation:compactHamiltonian}
220 \end{equation}
221 where $I$ is an identity matrix and $J$ is a skew-symmetrix matrix
222 ($ J^T  =  - J $).
217  
224 %\subsection{\label{introSection:canonicalTransformation}Canonical
225 Transformation}
226
227 \section{\label{introSection:geometricIntegratos}Geometric Integrators}
228
229 \subsection{\label{introSection:symplecticMaps}Symplectic Maps and Methods}
230
231 \subsection{\label{Construction of Symplectic Methods}}
232
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}
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
464 + Dynamics simulation. By comparing experimental values with the
465 + calculated properties, one can determine the accuracy of the
466 + simulation and the quality of the underlying model. However, both of
467 + experiment and computer simulation are usually performed during a
468 + certain time interval and the measurements are averaged over a
469 + period of them which is different from the average behavior of
470 + many-body system in Statistical Mechanics. Fortunately, Ergodic
471 + Hypothesis is proposed to make a connection between time average and
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(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 $\langle  A(q , p) \rangle_t$ 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
495 + motions. They usually begin with an initial conditionals and move
496 + the objects in the direction governed by the differential equations.
497 + However, most of them ignore the hidden physical law contained
498 + within the equations. Since 1990, geometric integrators, which
499 + preserve various phase-flow invariants such as symplectic structure,
500 + volume and time reversal symmetry, are developed to address this
501 + issue. The velocity verlet method, which happens to be a simple
502 + example of symplectic integrator, continues to gain its popularity
503 + in molecular dynamics community. This fact can be partly explained
504 + by its geometric nature.
505 +
506 + \subsection{\label{introSection:symplecticManifold}Symplectic Manifold}
507 + A \emph{manifold} is an abstract mathematical space. It locally
508 + looks like Euclidean space, but when viewed globally, it may have
509 + more complicate structure. A good example of manifold is the surface
510 + of Earth. It seems to be flat locally, but it is round if viewed as
511 + a whole. A \emph{differentiable manifold} (also known as
512 + \emph{smooth manifold}) is a manifold with an open cover in which
513 + the covering neighborhoods are all smoothly isomorphic to one
514 + another. In other words,it is possible to apply calculus on
515 + \emph{differentiable manifold}. A \emph{symplectic manifold} is
516 + defined as a pair $(M, \omega)$ which consisting of a
517 + \emph{differentiable manifold} $M$ and a close, non-degenerated,
518 + bilinear symplectic form, $\omega$. A symplectic form on a vector
519 + space $V$ is a function $\omega(x, y)$ which satisfies
520 + $\omega(\lambda_1x_1+\lambda_2x_2, y) = \lambda_1\omega(x_1, y)+
521 + \lambda_2\omega(x_2, y)$, $\omega(x, y) = - \omega(y, x)$ and
522 + $\omega(x, x) = 0$. Cross product operation in vector field is an
523 + example of symplectic form.
524 +
525 + One of the motivations to study \emph{symplectic manifold} in
526 + Hamiltonian Mechanics is that a symplectic manifold can represent
527 + all possible configurations of the system and the phase space of the
528 + system can be described by it's cotangent bundle. Every symplectic
529 + manifold is even dimensional. For instance, in Hamilton equations,
530 + coordinate and momentum always appear in pairs.
531 +
532 + Let  $(M,\omega)$ and $(N, \eta)$ be symplectic manifolds. A map
533 + \[
534 + f : M \rightarrow N
535 + \]
536 + is a \emph{symplectomorphism} if it is a \emph{diffeomorphims} and
537 + the \emph{pullback} of $\eta$ under f is equal to $\omega$.
538 + Canonical transformation is an example of symplectomorphism in
539 + classical mechanics.
540 +
541 + \subsection{\label{introSection:ODE}Ordinary Differential Equations}
542 +
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 + 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 + \subsection{\label{introSection:exactFlow}Exact Flow}
575 +
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 + 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 + \subsection{\label{introSection:geometricProperties}Geometric Properties}
614 +
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 + Let $\varphi$ be the flow of Hamiltonian vector field, $\varphi$ is
620 + a \emph{symplectic} flow if it satisfies,
621 + \begin{equation}
622 + {\varphi '}^T J \varphi ' = J.
623 + \end{equation}
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 + \[
642 + \dot y = \tilde f(y) = ((dh \cdot f)h^{ - 1} )(y).
643 + \]
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 + \[
652 + \frac{{dG(x(t))}}{{dt}} = 0.
653 + \]
654 + Using chain rule, one may obtain,
655 + \[
656 + \sum\limits_i {\frac{{dG}}{{dx_i }}} f_i (x) = f \bullet \nabla G,
657 + \]
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 + \[
671 + \frac{d}{{dt}}G(x(t)) = \left\{ {G,H} \right\}(x(t)).
672 + \]
673 + Therefore, the sufficient condition for $G$ to be the \emph{first
674 + integral} of a Hamiltonian system is
675 + \[
676 + \left\{ {G,H} \right\} = 0.
677 + \]
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 + 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 + 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{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 + \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:positionVerlet1}
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 + \[
845 + [X,Y] = XY - YX .
846 + \]
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 + \[
868 + \varphi _h^{(4)}  = \varphi _{\alpha h}^{(2)}  \circ \varphi _{\beta
869 + h}^{(2)}  \circ \varphi _{\alpha h}^{(2)}
870 + \]
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 + \[
880 + \alpha  =  - \frac{{2^{1/(2n + 1)} }}{{2 - 2^{1/(2n + 1)} }},\beta =
881 + \frac{{2^{1/(2n + 1)} }}{{2 - 2^{1/(2n + 1)} }} .
882 + \]
883 +
884   \section{\label{introSection:molecularDynamics}Molecular Dynamics}
885  
886   As a special discipline of molecular modeling, Molecular dynamics
# Line 252 | Line 890 | dynamical information.
890  
891   \subsection{\label{introSec:mdInit}Initialization}
892  
893 + \subsection{\label{introSec:forceEvaluation}Force Evaluation}
894 +
895   \subsection{\label{introSection:mdIntegration} Integration of the Equations of Motion}
896  
897   \section{\label{introSection:rigidBody}Dynamics of Rigid Bodies}
898  
899 < A rigid body is a body in which the distance between any two given
900 < points of a rigid body remains constant regardless of external
901 < forces exerted on it. A rigid body therefore conserves its shape
902 < during its motion.
