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 2699 by tim, Mon Apr 10 05:35:55 2006 UTC vs.
Revision 2776 by tim, Thu May 25 21:32:14 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 221 | Line 221 | Statistical Mechanics concepts presented in this disse
221   The thermodynamic behaviors and properties of Molecular Dynamics
222   simulation are governed by the principle of Statistical Mechanics.
223   The following section will give a brief introduction to some of the
224 < Statistical Mechanics concepts presented in this dissertation.
224 > Statistical Mechanics concepts and theorem presented in this
225 > dissertation.
226  
227 < \subsection{\label{introSection:ensemble}Ensemble and Phase Space}
227 > \subsection{\label{introSection:ensemble}Phase Space and Ensemble}
228  
229 + Mathematically, phase space is the space which represents all
230 + possible states. Each possible state of the system corresponds to
231 + one unique point in the phase space. For mechanical systems, the
232 + phase space usually consists of all possible values of position and
233 + momentum variables. Consider a dynamic system in a cartesian space,
234 + where each of the $6f$ coordinates and momenta is assigned to one of
235 + $6f$ mutually orthogonal axes, the phase space of this system is a
236 + $6f$ dimensional space. A point, $x = (q_1 , \ldots ,q_f ,p_1 ,
237 + \ldots ,p_f )$, with a unique set of values of $6f$ coordinates and
238 + momenta is a phase space vector.
239 +
240 + A microscopic state or microstate of a classical system is
241 + specification of the complete phase space vector of a system at any
242 + instant in time. An ensemble is defined as a collection of systems
243 + sharing one or more macroscopic characteristics but each being in a
244 + unique microstate. The complete ensemble is specified by giving all
245 + systems or microstates consistent with the common macroscopic
246 + characteristics of the ensemble. Although the state of each
247 + individual system in the ensemble could be precisely described at
248 + any instance in time by a suitable phase space vector, when using
249 + ensembles for statistical purposes, there is no need to maintain
250 + distinctions between individual systems, since the numbers of
251 + systems at any time in the different states which correspond to
252 + different regions of the phase space are more interesting. Moreover,
253 + in the point of view of statistical mechanics, one would prefer to
254 + use ensembles containing a large enough population of separate
255 + members so that the numbers of systems in such different states can
256 + be regarded as changing continuously as we traverse different
257 + regions of the phase space. The condition of an ensemble at any time
258 + can be regarded as appropriately specified by the density $\rho$
259 + with which representative points are distributed over the phase
260 + space. The density of distribution for an ensemble with $f$ degrees
261 + of freedom is defined as,
262 + \begin{equation}
263 + \rho  = \rho (q_1 , \ldots ,q_f ,p_1 , \ldots ,p_f ,t).
264 + \label{introEquation:densityDistribution}
265 + \end{equation}
266 + Governed by the principles of mechanics, the phase points change
267 + their value which would change the density at any time at phase
268 + space. Hence, the density of distribution is also to be taken as a
269 + function of the time.
270 +
271 + The number of systems $\delta N$ at time $t$ can be determined by,
272 + \begin{equation}
273 + \delta N = \rho (q,p,t)dq_1  \ldots dq_f dp_1  \ldots dp_f.
274 + \label{introEquation:deltaN}
275 + \end{equation}
276 + Assuming a large enough population of systems are exploited, we can
277 + sufficiently approximate $\delta N$ without introducing
278 + discontinuity when we go from one region in the phase space to
279 + another. By integrating over the whole phase space,
280 + \begin{equation}
281 + N = \int { \ldots \int {\rho (q,p,t)dq_1 } ...dq_f dp_1 } ...dp_f
282 + \label{introEquation:totalNumberSystem}
283 + \end{equation}
284 + gives us an expression for the total number of the systems. Hence,
285 + the probability per unit in the phase space can be obtained by,
286 + \begin{equation}
287 + \frac{{\rho (q,p,t)}}{N} = \frac{{\rho (q,p,t)}}{{\int { \ldots \int
288 + {\rho (q,p,t)dq_1 } ...dq_f dp_1 } ...dp_f }}.
289 + \label{introEquation:unitProbability}
290 + \end{equation}
291 + With the help of Equation(\ref{introEquation:unitProbability}) and
292 + the knowledge of the system, it is possible to calculate the average
293 + value of any desired quantity which depends on the coordinates and
294 + momenta of the system. Even when the dynamics of the real system is
295 + complex, or stochastic, or even discontinuous, the average
296 + properties of the ensemble of possibilities as a whole may still
297 + remain well defined. For a classical system in thermal equilibrium
298 + with its environment, the ensemble average of a mechanical quantity,
299 + $\langle A(q , p) \rangle_t$, takes the form of an integral over the
300 + phase space of the system,
301 + \begin{equation}
302 + \langle  A(q , p) \rangle_t = \frac{{\int { \ldots \int {A(q,p)\rho
303 + (q,p,t)dq_1 } ...dq_f dp_1 } ...dp_f }}{{\int { \ldots \int {\rho
304 + (q,p,t)dq_1 } ...dq_f dp_1 } ...dp_f }}
305 + \label{introEquation:ensembelAverage}
306 + \end{equation}
307 +
308 + There are several different types of ensembles with different
309 + statistical characteristics. As a function of macroscopic
310 + parameters, such as temperature \textit{etc}, partition function can
311 + be used to describe the statistical properties of a system in
312 + thermodynamic equilibrium.
313 +
314 + As an ensemble of systems, each of which is known to be thermally
315 + isolated and conserve energy, Microcanonical ensemble(NVE) has a
316 + partition function like,
317 + \begin{equation}
318 + \Omega (N,V,E) = e^{\beta TS} \label{introEquation:NVEPartition}.
319 + \end{equation}
320 + A canonical ensemble(NVT)is an ensemble of systems, each of which
321 + can share its energy with a large heat reservoir. The distribution
322 + of the total energy amongst the possible dynamical states is given
323 + by the partition function,
324 + \begin{equation}
325 + \Omega (N,V,T) = e^{ - \beta A}
326 + \label{introEquation:NVTPartition}
327 + \end{equation}
328 + Here, $A$ is the Helmholtz free energy which is defined as $ A = U -
329 + TS$. Since most experiment are carried out under constant pressure
330 + condition, isothermal-isobaric ensemble(NPT) play a very important
331 + role in molecular simulation. The isothermal-isobaric ensemble allow
332 + the system to exchange energy with a heat bath of temperature $T$
333 + and to change the volume as well. Its partition function is given as
334 + \begin{equation}
335 + \Delta (N,P,T) =  - e^{\beta G}.
336 + \label{introEquation:NPTPartition}
337 + \end{equation}
338 + Here, $G = U - TS + PV$, and $G$ is called Gibbs free energy.
339 +
340 + \subsection{\label{introSection:liouville}Liouville's theorem}
341 +
342 + The Liouville's theorem is the foundation on which statistical
343 + mechanics rests. It describes the time evolution of phase space
344 + distribution function. In order to calculate the rate of change of
345 + $\rho$, we begin from Equation(\ref{introEquation:deltaN}). If we
346 + consider the two faces perpendicular to the $q_1$ axis, which are
347 + located at $q_1$ and $q_1 + \delta q_1$, the number of phase points
348 + leaving the opposite face is given by the expression,
349 + \begin{equation}
350 + \left( {\rho  + \frac{{\partial \rho }}{{\partial q_1 }}\delta q_1 }
351 + \right)\left( {\dot q_1  + \frac{{\partial \dot q_1 }}{{\partial q_1
352 + }}\delta q_1 } \right)\delta q_2  \ldots \delta q_f \delta p_1
353 + \ldots \delta p_f .