899 > Rigid bodies are frequently involved in the modeling of different
900 > areas, from engineering, physics, to chemistry. For example,
901 > missiles and vehicle are usually modeled by rigid bodies.  The
902 > movement of the objects in 3D gaming engine or other physics
903 > simulator is governed by the rigid body dynamics. In molecular
904 > simulation, rigid body is used to simplify the model in
905 > protein-protein docking study{\cite{Gray03}}.
906  
907 < Applications of dynamics of rigid bodies.
907 > It is very important to develop stable and efficient methods to
908 > integrate the equations of motion of orientational degrees of
909 > freedom. Euler angles are the nature choice to describe the
910 > rotational degrees of freedom. However, due to its singularity, the
911 > numerical integration of corresponding equations of motion is very
912 > inefficient and inaccurate. Although an alternative integrator using
913 > different sets of Euler angles can overcome this difficulty\cite{},
914 > the computational penalty and the lost of angular momentum
915 > conservation still remain. A singularity free representation
916 > utilizing quaternions was developed by Evans in 1977. Unfortunately,
917 > this approach suffer from the nonseparable Hamiltonian resulted from
918 > quaternion representation, which prevents the symplectic algorithm
919 > to be utilized. Another different approach is to apply holonomic
920 > constraints to the atoms belonging to the rigid body. Each atom
921 > moves independently under the normal forces deriving from potential
922 > energy and constraint forces which are used to guarantee the
923 > rigidness. However, due to their iterative nature, SHAKE and Rattle
924 > algorithm converge very slowly when the number of constraint
925 > increases.
926  
927 <
928 < %\subsection{\label{introSection:poissonBrackets}Poisson Brackets}
927 > The break through in geometric literature suggests that, in order to
928 > develop a long-term integration scheme, one should preserve the
929 > symplectic structure of the flow. Introducing conjugate momentum to
930 > rotation matrix $A$ and re-formulating Hamiltonian's equation, a
931 > symplectic integrator, RSHAKE, was proposed to evolve the
932 > Hamiltonian system in a constraint manifold by iteratively
933 > satisfying the orthogonality constraint $A_t A = 1$. An alternative
934 > method using quaternion representation was developed by Omelyan.
935 > However, both of these methods are iterative and inefficient. In
936 > this section, we will present a symplectic Lie-Poisson integrator
937 > for rigid body developed by Dullweber and his
938 > coworkers\cite{Dullweber1997} in depth.
939  
940 < \section{\label{introSection:correlationFunctions}Correlation Functions}
940 > \subsection{\label{introSection:constrainedHamiltonianRB}Constrained Hamiltonian for Rigid Body}
941 > The motion of the rigid body is Hamiltonian with the Hamiltonian
942 > function
943 > \begin{equation}
944 > H = \frac{1}{2}(p^T m^{ - 1} p) + \frac{1}{2}tr(PJ^{ - 1} P) +
945 > V(q,Q) + \frac{1}{2}tr[(QQ^T  - 1)\Lambda ].
946 > \label{introEquation:RBHamiltonian}
947 > \end{equation}
948 > Here, $q$ and $Q$  are the position and rotation matrix for the
949 > rigid-body, $p$ and $P$  are conjugate momenta to $q$  and $Q$ , and
950 > $J$, a diagonal matrix, is defined by
951 > \[
952 > I_{ii}^{ - 1}  = \frac{1}{2}\sum\limits_{i \ne j} {J_{jj}^{ - 1} }
953 > \]
954 > where $I_{ii}$ is the diagonal element of the inertia tensor. This
955 > constrained Hamiltonian equation subjects to a holonomic constraint,
956 > \begin{equation}
957 > Q^T Q = 1$, \label{introEquation:orthogonalConstraint}
958 > \end{equation}
959 > which is used to ensure rotation matrix's orthogonality.
960 > Differentiating \ref{introEquation:orthogonalConstraint} and using
961 > Equation \ref{introEquation:RBMotionMomentum}, one may obtain,
962 > \begin{equation}
963 > Q^T PJ^{ - 1}  + J^{ - 1} P^T Q = 0 . \\
964 > \label{introEquation:RBFirstOrderConstraint}
965 > \end{equation}
966  
967 + Using Equation (\ref{introEquation:motionHamiltonianCoordinate},
968 + \ref{introEquation:motionHamiltonianMomentum}), one can write down
969 + the equations of motion,
970 + \[
971 + \begin{array}{c}
972 + \frac{{dq}}{{dt}} = \frac{p}{m} \label{introEquation:RBMotionPosition}\\
973 + \frac{{dp}}{{dt}} =  - \nabla _q V(q,Q) \label{introEquation:RBMotionMomentum}\\
974 + \frac{{dQ}}{{dt}} = PJ^{ - 1}  \label{introEquation:RBMotionRotation}\\
975 + \frac{{dP}}{{dt}} =  - \nabla _Q V(q,Q) - 2Q\Lambda . \label{introEquation:RBMotionP}\\
976 + \end{array}
977 + \]
978 +
979 + In general, there are two ways to satisfy the holonomic constraints.
980 + We can use constraint force provided by lagrange multiplier on the
981 + normal manifold to keep the motion on constraint space. Or we can
982 + simply evolve the system in constraint manifold. The two method are
983 + proved to be equivalent. The holonomic constraint and equations of
984 + motions define a constraint manifold for rigid body
985 + \[
986 + M = \left\{ {(Q,P):Q^T Q = 1,Q^T PJ^{ - 1}  + J^{ - 1} P^T Q = 0}
987 + \right\}.
988 + \]
989 +
990 + Unfortunately, this constraint manifold is not the cotangent bundle
991 + $T_{\star}SO(3)$. However, it turns out that under symplectic
992 + transformation, the cotangent space and the phase space are
993 + diffeomorphic. Introducing
994 + \[
995 + \tilde Q = Q,\tilde P = \frac{1}{2}\left( {P - QP^T Q} \right),
996 + \]
997 + the mechanical system subject to a holonomic constraint manifold $M$
998 + can be re-formulated as a Hamiltonian system on the cotangent space
999 + \[
1000 + T^* SO(3) = \left\{ {(\tilde Q,\tilde P):\tilde Q^T \tilde Q =
1001 + 1,\tilde Q^T \tilde PJ^{ - 1}  + J^{ - 1} P^T \tilde Q = 0} \right\}
1002 + \]
1003 +
1004 + For a body fixed vector $X_i$ with respect to the center of mass of
1005 + the rigid body, its corresponding lab fixed vector $X_0^{lab}$  is
1006 + given as
1007 + \begin{equation}
1008 + X_i^{lab} = Q X_i + q.
1009 + \end{equation}
1010 + Therefore, potential energy $V(q,Q)$ is defined by
1011 + \[
1012 + V(q,Q) = V(Q X_0 + q).
1013 + \]
1014 + Hence, the force and torque are given by
1015 + \[
1016 + \nabla _q V(q,Q) = F(q,Q) = \sum\limits_i {F_i (q,Q)},
1017 + \]
1018 + and
1019 + \[
1020 + \nabla _Q V(q,Q) = F(q,Q)X_i^t
1021 + \]
1022 + respectively.
1023 +
1024 + As a common choice to describe the rotation dynamics of the rigid
1025 + body, angular momentum on body frame $\Pi  = Q^t P$ is introduced to
1026 + rewrite the equations of motion,
1027 + \begin{equation}
1028 + \begin{array}{l}
1029 + \mathop \Pi \limits^ \bullet   = J^{ - 1} \Pi ^T \Pi  + Q^T \sum\limits_i {F_i (q,Q)X_i^T }  - \Lambda  \\
1030 + \mathop Q\limits^{{\rm{   }} \bullet }  = Q\Pi {\rm{ }}J^{ - 1}  \\
1031 + \end{array}
1032 + \label{introEqaution:RBMotionPI}
1033 + \end{equation}
1034 + , as well as holonomic constraints,
1035 + \[
1036 + \begin{array}{l}
1037 + \Pi J^{ - 1}  + J^{ - 1} \Pi ^t  = 0 \\
1038 + Q^T Q = 1 \\
1039 + \end{array}
1040 + \]
1041 +
1042 + For a vector $v(v_1 ,v_2 ,v_3 ) \in R^3$ and a matrix $\hat v \in
1043 + so(3)^ \star$, the hat-map isomorphism,
1044 + \begin{equation}
1045 + v(v_1 ,v_2 ,v_3 ) \Leftrightarrow \hat v = \left(
1046 + {\begin{array}{*{20}c}
1047 +   0 & { - v_3 } & {v_2 }  \\
1048 +   {v_3 } & 0 & { - v_1 }  \\
1049 +   { - v_2 } & {v_1 } & 0  \\
1050 + \end{array}} \right),
1051 + \label{introEquation:hatmapIsomorphism}
1052 + \end{equation}
1053 + will let us associate the matrix products with traditional vector
1054 + operations
1055 + \[
1056 + \hat vu = v \times u
1057 + \]
1058 +
1059 + Using \ref{introEqaution:RBMotionPI}, one can construct a skew
1060 + matrix,
1061 + \begin{equation}
1062 + (\mathop \Pi \limits^ \bullet   - \mathop \Pi \limits^ \bullet  ^T
1063 + ){\rm{ }} = {\rm{ }}(\Pi  - \Pi ^T ){\rm{ }}(J^{ - 1} \Pi  + \Pi J^{
1064 + - 1} ) + \sum\limits_i {[Q^T F_i (r,Q)X_i^T  - X_i F_i (r,Q)^T Q]} -
1065 + (\Lambda  - \Lambda ^T ) . \label{introEquation:skewMatrixPI}
1066 + \end{equation}
1067 + Since $\Lambda$ is symmetric, the last term of Equation
1068 + \ref{introEquation:skewMatrixPI} is zero, which implies the Lagrange
1069 + multiplier $\Lambda$ is absent from the equations of motion. This
1070 + unique property eliminate the requirement of iterations which can
1071 + not be avoided in other methods\cite{}.
1072 +
1073 + Applying hat-map isomorphism, we obtain the equation of motion for
1074 + angular momentum on body frame
1075 + \begin{equation}
1076 + \dot \pi  = \pi  \times I^{ - 1} \pi  + \sum\limits_i {\left( {Q^T
1077 + F_i (r,Q)} \right) \times X_i }.
1078 + \label{introEquation:bodyAngularMotion}
1079 + \end{equation}
1080 + In the same manner, the equation of motion for rotation matrix is
1081 + given by
1082 + \[
1083 + \dot Q = Qskew(I^{ - 1} \pi )
1084 + \]
1085 +
1086 + \subsection{\label{introSection:SymplecticFreeRB}Symplectic
1087 + Lie-Poisson Integrator for Free Rigid Body}
1088 +
1089 + If there is not external forces exerted on the rigid body, the only
1090 + contribution to the rotational is from the kinetic potential (the
1091 + first term of \ref{ introEquation:bodyAngularMotion}). The free
1092 + rigid body is an example of Lie-Poisson system with Hamiltonian
1093 + function
1094 + \begin{equation}
1095 + T^r (\pi ) = T_1 ^r (\pi _1 ) + T_2^r (\pi _2 ) + T_3^r (\pi _3 )
1096 + \label{introEquation:rotationalKineticRB}
1097 + \end{equation}
1098 + where $T_i^r (\pi _i ) = \frac{{\pi _i ^2 }}{{2I_i }}$ and
1099 + Lie-Poisson structure matrix,
1100 + \begin{equation}
1101 + J(\pi ) = \left( {\begin{array}{*{20}c}
1102 +   0 & {\pi _3 } & { - \pi _2 }  \\
1103 +   { - \pi _3 } & 0 & {\pi _1 }  \\
1104 +   {\pi _2 } & { - \pi _1 } & 0  \\
1105 + \end{array}} \right)
1106 + \end{equation}
1107 + Thus, the dynamics of free rigid body is governed by
1108 + \begin{equation}
1109 + \frac{d}{{dt}}\pi  = J(\pi )\nabla _\pi  T^r (\pi )
1110 + \end{equation}
1111 +
1112 + One may notice that each $T_i^r$ in Equation
1113 + \ref{introEquation:rotationalKineticRB} can be solved exactly. For
1114 + instance, the equations of motion due to $T_1^r$ are given by
1115 + \begin{equation}
1116 + \frac{d}{{dt}}\pi  = R_1 \pi ,\frac{d}{{dt}}Q = QR_1
1117 + \label{introEqaution:RBMotionSingleTerm}
1118 + \end{equation}
1119 + where
1120 + \[ R_1  = \left( {\begin{array}{*{20}c}
1121 +   0 & 0 & 0  \\
1122 +   0 & 0 & {\pi _1 }  \\
1123 +   0 & { - \pi _1 } & 0  \\
1124 + \end{array}} \right).