354 + \end{equation}
355 + Summing all over the phase space, we obtain
356 + \begin{equation}
357 + \frac{{d(\delta N)}}{{dt}} =  - \sum\limits_{i = 1}^f {\left[ {\rho
358 + \left( {\frac{{\partial \dot q_i }}{{\partial q_i }} +
359 + \frac{{\partial \dot p_i }}{{\partial p_i }}} \right) + \left(
360 + {\frac{{\partial \rho }}{{\partial q_i }}\dot q_i  + \frac{{\partial
361 + \rho }}{{\partial p_i }}\dot p_i } \right)} \right]} \delta q_1
362 + \ldots \delta q_f \delta p_1  \ldots \delta p_f .
363 + \end{equation}
364 + Differentiating the equations of motion in Hamiltonian formalism
365 + (\ref{introEquation:motionHamiltonianCoordinate},
366 + \ref{introEquation:motionHamiltonianMomentum}), we can show,
367 + \begin{equation}
368 + \sum\limits_i {\left( {\frac{{\partial \dot q_i }}{{\partial q_i }}
369 + + \frac{{\partial \dot p_i }}{{\partial p_i }}} \right)}  = 0 ,
370 + \end{equation}
371 + which cancels the first terms of the right hand side. Furthermore,
372 + divining $ \delta q_1  \ldots \delta q_f \delta p_1  \ldots \delta
373 + p_f $ in both sides, we can write out Liouville's theorem in a
374 + simple form,
375 + \begin{equation}
376 + \frac{{\partial \rho }}{{\partial t}} + \sum\limits_{i = 1}^f
377 + {\left( {\frac{{\partial \rho }}{{\partial q_i }}\dot q_i  +
378 + \frac{{\partial \rho }}{{\partial p_i }}\dot p_i } \right)}  = 0 .
379 + \label{introEquation:liouvilleTheorem}
380 + \end{equation}
381 +
382 + Liouville's theorem states that the distribution function is
383 + constant along any trajectory in phase space. In classical
384 + statistical mechanics, since the number of particles in the system
385 + is huge, we may be able to believe the system is stationary,
386 + \begin{equation}
387 + \frac{{\partial \rho }}{{\partial t}} = 0.
388 + \label{introEquation:stationary}
389 + \end{equation}
390 + In such stationary system, the density of distribution $\rho$ can be
391 + connected to the Hamiltonian $H$ through Maxwell-Boltzmann
392 + distribution,
393 + \begin{equation}
394 + \rho  \propto e^{ - \beta H}
395 + \label{introEquation:densityAndHamiltonian}
396 + \end{equation}
397 +
398 + \subsubsection{\label{introSection:phaseSpaceConservation}Conservation of Phase Space}
399 + Lets consider a region in the phase space,
400 + \begin{equation}
401 + \delta v = \int { \ldots \int {dq_1 } ...dq_f dp_1 } ..dp_f .
402 + \end{equation}
403 + If this region is small enough, the density $\rho$ can be regarded
404 + as uniform over the whole phase space. Thus, the number of phase
405 + points inside this region is given by,
406 + \begin{equation}
407 + \delta N = \rho \delta v = \rho \int { \ldots \int {dq_1 } ...dq_f
408 + dp_1 } ..dp_f.
409 + \end{equation}
410 +
411 + \begin{equation}
412 + \frac{{d(\delta N)}}{{dt}} = \frac{{d\rho }}{{dt}}\delta v + \rho
413 + \frac{d}{{dt}}(\delta v) = 0.
414 + \end{equation}
415 + With the help of stationary assumption
416 + (\ref{introEquation:stationary}), we obtain the principle of the
417 + \emph{conservation of extension in phase space},
418 + \begin{equation}
419 + \frac{d}{{dt}}(\delta v) = \frac{d}{{dt}}\int { \ldots \int {dq_1 }
420 + ...dq_f dp_1 } ..dp_f  = 0.
421 + \label{introEquation:volumePreserving}
422 + \end{equation}
423 +
424 + \subsubsection{\label{introSection:liouvilleInOtherForms}Liouville's Theorem in Other Forms}
425 +
426 + Liouville's theorem can be expresses in a variety of different forms
427 + which are convenient within different contexts. For any two function
428 + $F$ and $G$ of the coordinates and momenta of a system, the Poisson
429 + bracket ${F, G}$ is defined as
430 + \begin{equation}
431 + \left\{ {F,G} \right\} = \sum\limits_i {\left( {\frac{{\partial
432 + F}}{{\partial q_i }}\frac{{\partial G}}{{\partial p_i }} -
433 + \frac{{\partial F}}{{\partial p_i }}\frac{{\partial G}}{{\partial
434 + q_i }}} \right)}.
435 + \label{introEquation:poissonBracket}
436 + \end{equation}
437 + Substituting equations of motion in Hamiltonian formalism(
438 + \ref{introEquation:motionHamiltonianCoordinate} ,
439 + \ref{introEquation:motionHamiltonianMomentum} ) into
440 + (\ref{introEquation:liouvilleTheorem}), we can rewrite Liouville's
441 + theorem using Poisson bracket notion,
442 + \begin{equation}
443 + \left( {\frac{{\partial \rho }}{{\partial t}}} \right) =  - \left\{
444 + {\rho ,H} \right\}.
445 + \label{introEquation:liouvilleTheromInPoissin}
446 + \end{equation}
447 + Moreover, the Liouville operator is defined as
448 + \begin{equation}
449 + iL = \sum\limits_{i = 1}^f {\left( {\frac{{\partial H}}{{\partial
450 + p_i }}\frac{\partial }{{\partial q_i }} - \frac{{\partial
451 + H}}{{\partial q_i }}\frac{\partial }{{\partial p_i }}} \right)}
452 + \label{introEquation:liouvilleOperator}
453 + \end{equation}
454 + In terms of Liouville operator, Liouville's equation can also be
455 + expressed as
456 + \begin{equation}
457 + \left( {\frac{{\partial \rho }}{{\partial t}}} \right) =  - iL\rho
458 + \label{introEquation:liouvilleTheoremInOperator}
459 + \end{equation}
460 +
461   \subsection{\label{introSection:ergodic}The Ergodic Hypothesis}
462  
463   Various thermodynamic properties can be calculated from Molecular
# Line 239 | 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 336 | Line 570 | The free rigid body is an example of Poisson system (a
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 $x$.
339 The free rigid body is an example of Poisson system (actually a
340 Lie-Poisson system) with Hamiltonian function of angular kinetic
341 energy.
342 \begin{equation}
343 J(\pi ) = \left( {\begin{array}{*{20}c}
344   0 & {\pi _3 } & { - \pi _2 }  \\
345   { - \pi _3 } & 0 & {\pi _1 }  \\
346   {\pi _2 } & { - \pi _1 } & 0  \\
347 \end{array}} \right)
348 \end{equation}
573  
574 < \begin{equation}
351 < H = \frac{1}{2}\left( {\frac{{\pi _1^2 }}{{I_1 }} + \frac{{\pi _2^2
352 < }}{{I_2 }} + \frac{{\pi _3^2 }}{{I_3 }}} \right)
353 < \end{equation}
574 > \subsection{\label{introSection:exactFlow}Exact Flow}
575  
355 \subsection{\label{introSection:geometricProperties}Geometric Properties}
576   Let $x(t)$ be the exact solution of the ODE system,
577   \begin{equation}
578   \frac{{dx}}{{dt}} = f(x) \label{introEquation:ODE}
# Line 362 | Line 582 | space to itself. In most cases, it is not easy to find
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. In most cases, it is not easy to find the exact
366 < flow $\varphi_\tau$. Instead, we use a approximate map, $\psi_\tau$,
367 < which is usually called integrator. The order of an integrator
368 < $\psi_\tau$ is $p$, if the Taylor series of $\psi_\tau$ agree to
369 < order $p$,
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. Let $\varphi$ be the flow of Hamiltonian
617 < vector field, $\varphi$ is a \emph{symplectic} flow if it satisfies,
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.
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
# Line 383 | Line 627 | symplectomorphism. As to the Poisson system,
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
630 > {\varphi '}^T J \varphi ' = J \circ \varphi
631   \end{equation}
632 < is the property must be preserved by the integrator. It is possible
633 < to construct a \emph{volume-preserving} flow for a source free($
634 < \nabla \cdot f = 0 $) ODE, if the flow satisfies $ \det d\varphi  =
635 < 1$. Changing the variables $y = h(x)$ in a
636 < ODE\ref{introEquation:ODE} will result in a new system,
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 $. When
399 < designing any numerical methods, one should always try to preserve
400 < the structural properties of the original ODE and its flow.
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
# Line 414 | Line 698 | and difficult to use\cite{}. In dissipative systems, v
698   \end{enumerate}
699  
700   Generating function tends to lead to methods which are cumbersome
701 < and difficult to use\cite{}. In dissipative systems, variational
702 < methods can capture the decay of energy accurately\cite{}. Since
703 < their geometrically unstable nature against non-Hamiltonian
704 < perturbations, ordinary implicit Runge-Kutta methods are not
705 < suitable for Hamiltonian system. Recently, various high-order
706 < explicit Runge--Kutta methods have been developed to overcome this
707 < instability \cite{}. However, due to computational penalty involved
708 < in implementing the Runge-Kutta methods, they do not attract too
709 < much attention from Molecular Dynamics community. Instead, splitting
710 < have been widely accepted since they exploit natural decompositions
711 < of the system\cite{Tuckerman92}. The main idea behind splitting
712 < methods is to decompose the discrete $\varphi_h$ as a composition of
713 < simpler flows,
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. Let $\phi$ and $\psi$ both be
724 < symplectic maps, it is easy to show that any composition of
725 < symplectic flows yields a symplectic map,
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.
747 > '\phi ' = \phi '^T J\phi ' = J,
748   \label{introEquation:SymplecticFlowComposition}
749   \end{equation}
750 < Suppose that a Hamiltonian system has a form with $H = T + V$
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: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 + \[
845 + [X,Y] = XY - YX .
846 + \]
847 + Applying Baker-Campbell-Hausdorff formula to Sprang splitting, we
848 + can obtain
849 + \begin{equation}
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{equation}
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
887 < has proven to be a powerful tool for studying the functions of
888 < biological systems, providing structural, thermodynamic and
889 < dynamical information.
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 > \[
895 > T = \sum\limits_{i = 1}^N {\frac{{m_i v_i^2 }}{{fk_B }}}
896 > \]
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 < \subsection{\label{introSec:mdInit}Initialization}
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. Sec.~\ref{introSection:Analysis}
917 > provides the theoretical tools for trajectory analysis.
918  
919 < \subsection{\label{introSection:mdIntegration} Integration of the Equations of Motion}
919 > \subsection{\label{introSec:initialSystemSettings}Initialization}
920  
921 < \section{\label{introSection:rigidBody}Dynamics of Rigid Bodies}
921 > \subsubsection{Preliminary preparation}
922  
923 < A rigid body is a body in which the distance between any two given
924 < points of a rigid body remains constant regardless of external
925 < forces exerted on it. A rigid body therefore conserves its shape
926 < during its motion.
923 > When selecting the starting structure of a molecule for molecular
924 > simulation, one may retrieve its Cartesian coordinates from public
925 > databases, such as RCSB Protein Data Bank \textit{etc}. Although
926 > thousands of crystal structures of molecules are discovered every
927 > year, many more remain unknown due to the difficulties of
928 > purification and crystallization. Even for the molecule with known
929 > structure, some important information is missing. For example, the
930 > missing hydrogen atom which acts as donor in hydrogen bonding must
931 > be added. Moreover, in order to include electrostatic interaction,
932 > one may need to specify the partial charges for individual atoms.
933 > Under some circumstances, we may even need to prepare the system in
934 > a special setup. For instance, when studying transport phenomenon in
935 > membrane system, we may prepare the lipids in bilayer structure
936 > instead of placing lipids randomly in solvent, since we are not
937 > interested in self-aggregation and it takes a long time to happen.
938  
939 < Applications of dynamics of rigid bodies.
939 > \subsubsection{Minimization}
940  
941 < \subsection{\label{introSection:lieAlgebra}Lie Algebra}
941 > It is quite possible that some of molecules in the system from
942 > preliminary preparation may be overlapped with each other. This
943 > close proximity leads to high potential energy which consequently
944 > jeopardizes any molecular dynamics simulations. To remove these
945 > steric overlaps, one typically performs energy minimization to find
946 > a more reasonable conformation. Several energy minimization methods
947 > have been developed to exploit the energy surface and to locate the
948 > local minimum. While converging slowly near the minimum, steepest
949 > descent method is extremely robust when systems are far from
950 > harmonic. Thus, it is often used to refine structure from
951 > crystallographic data. Relied on the gradient or hessian, advanced
952 > methods like conjugate gradient and Newton-Raphson converge rapidly
953 > to a local minimum, while become unstable if the energy surface is
954 > far from quadratic. Another factor must be taken into account, when
955 > choosing energy minimization method, is the size of the system.
956 > Steepest descent and conjugate gradient can deal with models of any
957 > size. Because of the limit of computation power to calculate hessian
958 > matrix and insufficient storage capacity to store them, most
959 > Newton-Raphson methods can not be used with very large models.
960  
961 < \subsection{\label{introSection:DLMMotionEquation}The Euler Equations of Rigid Body Motion}
961 > \subsubsection{Heating}
962  
963 < \subsection{\label{introSection:otherRBMotionEquation}Other Formulations for Rigid Body Motion}
963 > Typically, Heating is performed by assigning random velocities
964 > according to a Gaussian distribution for a temperature. Beginning at
965 > a lower temperature and gradually increasing the temperature by
966 > assigning greater random velocities, we end up with setting the
967 > temperature of the system to a final temperature at which the
968 > simulation will be conducted. In heating phase, we should also keep
969 > the system from drifting or rotating as a whole. Equivalently, the
970 > net linear momentum and angular momentum of the system should be
971 > shifted to zero.
972  
973 < %\subsection{\label{introSection:poissonBrackets}Poisson Brackets}
973 > \subsubsection{Equilibration}
974  
975 < \section{\label{introSection:correlationFunctions}Correlation Functions}
975 > The purpose of equilibration is to allow the system to evolve
976 > spontaneously for a period of time and reach equilibrium. The
977 > procedure is continued until various statistical properties, such as
978 > temperature, pressure, energy, volume and other structural
979 > properties \textit{etc}, become independent of time. Strictly
980 > speaking, minimization and heating are not necessary, provided the
981 > equilibration process is long enough. However, these steps can serve
982 > as a means to arrive at an equilibrated structure in an effective
983 > way.
984  
985 < \section{\label{introSection:langevinDynamics}Langevin Dynamics}
985 > \subsection{\label{introSection:production}Production}
986  
987 < \subsection{\label{introSection:LDIntroduction}Introduction and application of Langevin Dynamics}
987 > Production run is the most important steps of the simulation, in
988 > which the equilibrated structure is used as a starting point and the
989 > motions of the molecules are collected for later analysis. In order
990 > to capture the macroscopic properties of the system, the molecular
991 > dynamics simulation must be performed in correct and efficient way.
992  
993 < \subsection{\label{introSection:generalizedLangevinDynamics}Generalized Langevin Dynamics}
993 > The most expensive part of a molecular dynamics simulation is the
994 > calculation of non-bonded forces, such as van der Waals force and
995 > Coulombic forces \textit{etc}. For a system of $N$ particles, the
996 > complexity of the algorithm for pair-wise interactions is $O(N^2 )$,
997 > which making large simulations prohibitive in the absence of any
998 > computation saving techniques.