1125 + \]
1126 + The solutions of Equation \ref{introEqaution:RBMotionSingleTerm} is
1127 + \[
1128 + \pi (\Delta t) = e^{\Delta tR_1 } \pi (0),Q(\Delta t) =
1129 + Q(0)e^{\Delta tR_1 }
1130 + \]
1131 + with
1132 + \[
1133 + e^{\Delta tR_1 }  = \left( {\begin{array}{*{20}c}
1134 +   0 & 0 & 0  \\
1135 +   0 & {\cos \theta _1 } & {\sin \theta _1 }  \\
1136 +   0 & { - \sin \theta _1 } & {\cos \theta _1 }  \\
1137 + \end{array}} \right),\theta _1  = \frac{{\pi _1 }}{{I_1 }}\Delta t.
1138 + \]
1139 + To reduce the cost of computing expensive functions in e^{\Delta
1140 + tR_1 }, we can use Cayley transformation,
1141 + \[
1142 + e^{\Delta tR_1 }  \approx (1 - \Delta tR_1 )^{ - 1} (1 + \Delta tR_1
1143 + )
1144 + \]
1145 +
1146 + The flow maps for $T_2^r$ and $T_2^r$ can be found in the same
1147 + manner.
1148 +
1149 + In order to construct a second-order symplectic method, we split the
1150 + angular kinetic Hamiltonian function can into five terms
1151 + \[
1152 + T^r (\pi ) = \frac{1}{2}T_1 ^r (\pi _1 ) + \frac{1}{2}T_2^r (\pi _2
1153 + ) + T_3^r (\pi _3 ) + \frac{1}{2}T_2^r (\pi _2 ) + \frac{1}{2}T_1 ^r
1154 + (\pi _1 )
1155 + \].
1156 + Concatenating flows corresponding to these five terms, we can obtain
1157 + an symplectic integrator,
1158 + \[
1159 + \varphi _{\Delta t,T^r }  = \varphi _{\Delta t/2,\pi _1 }  \circ
1160 + \varphi _{\Delta t/2,\pi _2 }  \circ \varphi _{\Delta t,\pi _3 }
1161 + \circ \varphi _{\Delta t/2,\pi _2 }  \circ \varphi _{\Delta t/2,\pi
1162 + _1 }.
1163 + \]
1164 +
1165 + The non-canonical Lie-Poisson bracket ${F, G}$ of two function
1166 + $F(\pi )$ and $G(\pi )$ is defined by
1167 + \[
1168 + \{ F,G\} (\pi ) = [\nabla _\pi  F(\pi )]^T J(\pi )\nabla _\pi  G(\pi
1169 + )
1170 + \]
1171 + If the Poisson bracket of a function $F$ with an arbitrary smooth
1172 + function $G$ is zero, $F$ is a \emph{Casimir}, which is the
1173 + conserved quantity in Poisson system. We can easily verify that the
1174 + norm of the angular momentum, $\parallel \pi
1175 + \parallel$, is a \emph{Casimir}. Let$ F(\pi ) = S(\frac{{\parallel
1176 + \pi \parallel ^2 }}{2})$ for an arbitrary function $ S:R \to R$ ,
1177 + then by the chain rule
1178 + \[
1179 + \nabla _\pi  F(\pi ) = S'(\frac{{\parallel \pi \parallel ^2
1180 + }}{2})\pi
1181 + \]
1182 + Thus $ [\nabla _\pi  F(\pi )]^T J(\pi ) =  - S'(\frac{{\parallel \pi
1183 + \parallel ^2 }}{2})\pi  \times \pi  = 0 $. This explicit
1184 + Lie-Poisson integrator is found to be extremely efficient and stable
1185 + which can be explained by the fact the small angle approximation is
1186 + used and the norm of the angular momentum is conserved.
1187 +
1188 + \subsection{\label{introSection:RBHamiltonianSplitting} Hamiltonian
1189 + Splitting for Rigid Body}
1190 +
1191 + The Hamiltonian of rigid body can be separated in terms of kinetic
1192 + energy and potential energy,
1193 + \[
1194 + H = T(p,\pi ) + V(q,Q)
1195 + \]
1196 + The equations of motion corresponding to potential energy and
1197 + kinetic energy are listed in the below table,
1198 + \begin{center}
1199 + \begin{tabular}{|l|l|}
1200 +  \hline
1201 +  % after \\: \hline or \cline{col1-col2} \cline{col3-col4} ...
1202 +  Potential & Kinetic \\
1203 +  $\frac{{dq}}{{dt}} = \frac{p}{m}$ & $\frac{d}{{dt}}q = p$ \\
1204 +  $\frac{d}{{dt}}p =  - \frac{{\partial V}}{{\partial q}}$ & $ \frac{d}{{dt}}p = 0$ \\
1205 +  $\frac{d}{{dt}}Q = 0$ & $ \frac{d}{{dt}}Q = Qskew(I^{ - 1} j)$ \\
1206 +  $ \frac{d}{{dt}}\pi  = \sum\limits_i {\left( {Q^T F_i (r,Q)} \right) \times X_i }$ & $\frac{d}{{dt}}\pi  = \pi  \times I^{ - 1} \pi$\\
1207 +  \hline
1208 + \end{tabular}
1209 + \end{center}
1210 + A second-order symplectic method is now obtained by the composition
1211 + of the flow maps,
1212 + \[
1213 + \varphi _{\Delta t}  = \varphi _{\Delta t/2,V}  \circ \varphi
1214 + _{\Delta t,T}  \circ \varphi _{\Delta t/2,V}.
1215 + \]
1216 + Moreover, \varphi _{\Delta t/2,V} can be divided into two sub-flows
1217 + which corresponding to force and torque respectively,
1218 + \[
1219 + \varphi _{\Delta t/2,V}  = \varphi _{\Delta t/2,F}  \circ \varphi
1220 + _{\Delta t/2,\tau }.
1221 + \]
1222 + Since the associated operators of $\varphi _{\Delta t/2,F} $ and
1223 + $\circ \varphi _{\Delta t/2,\tau }$ are commuted, the composition
1224 + order inside \varphi _{\Delta t/2,V} does not matter.
1225 +
1226 + Furthermore, kinetic potential can be separated to translational
1227 + kinetic term, $T^t (p)$, and rotational kinetic term, $T^r (\pi )$,
1228 + \begin{equation}
1229 + T(p,\pi ) =T^t (p) + T^r (\pi ).
1230 + \end{equation}
1231 + where $ T^t (p) = \frac{1}{2}p^T m^{ - 1} p $ and $T^r (\pi )$ is
1232 + defined by \ref{introEquation:rotationalKineticRB}. Therefore, the
1233 + corresponding flow maps are given by
1234 + \[
1235 + \varphi _{\Delta t,T}  = \varphi _{\Delta t,T^t }  \circ \varphi
1236 + _{\Delta t,T^r }.
1237 + \]
1238 + Finally, we obtain the overall symplectic flow maps for free moving
1239 + rigid body
1240 + \begin{equation}
1241 + \begin{array}{c}
1242 + \varphi _{\Delta t}  = \varphi _{\Delta t/2,F}  \circ \varphi _{\Delta t/2,\tau }  \\
1243 +  \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 }  \\
1244 +  \circ \varphi _{\Delta t/2,\tau }  \circ \varphi _{\Delta t/2,F}  .\\
1245 + \end{array}
1246 + \label{introEquation:overallRBFlowMaps}
1247 + \end{equation}
1248 +
1249   \section{\label{introSection:langevinDynamics}Langevin Dynamics}
1250 + As an alternative to newtonian dynamics, Langevin dynamics, which
1251 + mimics a simple heat bath with stochastic and dissipative forces,
1252 + has been applied in a variety of studies. This section will review
1253 + the theory of Langevin dynamics simulation. A brief derivation of
1254 + generalized Langevin Dynamics will be given first. Follow that, we
1255 + will discuss the physical meaning of the terms appearing in the
1256 + equation as well as the calculation of friction tensor from
1257 + hydrodynamics theory.
1258  
1259   \subsection{\label{introSection:generalizedLangevinDynamics}Generalized Langevin Dynamics}
1260  
1261 < \subsection{\label{introSection:hydroynamics}Hydrodynamics}
1261 > \begin{equation}
1262 > H = \frac{{p^2 }}{{2m}} + U(x) + H_B  + \Delta U(x,x_1 , \ldots x_N)
1263 > \label{introEquation:bathGLE}
1264 > \end{equation}
1265 > where $H_B$ is harmonic bath Hamiltonian,
1266 > \[
1267 > H_B  =\sum\limits_{\alpha  = 1}^N {\left\{ {\frac{{p_\alpha ^2
1268 > }}{{2m_\alpha  }} + \frac{1}{2}m_\alpha  w_\alpha ^2 } \right\}}
1269 > \]
1270 > and $\Delta U$ is bilinear system-bath coupling,
1271 > \[
1272 > \Delta U =  - \sum\limits_{\alpha  = 1}^N {g_\alpha  x_\alpha  x}
1273 > \]
1274 > Completing the square,
1275 > \[
1276 > H_B  + \Delta U = \sum\limits_{\alpha  = 1}^N {\left\{
1277 > {\frac{{p_\alpha ^2 }}{{2m_\alpha  }} + \frac{1}{2}m_\alpha
1278 > w_\alpha ^2 \left( {x_\alpha   - \frac{{g_\alpha  }}{{m_\alpha
1279 > w_\alpha ^2 }}x} \right)^2 } \right\}}  - \sum\limits_{\alpha  =
1280 > 1}^N {\frac{{g_\alpha ^2 }}{{2m_\alpha  w_\alpha ^2 }}} x^2
1281 > \]
1282 > and putting it back into Eq.~\ref{introEquation:bathGLE},
1283 > \[
1284 > H = \frac{{p^2 }}{{2m}} + W(x) + \sum\limits_{\alpha  = 1}^N
1285 > {\left\{ {\frac{{p_\alpha ^2 }}{{2m_\alpha  }} + \frac{1}{2}m_\alpha
1286 > w_\alpha ^2 \left( {x_\alpha   - \frac{{g_\alpha  }}{{m_\alpha
1287 > w_\alpha ^2 }}x} \right)^2 } \right\}}
1288 > \]
1289 > where
1290 > \[
1291 > W(x) = U(x) - \sum\limits_{\alpha  = 1}^N {\frac{{g_\alpha ^2
1292 > }}{{2m_\alpha  w_\alpha ^2 }}} x^2
1293 > \]
1294 > Since the first two terms of the new Hamiltonian depend only on the
1295 > system coordinates, we can get the equations of motion for
1296 > Generalized Langevin Dynamics by Hamilton's equations
1297 > \ref{introEquation:motionHamiltonianCoordinate,
1298 > introEquation:motionHamiltonianMomentum},
1299 > \begin{align}
1300 > \dot p &=  - \frac{{\partial H}}{{\partial x}}
1301 >       &= m\ddot x
1302 >       &= - \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)}
1303 > \label{introEquation:Lp5}
1304 > \end{align}
1305 > , and
1306 > \begin{align}
1307 > \dot p_\alpha   &=  - \frac{{\partial H}}{{\partial x_\alpha  }}
1308 >                &= m\ddot x_\alpha
1309 >                &= \- m_\alpha  w_\alpha ^2 \left( {x_\alpha   - \frac{{g_\alpha}}{{m_\alpha  w_\alpha ^2 }}x} \right)
1310 > \end{align}
1311 >
1312 > \subsection{\label{introSection:laplaceTransform}The Laplace Transform}
1313 >
1314 > \[
1315 > L(x) = \int_0^\infty  {x(t)e^{ - pt} dt}
1316 > \]
1317 >
1318 > \[
1319 > L(x + y) = L(x) + L(y)
1320 > \]
1321 >
1322 > \[
1323 > L(ax) = aL(x)
1324 > \]
1325 >
1326 > \[
1327 > L(\dot x) = pL(x) - px(0)
1328 > \]
1329 >
1330 > \[
1331 > L(\ddot x) = p^2 L(x) - px(0) - \dot x(0)
1332 > \]
1333 >
1334 > \[
1335 > L\left( {\int_0^t {g(t - \tau )h(\tau )d\tau } } \right) = G(p)H(p)
1336 > \]
1337 >
1338 > Some relatively important transformation,
1339 > \[
1340 > L(\cos at) = \frac{p}{{p^2  + a^2 }}
1341 > \]
1342 >
1343 > \[
1344 > L(\sin at) = \frac{a}{{p^2  + a^2 }}
1345 > \]
1346 >
1347 > \[
1348 > L(1) = \frac{1}{p}
1349 > \]
1350 >
1351 > First, the bath coordinates,
1352 > \[
1353 > p^2 L(x_\alpha  ) - px_\alpha  (0) - \dot x_\alpha  (0) =  - \omega
1354 > _\alpha ^2 L(x_\alpha  ) + \frac{{g_\alpha  }}{{\omega _\alpha
1355 > }}L(x)
1356 > \]
1357 > \[
1358 > L(x_\alpha  ) = \frac{{\frac{{g_\alpha  }}{{\omega _\alpha  }}L(x) +
1359 > px_\alpha  (0) + \dot x_\alpha  (0)}}{{p^2  + \omega _\alpha ^2 }}
1360 > \]
1361 > Then, the system coordinates,
1362 > \begin{align}
1363 > mL(\ddot x) &=  - \frac{1}{p}\frac{{\partial W(x)}}{{\partial x}} -
1364 > \sum\limits_{\alpha  = 1}^N {\left\{ {\frac{{\frac{{g_\alpha