999  
1000 + A natural approach to avoid system size issue is to represent the
1001 + bulk behavior by a finite number of the particles. However, this
1002 + approach will suffer from the surface effect. To offset this,
1003 + \textit{Periodic boundary condition} is developed to simulate bulk
1004 + properties with a relatively small number of particles. In this
1005 + method, the simulation box is replicated throughout space to form an
1006 + infinite lattice. During the simulation, when a particle moves in
1007 + the primary cell, its image in other cells move in exactly the same
1008 + direction with exactly the same orientation. Thus, as a particle
1009 + leaves the primary cell, one of its images will enter through the
1010 + opposite face.
1011 + %\begin{figure}
1012 + %\centering
1013 + %\includegraphics[width=\linewidth]{pbcFig.eps}
1014 + %\caption[An illustration of periodic boundary conditions]{A 2-D
1015 + %illustration of periodic boundary conditions. As one particle leaves
1016 + %the right of the simulation box, an image of it enters the left.}
1017 + %\label{introFig:pbc}
1018 + %\end{figure}
1019 +
1020 + %cutoff and minimum image convention
1021 + Another important technique to improve the efficiency of force
1022 + evaluation is to apply cutoff where particles farther than a
1023 + predetermined distance, are not included in the calculation
1024 + \cite{Frenkel1996}. The use of a cutoff radius will cause a
1025 + discontinuity in the potential energy curve. Fortunately, one can
1026 + shift the potential to ensure the potential curve go smoothly to
1027 + zero at the cutoff radius. Cutoff strategy works pretty well for
1028 + Lennard-Jones interaction because of its short range nature.
1029 + However, simply truncating the electrostatic interaction with the
1030 + use of cutoff has been shown to lead to severe artifacts in
1031 + simulations. Ewald summation, in which the slowly conditionally
1032 + convergent Coulomb potential is transformed into direct and
1033 + reciprocal sums with rapid and absolute convergence, has proved to
1034 + minimize the periodicity artifacts in liquid simulations. Taking the
1035 + advantages of the fast Fourier transform (FFT) for calculating
1036 + discrete Fourier transforms, the particle mesh-based methods are
1037 + accelerated from $O(N^{3/2})$ to $O(N logN)$. An alternative
1038 + approach is \emph{fast multipole method}, which treats Coulombic
1039 + interaction exactly at short range, and approximate the potential at
1040 + long range through multipolar expansion. In spite of their wide
1041 + acceptances at the molecular simulation community, these two methods
1042 + are hard to be implemented correctly and efficiently. Instead, we
1043 + use a damped and charge-neutralized Coulomb potential method
1044 + developed by Wolf and his coworkers. The shifted Coulomb potential
1045 + for particle $i$ and particle $j$ at distance $r_{rj}$ is given by:
1046   \begin{equation}
1047 < H = \frac{{p^2 }}{{2m}} + U(x) + H_B  + \Delta U(x,x_1 , \ldots x_N)
1048 < \label{introEquation:bathGLE}
1047 > V(r_{ij})= \frac{q_i q_j \textrm{erfc}(\alpha
1048 > r_{ij})}{r_{ij}}-\lim_{r_{ij}\rightarrow
1049 > R_\textrm{c}}\left\{\frac{q_iq_j \textrm{erfc}(\alpha
1050 > r_{ij})}{r_{ij}}\right\}. \label{introEquation:shiftedCoulomb}
1051   \end{equation}
1052 < where $H_B$ is harmonic bath Hamiltonian,
1052 > where $\alpha$ is the convergence parameter. Due to the lack of
1053 > inherent periodicity and rapid convergence,this method is extremely
1054 > efficient and easy to implement.
1055 > %\begin{figure}
1056 > %\centering
1057 > %\includegraphics[width=\linewidth]{pbcFig.eps}
1058 > %\caption[An illustration of shifted Coulomb potential]{An illustration of shifted Coulomb potential.}
1059 > %\label{introFigure:shiftedCoulomb}
1060 > %\end{figure}
1061 >
1062 > %multiple time step
1063 >
1064 > \subsection{\label{introSection:Analysis} Analysis}
1065 >
1066 > Recently, advanced visualization technique are widely applied to
1067 > monitor the motions of molecules. Although the dynamics of the
1068 > system can be described qualitatively from animation, quantitative
1069 > trajectory analysis are more appreciable. According to the
1070 > principles of Statistical Mechanics,
1071 > Sec.~\ref{introSection:statisticalMechanics}, one can compute
1072 > thermodynamics properties, analyze fluctuations of structural
1073 > parameters, and investigate time-dependent processes of the molecule
1074 > from the trajectories.
1075 >
1076 > \subsubsection{\label{introSection:thermodynamicsProperties}Thermodynamics Properties}
1077 >
1078 > Thermodynamics properties, which can be expressed in terms of some
1079 > function of the coordinates and momenta of all particles in the
1080 > system, can be directly computed from molecular dynamics. The usual
1081 > way to measure the pressure is based on virial theorem of Clausius
1082 > which states that the virial is equal to $-3Nk_BT$. For a system
1083 > with forces between particles, the total virial, $W$, contains the
1084 > contribution from external pressure and interaction between the
1085 > particles:
1086   \[
1087 < H_B  =\sum\limits_{\alpha  = 1}^N {\left\{ {\frac{{p_\alpha ^2
1088 < }}{{2m_\alpha  }} + \frac{1}{2}m_\alpha  w_\alpha ^2 } \right\}}
1087 > W =  - 3PV + \left\langle {\sum\limits_{i < j} {r{}_{ij} \cdot
1088 > f_{ij} } } \right\rangle
1089   \]
1090 < and $\Delta U$ is bilinear system-bath coupling,
1090 > where $f_{ij}$ is the force between particle $i$ and $j$ at a
1091 > distance $r_{ij}$. Thus, the expression for the pressure is given
1092 > by:
1093 > \begin{equation}
1094 > P = \frac{{Nk_B T}}{V} - \frac{1}{{3V}}\left\langle {\sum\limits_{i
1095 > < j} {r{}_{ij} \cdot f_{ij} } } \right\rangle
1096 > \end{equation}
1097 >
1098 > \subsubsection{\label{introSection:structuralProperties}Structural Properties}
1099 >
1100 > Structural Properties of a simple fluid can be described by a set of
1101 > distribution functions. Among these functions,\emph{pair
1102 > distribution function}, also known as \emph{radial distribution
1103 > function}, is of most fundamental importance to liquid-state theory.
1104 > Pair distribution function can be gathered by Fourier transforming
1105 > raw data from a series of neutron diffraction experiments and
1106 > integrating over the surface factor \cite{Powles73}. The experiment
1107 > result can serve as a criterion to justify the correctness of the
1108 > theory. Moreover, various equilibrium thermodynamic and structural
1109 > properties can also be expressed in terms of radial distribution
1110 > function \cite{allen87:csl}.
1111 >
1112 > A pair distribution functions $g(r)$ gives the probability that a
1113 > particle $i$ will be located at a distance $r$ from a another
1114 > particle $j$ in the system
1115   \[
1116 < \Delta U =  - \sum\limits_{\alpha  = 1}^N {g_\alpha  x_\alpha  x}
1116 > g(r) = \frac{V}{{N^2 }}\left\langle {\sum\limits_i {\sum\limits_{j
1117 > \ne i} {\delta (r - r_{ij} )} } } \right\rangle.