1365 > }}{{\omega _\alpha  }}L(x) + px_\alpha  (0) + \dot x_\alpha
1366 > (0)}}{{p^2  + \omega _\alpha ^2 }} - \frac{{g_\alpha ^2 }}{{m_\alpha
1367 > }}\omega _\alpha ^2 L(x)} \right\}}
1368 > %
1369 > &= - \frac{1}{p}\frac{{\partial W(x)}}{{\partial x}} -
1370 > \sum\limits_{\alpha  = 1}^N {\left\{ { - \frac{{g_\alpha ^2 }}{{m_\alpha  \omega _\alpha ^2 }}\frac{p}{{p^2  + \omega _\alpha ^2 }}pL(x)
1371 > - \frac{p}{{p^2  + \omega _\alpha ^2 }}g_\alpha  x_\alpha  (0)
1372 > - \frac{1}{{p^2  + \omega _\alpha ^2 }}g_\alpha  \dot x_\alpha  (0)} \right\}}
1373 > \end{align}
1374 > Then, the inverse transform,
1375 >
1376 > \begin{align}
1377 > m\ddot x &=  - \frac{{\partial W(x)}}{{\partial x}} -
1378 > \sum\limits_{\alpha  = 1}^N {\left\{ {\left( { - \frac{{g_\alpha ^2
1379 > }}{{m_\alpha  \omega _\alpha ^2 }}} \right)\int_0^t {\cos (\omega
1380 > _\alpha  t)\dot x(t - \tau )d\tau  - \left[ {g_\alpha  x_\alpha  (0)
1381 > - \frac{{g_\alpha  }}{{m_\alpha  \omega _\alpha  }}} \right]\cos
1382 > (\omega _\alpha  t) - \frac{{g_\alpha  \dot x_\alpha  (0)}}{{\omega
1383 > _\alpha  }}\sin (\omega _\alpha  t)} } \right\}}
1384 > %
1385 > &= - \frac{{\partial W(x)}}{{\partial x}} - \int_0^t
1386 > {\sum\limits_{\alpha  = 1}^N {\left( { - \frac{{g_\alpha ^2
1387 > }}{{m_\alpha  \omega _\alpha ^2 }}} \right)\cos (\omega _\alpha
1388 > t)\dot x(t - \tau )d} \tau }  + \sum\limits_{\alpha  = 1}^N {\left\{
1389 > {\left[ {g_\alpha  x_\alpha  (0) - \frac{{g_\alpha  }}{{m_\alpha
1390 > \omega _\alpha  }}} \right]\cos (\omega _\alpha  t) +
1391 > \frac{{g_\alpha  \dot x_\alpha  (0)}}{{\omega _\alpha  }}\sin
1392 > (\omega _\alpha  t)} \right\}}
1393 > \end{align}
1394 >
1395 > \begin{equation}
1396 > m\ddot x =  - \frac{{\partial W}}{{\partial x}} - \int_0^t {\xi
1397 > (t)\dot x(t - \tau )d\tau }  + R(t)
1398 > \label{introEuqation:GeneralizedLangevinDynamics}
1399 > \end{equation}
1400 > %where $ {\xi (t)}$ is friction kernel, $R(t)$ is random force and
1401 > %$W$ is the potential of mean force. $W(x) =  - kT\ln p(x)$
1402 > \[
1403 > \xi (t) = \sum\limits_{\alpha  = 1}^N {\left( { - \frac{{g_\alpha ^2
1404 > }}{{m_\alpha  \omega _\alpha ^2 }}} \right)\cos (\omega _\alpha  t)}
1405 > \]
1406 > For an infinite harmonic bath, we can use the spectral density and
1407 > an integral over frequencies.
1408 >
1409 > \[
1410 > R(t) = \sum\limits_{\alpha  = 1}^N {\left( {g_\alpha  x_\alpha  (0)
1411 > - \frac{{g_\alpha ^2 }}{{m_\alpha  \omega _\alpha ^2 }}x(0)}
1412 > \right)\cos (\omega _\alpha  t)}  + \frac{{\dot x_\alpha
1413 > (0)}}{{\omega _\alpha  }}\sin (\omega _\alpha  t)
1414 > \]
1415 > The random forces depend only on initial conditions.
1416 >
1417 > \subsubsection{\label{introSection:secondFluctuationDissipation}The Second Fluctuation Dissipation Theorem}
1418 > So we can define a new set of coordinates,
1419 > \[
1420 > q_\alpha  (t) = x_\alpha  (t) - \frac{1}{{m_\alpha  \omega _\alpha
1421 > ^2 }}x(0)
1422 > \]
1423 > This makes
1424 > \[
1425 > R(t) = \sum\limits_{\alpha  = 1}^N {g_\alpha  q_\alpha  (t)}
1426 > \]
1427 > And since the $q$ coordinates are harmonic oscillators,
1428 > \[
1429 > \begin{array}{l}
1430 > \left\langle {q_\alpha  (t)q_\alpha  (0)} \right\rangle  = \left\langle {q_\alpha ^2 (0)} \right\rangle \cos (\omega _\alpha  t) \\
1431 > \left\langle {q_\alpha  (t)q_\beta  (0)} \right\rangle  = \delta _{\alpha \beta } \left\langle {q_\alpha  (t)q_\alpha  (0)} \right\rangle  \\
1432 > \end{array}
1433 > \]
1434 >
1435 > \begin{align}
1436 > \left\langle {R(t)R(0)} \right\rangle  &= \sum\limits_\alpha
1437 > {\sum\limits_\beta  {g_\alpha  g_\beta  \left\langle {q_\alpha
1438 > (t)q_\beta  (0)} \right\rangle } }
1439 > %
1440 > &= \sum\limits_\alpha  {g_\alpha ^2 \left\langle {q_\alpha ^2 (0)}
1441 > \right\rangle \cos (\omega _\alpha  t)}
1442 > %
1443 > &= kT\xi (t)
1444 > \end{align}
1445 >
1446 > \begin{equation}
1447 > \xi (t) = \left\langle {R(t)R(0)} \right\rangle
1448 > \label{introEquation:secondFluctuationDissipation}
1449 > \end{equation}
1450 >
1451 > \subsection{\label{introSection:frictionTensor} Friction Tensor}
1452 > Theoretically, the friction kernel can be determined using velocity
1453 > autocorrelation function. However, this approach become impractical
1454 > when the system become more and more complicate. Instead, various
1455 > approaches based on hydrodynamics have been developed to calculate
1456 > the friction coefficients. The friction effect is isotropic in
1457 > Equation, \zeta can be taken as a scalar. In general, friction
1458 > tensor \Xi is a $6\times 6$ matrix given by
1459 > \[
1460 > \Xi  = \left( {\begin{array}{*{20}c}
1461 >   {\Xi _{}^{tt} } & {\Xi _{}^{rt} }  \\
1462 >   {\Xi _{}^{tr} } & {\Xi _{}^{rr} }  \\
1463 > \end{array}} \right).