1118   \]
1119 < Completing the square,
1119 > Note that the delta function can be replaced by a histogram in
1120 > computer simulation. Figure
1121 > \ref{introFigure:pairDistributionFunction} shows a typical pair
1122 > distribution function for the liquid argon system. The occurrence of
1123 > several peaks in the plot of $g(r)$ suggests that it is more likely
1124 > to find particles at certain radial values than at others. This is a
1125 > result of the attractive interaction at such distances. Because of
1126 > the strong repulsive forces at short distance, the probability of
1127 > locating particles at distances less than about 2.5{\AA} from each
1128 > other is essentially zero.
1129 >
1130 > %\begin{figure}
1131 > %\centering
1132 > %\includegraphics[width=\linewidth]{pdf.eps}
1133 > %\caption[Pair distribution function for the liquid argon
1134 > %]{Pair distribution function for the liquid argon}
1135 > %\label{introFigure:pairDistributionFunction}
1136 > %\end{figure}
1137 >
1138 > \subsubsection{\label{introSection:timeDependentProperties}Time-dependent
1139 > Properties}
1140 >
1141 > Time-dependent properties are usually calculated using \emph{time
1142 > correlation function}, which correlates random variables $A$ and $B$
1143 > at two different time
1144 > \begin{equation}
1145 > C_{AB} (t) = \left\langle {A(t)B(0)} \right\rangle.
1146 > \label{introEquation:timeCorrelationFunction}
1147 > \end{equation}
1148 > If $A$ and $B$ refer to same variable, this kind of correlation
1149 > function is called \emph{auto correlation function}. One example of
1150 > auto correlation function is velocity auto-correlation function
1151 > which is directly related to transport properties of molecular
1152 > liquids:
1153   \[
1154 < H_B  + \Delta U = \sum\limits_{\alpha  = 1}^N {\left\{
1155 < {\frac{{p_\alpha ^2 }}{{2m_\alpha  }} + \frac{1}{2}m_\alpha
501 < w_\alpha ^2 \left( {x_\alpha   - \frac{{g_\alpha  }}{{m_\alpha
502 < w_\alpha ^2 }}x} \right)^2 } \right\}}  - \sum\limits_{\alpha  =
503 < 1}^N {\frac{{g_\alpha ^2 }}{{2m_\alpha  w_\alpha ^2 }}} x^2
1154 > D = \frac{1}{3}\int\limits_0^\infty  {\left\langle {v(t) \cdot v(0)}
1155 > \right\rangle } dt
1156   \]
1157 < and putting it back into Eq.~\ref{introEquation:bathGLE},
1157 > where $D$ is diffusion constant. Unlike velocity autocorrelation
1158 > function which is averaging over time origins and over all the
1159 > atoms, dipole autocorrelation are calculated for the entire system.
1160 > The dipole autocorrelation function is given by:
1161   \[
1162 < H = \frac{{p^2 }}{{2m}} + W(x) + \sum\limits_{\alpha  = 1}^N
1163 < {\left\{ {\frac{{p_\alpha ^2 }}{{2m_\alpha  }} + \frac{1}{2}m_\alpha
509 < w_\alpha ^2 \left( {x_\alpha   - \frac{{g_\alpha  }}{{m_\alpha
510 < w_\alpha ^2 }}x} \right)^2 } \right\}}
1162 > c_{dipole}  = \left\langle {u_{tot} (t) \cdot u_{tot} (t)}
1163 > \right\rangle
1164   \]
1165 < where
1165 > Here $u_{tot}$ is the net dipole of the entire system and is given
1166 > by
1167 > \[
1168 > u_{tot} (t) = \sum\limits_i {u_i (t)}
1169 > \]
1170 > In principle, many time correlation functions can be related with
1171 > Fourier transforms of the infrared, Raman, and inelastic neutron
1172 > scattering spectra of molecular liquids. In practice, one can
1173 > extract the IR spectrum from the intensity of dipole fluctuation at
1174 > each frequency using the following relationship:
1175 > \[
1176 > \hat c_{dipole} (v) = \int_{ - \infty }^\infty  {c_{dipole} (t)e^{ -
1177 > i2\pi vt} dt}
1178 > \]
1179 >
1180 > \section{\label{introSection:rigidBody}Dynamics of Rigid Bodies}
1181 >
1182 > Rigid bodies are frequently involved in the modeling of different
1183 > areas, from engineering, physics, to chemistry. For example,
1184 > missiles and vehicle are usually modeled by rigid bodies.  The
1185 > movement of the objects in 3D gaming engine or other physics
1186 > simulator is governed by the rigid body dynamics. In molecular
1187 > simulation, rigid body is used to simplify the model in
1188 > protein-protein docking study{\cite{Gray03}}.
1189 >
1190 > It is very important to develop stable and efficient methods to
1191 > integrate the equations of motion of orientational degrees of
1192 > freedom. Euler angles are the nature choice to describe the
1193 > rotational degrees of freedom. However, due to its singularity, the
1194 > numerical integration of corresponding equations of motion is very
1195 > inefficient and inaccurate. Although an alternative integrator using
1196 > different sets of Euler angles can overcome this difficulty\cite{},
1197 > the computational penalty and the lost of angular momentum
1198 > conservation still remain. A singularity free representation
1199 > utilizing quaternions was developed by Evans in 1977. Unfortunately,
1200 > this approach suffer from the nonseparable Hamiltonian resulted from
1201 > quaternion representation, which prevents the symplectic algorithm
1202 > to be utilized. Another different approach is to apply holonomic
1203 > constraints to the atoms belonging to the rigid body. Each atom
1204 > moves independently under the normal forces deriving from potential
1205 > energy and constraint forces which are used to guarantee the
1206 > rigidness. However, due to their iterative nature, SHAKE and Rattle
1207 > algorithm converge very slowly when the number of constraint
1208 > increases.
1209 >
1210 > The break through in geometric literature suggests that, in order to
1211 > develop a long-term integration scheme, one should preserve the
1212 > symplectic structure of the flow. Introducing conjugate momentum to
1213 > rotation matrix $Q$ and re-formulating Hamiltonian's equation, a
1214 > symplectic integrator, RSHAKE, was proposed to evolve the
1215 > Hamiltonian system in a constraint manifold by iteratively
1216 > satisfying the orthogonality constraint $Q_T Q = 1$. An alternative
1217 > method using quaternion representation was developed by Omelyan.
1218 > However, both of these methods are iterative and inefficient. In
1219 > this section, we will present a symplectic Lie-Poisson integrator
1220 > for rigid body developed by Dullweber and his
1221 > coworkers\cite{Dullweber1997} in depth.
1222 >
1223 > \subsection{\label{introSection:constrainedHamiltonianRB}Constrained Hamiltonian for Rigid Body}
1224 > The motion of the rigid body is Hamiltonian with the Hamiltonian
1225 > function
1226 > \begin{equation}
1227 > H = \frac{1}{2}(p^T m^{ - 1} p) + \frac{1}{2}tr(PJ^{ - 1} P) +
1228 > V(q,Q) + \frac{1}{2}tr[(QQ^T  - 1)\Lambda ].