1464 > \]
1465 > Here, $ {\Xi^{tt} }$ and $ {\Xi^{rr} }$ are translational friction
1466 > tensor and rotational friction tensor respectively, while ${\Xi^{tr}
1467 > }$ is translation-rotation coupling tensor and $ {\Xi^{rt} }$ is
1468 > rotation-translation coupling tensor.
1469 >
1470 > \[
1471 > \left( \begin{array}{l}
1472 > F_t  \\
1473 > \tau  \\
1474 > \end{array} \right) =  - \left( {\begin{array}{*{20}c}
1475 >   {\Xi ^{tt} } & {\Xi ^{rt} }  \\
1476 >   {\Xi ^{tr} } & {\Xi ^{rr} }  \\
1477 > \end{array}} \right)\left( \begin{array}{l}
1478 > v \\
1479 > w \\
1480 > \end{array} \right)
1481 > \]
1482 >
1483 > \subsubsection{\label{introSection:analyticalApproach}The Friction Tensor for Regular Shape}
1484 > For a spherical particle, the translational and rotational friction
1485 > constant can be calculated from Stoke's law,
1486 > \[
1487 > \Xi ^{tt}  = \left( {\begin{array}{*{20}c}
1488 >   {6\pi \eta R} & 0 & 0  \\
1489 >   0 & {6\pi \eta R} & 0  \\
1490 >   0 & 0 & {6\pi \eta R}  \\
1491 > \end{array}} \right)
1492 > \]
1493 > and
1494 > \[
1495 > \Xi ^{rr}  = \left( {\begin{array}{*{20}c}
1496 >   {8\pi \eta R^3 } & 0 & 0  \\
1497 >   0 & {8\pi \eta R^3 } & 0  \\
1498 >   0 & 0 & {8\pi \eta R^3 }  \\
1499 > \end{array}} \right)
1500 > \]
1501 > where $\eta$ is the viscosity of the solvent and $R$ is the
1502 > hydrodynamics radius.
1503 >
1504 > Other non-spherical particles have more complex properties.
1505 >
1506 > \[
1507 > S = \frac{2}{{\sqrt {a^2  - b^2 } }}\ln \frac{{a + \sqrt {a^2  - b^2
1508 > } }}{b}
1509 > \]
1510 >
1511 >
1512 > \[
1513 > S = \frac{2}{{\sqrt {b^2  - a^2 } }}arctg\frac{{\sqrt {b^2  - a^2 }
1514 > }}{a}
1515 > \]
1516 >
1517 > \[
1518 > \begin{array}{l}
1519 > \Xi _a^{tt}  = 16\pi \eta \frac{{a^2  - b^2 }}{{(2a^2  - b^2 )S - 2a}} \\
1520 > \Xi _b^{tt}  = \Xi _c^{tt}  = 32\pi \eta \frac{{a^2  - b^2 }}{{(2a^2  - 3b^2 )S + 2a}} \\
1521 > \end{array}
1522 > \]
1523 >
1524 > \[
1525 > \begin{array}{l}
1526 > \Xi _a^{rr}  = \frac{{32\pi }}{3}\eta \frac{{(a^2  - b^2 )b^2 }}{{2a - b^2 S}} \\
1527 > \Xi _b^{rr}  = \Xi _c^{rr}  = \frac{{32\pi }}{3}\eta \frac{{(a^4  - b^4 )}}{{(2a^2  - b^2 )S - 2a}} \\
1528 > \end{array}
1529 > \]
1530 >
1531 >
1532 > \subsubsection{\label{introSection:approximationApproach}The Friction Tensor for Arbitrary Shape}
1533 > Unlike spherical and other regular shaped molecules, there is not
1534 > analytical solution for friction tensor of any arbitrary shaped
1535 > rigid molecules. The ellipsoid of revolution model and general
1536 > triaxial ellipsoid model have been used to approximate the
1537 > hydrodynamic properties of rigid bodies. However, since the mapping
1538 > from all possible ellipsoidal space, $r$-space, to all possible
1539 > combination of rotational diffusion coefficients, $D$-space is not
1540 > unique\cite{Wegener79} as well as the intrinsic coupling between
1541 > translational and rotational motion of rigid body\cite{}, general
1542 > ellipsoid is not always suitable for modeling arbitrarily shaped
1543 > rigid molecule. A number of studies have been devoted to determine
1544 > the friction tensor for irregularly shaped rigid bodies using more
1545 > advanced method\cite{} where the molecule of interest was modeled by
1546 > combinations of spheres(beads)\cite{} and the hydrodynamics
1547 > properties of the molecule can be calculated using the hydrodynamic
1548 > interaction tensor. Let us consider a rigid assembly of $N$ beads
1549 > immersed in a continuous medium. Due to hydrodynamics interaction,
1550 > the ``net'' velocity of $i$th bead, $v'_i$ is different than its
1551 > unperturbed velocity $v_i$,
1552 > \[
1553 > v'_i  = v_i  - \sum\limits_{j \ne i} {T_{ij} F_j }
1554 > \]
1555 > where $F_i$ is the frictional force, and $T_{ij}$ is the
1556 > hydrodynamic interaction tensor. The friction force of $i$th bead is
1557 > proportional to its ``net'' velocity
1558 > \begin{equation}
1559 > F_i  = \zeta _i v_i  - \zeta _i \sum\limits_{j \ne i} {T_{ij} F_j }.
1560 > \label{introEquation:tensorExpression}
1561 > \end{equation}
1562 > This equation is the basis for deriving the hydrodynamic tensor. In
1563 > 1930, Oseen and Burgers gave a simple solution to Equation
1564 > \ref{introEquation:tensorExpression}
1565 > \begin{equation}
1566 > T_{ij}  = \frac{1}{{8\pi \eta r_{ij} }}\left( {I + \frac{{R_{ij}
1567 > R_{ij}^T }}{{R_{ij}^2 }}} \right).