1229 > \label{introEquation:RBHamiltonian}
1230 > \end{equation}
1231 > Here, $q$ and $Q$  are the position and rotation matrix for the
1232 > rigid-body, $p$ and $P$  are conjugate momenta to $q$  and $Q$ , and
1233 > $J$, a diagonal matrix, is defined by
1234   \[
1235 < W(x) = U(x) - \sum\limits_{\alpha  = 1}^N {\frac{{g_\alpha ^2
515 < }}{{2m_\alpha  w_\alpha ^2 }}} x^2
1235 > I_{ii}^{ - 1}  = \frac{1}{2}\sum\limits_{i \ne j} {J_{jj}^{ - 1} }
1236   \]
1237 < Since the first two terms of the new Hamiltonian depend only on the
1238 < system coordinates, we can get the equations of motion for
1239 < Generalized Langevin Dynamics by Hamilton's equations
1240 < \ref{introEquation:motionHamiltonianCoordinate,
1241 < introEquation:motionHamiltonianMomentum},
1242 < \begin{align}
1243 < \dot p &=  - \frac{{\partial H}}{{\partial x}}
1244 <       &= m\ddot x
1245 <       &= - \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)}
1246 < \label{introEq:Lp5}
1247 < \end{align}
1248 < , and
529 < \begin{align}
530 < \dot p_\alpha   &=  - \frac{{\partial H}}{{\partial x_\alpha  }}
531 <                &= m\ddot x_\alpha
532 <                &= \- m_\alpha  w_\alpha ^2 \left( {x_\alpha   - \frac{{g_\alpha}}{{m_\alpha  w_\alpha ^2 }}x} \right)
533 < \end{align}
1237 > where $I_{ii}$ is the diagonal element of the inertia tensor. This
1238 > constrained Hamiltonian equation subjects to a holonomic constraint,
1239 > \begin{equation}
1240 > Q^T Q = 1, \label{introEquation:orthogonalConstraint}
1241 > \end{equation}
1242 > which is used to ensure rotation matrix's orthogonality.
1243 > Differentiating \ref{introEquation:orthogonalConstraint} and using
1244 > Equation \ref{introEquation:RBMotionMomentum}, one may obtain,
1245 > \begin{equation}
1246 > Q^T PJ^{ - 1}  + J^{ - 1} P^T Q = 0 . \\
1247 > \label{introEquation:RBFirstOrderConstraint}
1248 > \end{equation}
1249  
1250 < \subsection{\label{introSection:laplaceTransform}The Laplace Transform}
1250 > Using Equation (\ref{introEquation:motionHamiltonianCoordinate},
1251 > \ref{introEquation:motionHamiltonianMomentum}), one can write down
1252 > the equations of motion,
1253 > \[
1254 > \begin{array}{c}
1255 > \frac{{dq}}{{dt}} = \frac{p}{m} \label{introEquation:RBMotionPosition}\\
1256 > \frac{{dp}}{{dt}} =  - \nabla _q V(q,Q) \label{introEquation:RBMotionMomentum}\\
1257 > \frac{{dQ}}{{dt}} = PJ^{ - 1}  \label{introEquation:RBMotionRotation}\\
1258 > \frac{{dP}}{{dt}} =  - \nabla _Q V(q,Q) - 2Q\Lambda . \label{introEquation:RBMotionP}\\
1259 > \end{array}
1260 > \]
1261  
1262 + In general, there are two ways to satisfy the holonomic constraints.
1263 + We can use constraint force provided by lagrange multiplier on the
1264 + normal manifold to keep the motion on constraint space. Or we can
1265 + simply evolve the system in constraint manifold. These two methods
1266 + are proved to be equivalent. The holonomic constraint and equations
1267 + of motions define a constraint manifold for rigid body
1268   \[
1269 < L(x) = \int_0^\infty  {x(t)e^{ - pt} dt}
1269 > M = \left\{ {(Q,P):Q^T Q = 1,Q^T PJ^{ - 1}  + J^{ - 1} P^T Q = 0}
1270 > \right\}.
1271   \]
1272  
1273 + Unfortunately, this constraint manifold is not the cotangent bundle
1274 + $T_{\star}SO(3)$. However, it turns out that under symplectic
1275 + transformation, the cotangent space and the phase space are
1276 + diffeomorphic. Introducing
1277   \[
1278 < L(x + y) = L(x) + L(y)
1278 > \tilde Q = Q,\tilde P = \frac{1}{2}\left( {P - QP^T Q} \right),
1279   \]
1280 + the mechanical system subject to a holonomic constraint manifold $M$
1281 + can be re-formulated as a Hamiltonian system on the cotangent space
1282 + \[
1283 + T^* SO(3) = \left\{ {(\tilde Q,\tilde P):\tilde Q^T \tilde Q =
1284 + 1,\tilde Q^T \tilde PJ^{ - 1}  + J^{ - 1} P^T \tilde Q = 0} \right\}
1285 + \]
1286  
1287 + For a body fixed vector $X_i$ with respect to the center of mass of
1288 + the rigid body, its corresponding lab fixed vector $X_0^{lab}$  is
1289 + given as
1290 + \begin{equation}
1291 + X_i^{lab} = Q X_i + q.
1292 + \end{equation}
1293 + Therefore, potential energy $V(q,Q)$ is defined by
1294   \[
1295 < L(ax) = aL(x)
1295 > V(q,Q) = V(Q X_0 + q).
1296   \]
1297 + Hence, the force and torque are given by
1298 + \[
1299 + \nabla _q V(q,Q) = F(q,Q) = \sum\limits_i {F_i (q,Q)},
1300 + \]
1301 + and
1302 + \[
1303 + \nabla _Q V(q,Q) = F(q,Q)X_i^t
1304 + \]
1305 + respectively.
1306  
1307 + As a common choice to describe the rotation dynamics of the rigid
1308 + body, angular momentum on body frame $\Pi  = Q^t P$ is introduced to
1309 + rewrite the equations of motion,
1310 + \begin{equation}
1311 + \begin{array}{l}
1312 + \mathop \Pi \limits^ \bullet   = J^{ - 1} \Pi ^T \Pi  + Q^T \sum\limits_i {F_i (q,Q)X_i^T }  - \Lambda  \\
1313 + \mathop Q\limits^{{\rm{   }} \bullet }  = Q\Pi {\rm{ }}J^{ - 1}  \\
1314 + \end{array}
1315 + \label{introEqaution:RBMotionPI}
1316 + \end{equation}
1317 + , as well as holonomic constraints,
1318   \[
1319 < L(\dot x) = pL(x) - px(0)
1319 > \begin{array}{l}
1320 > \Pi J^{ - 1}  + J^{ - 1} \Pi ^t  = 0 \\
1321 > Q^T Q = 1 \\
1322 > \end{array}
1323   \]
1324  
1325 + For a vector $v(v_1 ,v_2 ,v_3 ) \in R^3$ and a matrix $\hat v \in
1326 + so(3)^ \star$, the hat-map isomorphism,
1327 + \begin{equation}
1328 + v(v_1 ,v_2 ,v_3 ) \Leftrightarrow \hat v = \left(
1329 + {\begin{array}{*{20}c}
1330 +   0 & { - v_3 } & {v_2 }  \\
1331 +   {v_3 } & 0 & { - v_1 }  \\
1332 +   { - v_2 } & {v_1 } & 0  \\
1333 + \end{array}} \right),
1334 + \label{introEquation:hatmapIsomorphism}
1335 + \end{equation}
1336 + will let us associate the matrix products with traditional vector
1337 + operations
1338   \[
1339 < L(\ddot x) = p^2 L(x) - px(0) - \dot x(0)
1339 > \hat vu = v \times u
1340   \]
1341  
1342 + Using \ref{introEqaution:RBMotionPI}, one can construct a skew
1343 + matrix,
1344 + \begin{equation}
1345 + (\mathop \Pi \limits^ \bullet   - \mathop \Pi \limits^ \bullet  ^T
1346 + ){\rm{ }} = {\rm{ }}(\Pi  - \Pi ^T ){\rm{ }}(J^{ - 1} \Pi  + \Pi J^{
1347 + - 1} ) + \sum\limits_i {[Q^T F_i (r,Q)X_i^T  - X_i F_i (r,Q)^T Q]} -
1348 + (\Lambda  - \Lambda ^T ) . \label{introEquation:skewMatrixPI}
1349 + \end{equation}
1350 + Since $\Lambda$ is symmetric, the last term of Equation
1351 + \ref{introEquation:skewMatrixPI} is zero, which implies the Lagrange
1352 + multiplier $\Lambda$ is absent from the equations of motion. This
1353 + unique property eliminate the requirement of iterations which can
1354 + not be avoided in other methods\cite{}.
1355 +
1356 + Applying hat-map isomorphism, we obtain the equation of motion for
1357 + angular momentum on body frame
1358 + \begin{equation}
1359 + \dot \pi  = \pi  \times I^{ - 1} \pi  + \sum\limits_i {\left( {Q^T
1360 + F_i (r,Q)} \right) \times X_i }.
1361 + \label{introEquation:bodyAngularMotion}
1362 + \end{equation}
1363 + In the same manner, the equation of motion for rotation matrix is
1364 + given by
1365   \[
1366 < L\left( {\int_0^t {g(t - \tau )h(\tau )d\tau } } \right) = G(p)H(p)
1366 > \dot Q = Qskew(I^{ - 1} \pi )
1367   \]
1368  
1369 < Some relatively important transformation,
1369 > \subsection{\label{introSection:SymplecticFreeRB}Symplectic
1370 > Lie-Poisson Integrator for Free Rigid Body}
1371 >
1372 > If there is not external forces exerted on the rigid body, the only
1373 > contribution to the rotational is from the kinetic potential (the
1374 > first term of \ref{ introEquation:bodyAngularMotion}). The free
1375 > rigid body is an example of Lie-Poisson system with Hamiltonian
1376 > function
1377 > \begin{equation}
1378 > T^r (\pi ) = T_1 ^r (\pi _1 ) + T_2^r (\pi _2 ) + T_3^r (\pi _3 )
1379 > \label{introEquation:rotationalKineticRB}
1380 > \end{equation}
1381 > where $T_i^r (\pi _i ) = \frac{{\pi _i ^2 }}{{2I_i }}$ and
1382 > Lie-Poisson structure matrix,
1383 > \begin{equation}
1384 > J(\pi ) = \left( {\begin{array}{*{20}c}
1385 >   0 & {\pi _3 } & { - \pi _2 }  \\
1386 >   { - \pi _3 } & 0 & {\pi _1 }  \\
1387 >   {\pi _2 } & { - \pi _1 } & 0  \\
1388 > \end{array}} \right)
1389 > \end{equation}
1390 > Thus, the dynamics of free rigid body is governed by
1391 > \begin{equation}
1392 > \frac{d}{{dt}}\pi  = J(\pi )\nabla _\pi  T^r (\pi )
1393 > \end{equation}
1394 >
1395 > One may notice that each $T_i^r$ in Equation
1396 > \ref{introEquation:rotationalKineticRB} can be solved exactly. For
1397 > instance, the equations of motion due to $T_1^r$ are given by
1398 > \begin{equation}
1399 > \frac{d}{{dt}}\pi  = R_1 \pi ,\frac{d}{{dt}}Q = QR_1
1400 > \label{introEqaution:RBMotionSingleTerm}
1401 > \end{equation}
1402 > where
1403 > \[ R_1  = \left( {\begin{array}{*{20}c}
1404 >   0 & 0 & 0  \\
1405 >   0 & 0 & {\pi _1 }  \\
1406 >   0 & { - \pi _1 } & 0  \\
1407 > \end{array}} \right).
1408 > \]
1409 > The solutions of Equation \ref{introEqaution:RBMotionSingleTerm} is
1410   \[
1411 < L(\cos at) = \frac{p}{{p^2  + a^2 }}
1411 > \pi (\Delta t) = e^{\Delta tR_1 } \pi (0),Q(\Delta t) =
1412 > Q(0)e^{\Delta tR_1 }
1413   \]
1414 + with
1415 + \[
1416 + e^{\Delta tR_1 }  = \left( {\begin{array}{*{20}c}
1417 +   0 & 0 & 0  \\
1418 +   0 & {\cos \theta _1 } & {\sin \theta _1 }  \\
1419 +   0 & { - \sin \theta _1 } & {\cos \theta _1 }  \\
1420 + \end{array}} \right),\theta _1  = \frac{{\pi _1 }}{{I_1 }}\Delta t.
1421 + \]
1422 + To reduce the cost of computing expensive functions in $e^{\Delta
1423 + tR_1 }$, we can use Cayley transformation,
1424 + \[
1425 + e^{\Delta tR_1 }  \approx (1 - \Delta tR_1 )^{ - 1} (1 + \Delta tR_1
1426 + )
1427 + \]
1428 + The flow maps for $T_2^r$ and $T_3^r$ can be found in the same
1429 + manner.
1430  
1431 + In order to construct a second-order symplectic method, we split the
1432 + angular kinetic Hamiltonian function can into five terms
1433   \[
1434 < L(\sin at) = \frac{a}{{p^2  + a^2 }}
1434 > T^r (\pi ) = \frac{1}{2}T_1 ^r (\pi _1 ) + \frac{1}{2}T_2^r (\pi _2
1435 > ) + T_3^r (\pi _3 ) + \frac{1}{2}T_2^r (\pi _2 ) + \frac{1}{2}T_1 ^r
1436 > (\pi _1 )
1437 > \].
1438 > Concatenating flows corresponding to these five terms, we can obtain
1439 > an symplectic integrator,
1440 > \[
1441 > \varphi _{\Delta t,T^r }  = \varphi _{\Delta t/2,\pi _1 }  \circ
1442 > \varphi _{\Delta t/2,\pi _2 }  \circ \varphi _{\Delta t,\pi _3 }
1443 > \circ \varphi _{\Delta t/2,\pi _2 }  \circ \varphi _{\Delta t/2,\pi
1444 > _1 }.
1445   \]
1446  
1447 + The non-canonical Lie-Poisson bracket ${F, G}$ of two function
1448 + $F(\pi )$ and $G(\pi )$ is defined by
1449   \[
1450 < L(1) = \frac{1}{p}
1450 > \{ F,G\} (\pi ) = [\nabla _\pi  F(\pi )]^T J(\pi )\nabla _\pi  G(\pi
1451 > )
1452   \]
1453 + If the Poisson bracket of a function $F$ with an arbitrary smooth
1454 + function $G$ is zero, $F$ is a \emph{Casimir}, which is the
1455 + conserved quantity in Poisson system. We can easily verify that the
1456 + norm of the angular momentum, $\parallel \pi
1457 + \parallel$, is a \emph{Casimir}. Let$ F(\pi ) = S(\frac{{\parallel
1458 + \pi \parallel ^2 }}{2})$ for an arbitrary function $ S:R \to R$ ,
1459 + then by the chain rule
1460 + \[
1461 + \nabla _\pi  F(\pi ) = S'(\frac{{\parallel \pi \parallel ^2
1462 + }}{2})\pi
1463 + \]
1464 + Thus $ [\nabla _\pi  F(\pi )]^T J(\pi ) =  - S'(\frac{{\parallel \pi
1465 + \parallel ^2 }}{2})\pi  \times \pi  = 0 $. This explicit
1466 + Lie-Poisson integrator is found to be extremely efficient and stable
1467 + which can be explained by the fact the small angle approximation is
1468 + used and the norm of the angular momentum is conserved.
1469  
1470 < First, the bath coordinates,
1470 > \subsection{\label{introSection:RBHamiltonianSplitting} Hamiltonian
1471 > Splitting for Rigid Body}
1472 >
1473 > The Hamiltonian of rigid body can be separated in terms of kinetic
1474 > energy and potential energy,
1475   \[
1476 < p^2 L(x_\alpha  ) - px_\alpha  (0) - \dot x_\alpha  (0) =  - \omega
577 < _\alpha ^2 L(x_\alpha  ) + \frac{{g_\alpha  }}{{\omega _\alpha
578 < }}L(x)
1476 > H = T(p,\pi ) + V(q,Q)
1477   \]
1478 + The equations of motion corresponding to potential energy and
1479 + kinetic energy are listed in the below table,
1480 + \begin{table}
1481 + \caption{Equations of motion due to Potential and Kinetic Energies}
1482 + \begin{center}
1483 + \begin{tabular}{|l|l|}
1484 +  \hline
1485 +  % after \\: \hline or \cline{col1-col2} \cline{col3-col4} ...