1568 > \label{introEquation:oseenTensor}
1569 > \end{equation}
1570 > Here $R_{ij}$ is the distance vector between bead $i$ and bead $j$.
1571 > A second order expression for element of different size was
1572 > introduced by Rotne and Prager\cite{} and improved by Garc\'{i}a de
1573 > la Torre and Bloomfield,
1574 > \begin{equation}
1575 > T_{ij}  = \frac{1}{{8\pi \eta R_{ij} }}\left[ {\left( {I +
1576 > \frac{{R_{ij} R_{ij}^T }}{{R_{ij}^2 }}} \right) + R\frac{{\sigma
1577 > _i^2  + \sigma _j^2 }}{{r_{ij}^2 }}\left( {\frac{I}{3} -
1578 > \frac{{R_{ij} R_{ij}^T }}{{R_{ij}^2 }}} \right)} \right].
1579 > \label{introEquation:RPTensorNonOverlapped}
1580 > \end{equation}
1581 > Both of the Equation \ref{introEquation:oseenTensor} and Equation
1582 > \ref{introEquation:RPTensorNonOverlapped} have an assumption $R_{ij}
1583 > \ge \sigma _i  + \sigma _j$. An alternative expression for
1584 > overlapping beads with the same radius, $\sigma$, is given by
1585 > \begin{equation}
1586 > T_{ij}  = \frac{1}{{6\pi \eta R_{ij} }}\left[ {\left( {1 -
1587 > \frac{2}{{32}}\frac{{R_{ij} }}{\sigma }} \right)I +
1588 > \frac{2}{{32}}\frac{{R_{ij} R_{ij}^T }}{{R_{ij} \sigma }}} \right]
1589 > \label{introEquation:RPTensorOverlapped}
1590 > \end{equation}
1591 >
1592 > %Bead Modeling
1593 >
1594 > \[
1595 > B = \left( {\begin{array}{*{20}c}
1596 >   {T_{11} } &  \ldots  & {T_{1N} }  \\
1597 >    \vdots  &  \ddots  &  \vdots   \\
1598 >   {T_{N1} } &  \cdots  & {T_{NN} }  \\
1599 > \end{array}} \right)
1600 > \]
1601 >
1602 > \[
1603 > C = B^{ - 1}  = \left( {\begin{array}{*{20}c}
1604 >   {C_{11} } &  \ldots  & {C_{1N} }  \\
1605 >    \vdots  &  \ddots  &  \vdots   \\
1606 >   {C_{N1} } &  \cdots  & {C_{NN} }  \\
1607 > \end{array}} \right)
1608 > \]
1609 >
1610 > \begin{equation}
1611 > \begin{array}{l}
1612 > \Xi _{}^{tt}  = \sum\limits_i {\sum\limits_j {C_{ij} } } , \\
1613 > \Xi _{}^{tr}  = \Xi _{}^{rt}  = \sum\limits_i {\sum\limits_j {U_i C_{ij} } } , \\
1614 > \Xi _{}^{rr}  =  - \sum\limits_i {\sum\limits_j {U_i C_{ij} } } U_j  \\
1615 > \end{array}
1616 > \end{equation}
1617 > where
1618 > \[
1619 > U_i  = \left( {\begin{array}{*{20}c}
1620 >   0 & { - z_i } & {y_i }  \\
1621 >   {z_i } & 0 & { - x_i }  \\
1622 >   { - y_i } & {x_i } & 0  \\
1623 > \end{array}} \right)
1624 > \]
1625 >
1626 > \[
1627 > r_{OR}  = \left( \begin{array}{l}
1628 > x_{OR}  \\
1629 > y_{OR}  \\
1630 > z_{OR}  \\
1631 > \end{array} \right) = \left( {\begin{array}{*{20}c}
1632 >   {\Xi _{yy}^{rr}  + \Xi _{zz}^{rr} } & { - \Xi _{xy}^{rr} } & { - \Xi _{xz}^{rr} }  \\
1633 >   { - \Xi _{yx}^{rr} } & {\Xi _{zz}^{rr}  + \Xi _{xx}^{rr} } & { - \Xi _{yz}^{rr} }  \\
1634 >   { - \Xi _{zx}^{rr} } & { - \Xi _{yz}^{rr} } & {\Xi _{xx}^{rr}  + \Xi _{yy}^{rr} }  \\
1635 > \end{array}} \right)^{ - 1} \left( \begin{array}{l}
1636 > \Xi _{yz}^{tr}  - \Xi _{zy}^{tr}  \\
1637 > \Xi _{zx}^{tr}  - \Xi _{xz}^{tr}  \\
1638 > \Xi _{xy}^{tr}  - \Xi _{yx}^{tr}  \\
1639 > \end{array} \right)
1640 > \]
1641 >
1642 > \[
1643 > U_{OR}  = \left( {\begin{array}{*{20}c}
1644 >   0 & { - z_{OR} } & {y_{OR} }  \\
1645 >   {z_i } & 0 & { - x_{OR} }  \\
1646 >   { - y_{OR} } & {x_{OR} } & 0  \\
1647 > \end{array}} \right)
1648 > \]
1649 >
1650 > \[
1651 > \begin{array}{l}
1652 > \Xi _R^{tt}  = \Xi _{}^{tt}  \\
1653 > \Xi _R^{tr}  = \Xi _R^{rt}  = \Xi _{}^{tr}  - U_{OR} \Xi _{}^{tt}  \\
1654 > \Xi _R^{rr}  = \Xi _{}^{rr}  - U_{OR} \Xi _{}^{tt} U_{OR}  + \Xi _{}^{tr} U_{OR}  - U_{OR} \Xi _{}^{tr} ^{^T }  \\
1655 > \end{array}
1656 > \]
1657 >
1658 > \[
1659 > D_R  = \left( {\begin{array}{*{20}c}
1660 >   {D_R^{tt} } & {D_R^{rt} }  \\
1661 >   {D_R^{tr} } & {D_R^{rr} }  \\
1662 > \end{array}} \right) = k_b T\left( {\begin{array}{*{20}c}
1663 >   {\Xi _R^{tt} } & {\Xi _R^{rt} }  \\
1664 >   {\Xi _R^{tr} } & {\Xi _R^{rr} }  \\
1665 > \end{array}} \right)^{ - 1}
1666 > \]
1667 >
1668 >
1669 > %Approximation Methods
1670 >
1671 > %\section{\label{introSection:correlationFunctions}Correlation Functions}

Diff Legend

Removed lines
+ Added lines
< Changed lines
> Changed lines