1486 +  Potential & Kinetic \\
1487 +  $\frac{{dq}}{{dt}} = \frac{p}{m}$ & $\frac{d}{{dt}}q = p$ \\
1488 +  $\frac{d}{{dt}}p =  - \frac{{\partial V}}{{\partial q}}$ & $ \frac{d}{{dt}}p = 0$ \\
1489 +  $\frac{d}{{dt}}Q = 0$ & $ \frac{d}{{dt}}Q = Qskew(I^{ - 1} j)$ \\
1490 +  $ \frac{d}{{dt}}\pi  = \sum\limits_i {\left( {Q^T F_i (r,Q)} \right) \times X_i }$ & $\frac{d}{{dt}}\pi  = \pi  \times I^{ - 1} \pi$\\
1491 +  \hline
1492 + \end{tabular}
1493 + \end{center}
1494 + \end{table}
1495 + A second-order symplectic method is now obtained by the
1496 + composition of the flow maps,
1497   \[
1498 < L(x_\alpha  ) = \frac{{\frac{{g_\alpha  }}{{\omega _\alpha  }}L(x) +
1499 < px_\alpha  (0) + \dot x_\alpha  (0)}}{{p^2  + \omega _\alpha ^2 }}
1498 > \varphi _{\Delta t}  = \varphi _{\Delta t/2,V}  \circ \varphi
1499 > _{\Delta t,T}  \circ \varphi _{\Delta t/2,V}.
1500   \]
1501 < Then, the system coordinates,
1502 < \begin{align}
1503 < mL(\ddot x) &=  - \frac{1}{p}\frac{{\partial W(x)}}{{\partial x}} -
1504 < \sum\limits_{\alpha  = 1}^N {\left\{ {\frac{{\frac{{g_\alpha
1505 < }}{{\omega _\alpha  }}L(x) + px_\alpha  (0) + \dot x_\alpha
1506 < (0)}}{{p^2  + \omega _\alpha ^2 }} - \frac{{g_\alpha ^2 }}{{m_\alpha
1507 < }}\omega _\alpha ^2 L(x)} \right\}}
1508 < %
1509 < &= - \frac{1}{p}\frac{{\partial W(x)}}{{\partial x}} -
593 < \sum\limits_{\alpha  = 1}^N {\left\{ { - \frac{{g_\alpha ^2 }}{{m_\alpha  \omega _\alpha ^2 }}\frac{p}{{p^2  + \omega _\alpha ^2 }}pL(x)
594 < - \frac{p}{{p^2  + \omega _\alpha ^2 }}g_\alpha  x_\alpha  (0)
595 < - \frac{1}{{p^2  + \omega _\alpha ^2 }}g_\alpha  \dot x_\alpha  (0)} \right\}}
596 < \end{align}
597 < Then, the inverse transform,
1501 > Moreover, $\varphi _{\Delta t/2,V}$ can be divided into two
1502 > sub-flows which corresponding to force and torque respectively,
1503 > \[
1504 > \varphi _{\Delta t/2,V}  = \varphi _{\Delta t/2,F}  \circ \varphi
1505 > _{\Delta t/2,\tau }.
1506 > \]
1507 > Since the associated operators of $\varphi _{\Delta t/2,F} $ and
1508 > $\circ \varphi _{\Delta t/2,\tau }$ are commuted, the composition
1509 > order inside $\varphi _{\Delta t/2,V}$ does not matter.
1510  
1511 + Furthermore, kinetic potential can be separated to translational
1512 + kinetic term, $T^t (p)$, and rotational kinetic term, $T^r (\pi )$,
1513 + \begin{equation}
1514 + T(p,\pi ) =T^t (p) + T^r (\pi ).
1515 + \end{equation}
1516 + where $ T^t (p) = \frac{1}{2}p^T m^{ - 1} p $ and $T^r (\pi )$ is
1517 + defined by \ref{introEquation:rotationalKineticRB}. Therefore, the
1518 + corresponding flow maps are given by
1519 + \[
1520 + \varphi _{\Delta t,T}  = \varphi _{\Delta t,T^t }  \circ \varphi
1521 + _{\Delta t,T^r }.
1522 + \]
1523 + Finally, we obtain the overall symplectic flow maps for free moving
1524 + rigid body
1525 + \begin{equation}
1526 + \begin{array}{c}
1527 + \varphi _{\Delta t}  = \varphi _{\Delta t/2,F}  \circ \varphi _{\Delta t/2,\tau }  \\
1528 +  \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 }  \\
1529 +  \circ \varphi _{\Delta t/2,\tau }  \circ \varphi _{\Delta t/2,F}  .\\
1530 + \end{array}
1531 + \label{introEquation:overallRBFlowMaps}
1532 + \end{equation}
1533 +
1534 + \section{\label{introSection:langevinDynamics}Langevin Dynamics}
1535 + As an alternative to newtonian dynamics, Langevin dynamics, which
1536 + mimics a simple heat bath with stochastic and dissipative forces,
1537 + has been applied in a variety of studies. This section will review
1538 + the theory of Langevin dynamics simulation. A brief derivation of
1539 + generalized Langevin equation will be given first. Follow that, we
1540 + will discuss the physical meaning of the terms appearing in the
1541 + equation as well as the calculation of friction tensor from
1542 + hydrodynamics theory.
1543 +
1544 + \subsection{\label{introSection:generalizedLangevinDynamics}Derivation of Generalized Langevin Equation}
1545 +
1546 + Harmonic bath model, in which an effective set of harmonic
1547 + oscillators are used to mimic the effect of a linearly responding
1548 + environment, has been widely used in quantum chemistry and
1549 + statistical mechanics. One of the successful applications of
1550 + Harmonic bath model is the derivation of Deriving Generalized
1551 + Langevin Dynamics. Lets consider a system, in which the degree of
1552 + freedom $x$ is assumed to couple to the bath linearly, giving a
1553 + Hamiltonian of the form
1554 + \begin{equation}
1555 + H = \frac{{p^2 }}{{2m}} + U(x) + H_B  + \Delta U(x,x_1 , \ldots x_N)
1556 + \label{introEquation:bathGLE}.
1557 + \end{equation}
1558 + Here $p$ is a momentum conjugate to $q$, $m$ is the mass associated
1559 + with this degree of freedom, $H_B$ is harmonic bath Hamiltonian,
1560 + \[
1561 + H_B  = \sum\limits_{\alpha  = 1}^N {\left\{ {\frac{{p_\alpha ^2
1562 + }}{{2m_\alpha  }} + \frac{1}{2}m_\alpha  \omega _\alpha ^2 }
1563 + \right\}}
1564 + \]
1565 + where the index $\alpha$ runs over all the bath degrees of freedom,
1566 + $\omega _\alpha$ are the harmonic bath frequencies, $m_\alpha$ are
1567 + the harmonic bath masses, and $\Delta U$ is bilinear system-bath
1568 + coupling,
1569 + \[
1570 + \Delta U =  - \sum\limits_{\alpha  = 1}^N {g_\alpha  x_\alpha  x}
1571 + \]
1572 + where $g_\alpha$ are the coupling constants between the bath and the
1573 + coordinate $x$. Introducing
1574 + \[
1575 + W(x) = U(x) - \sum\limits_{\alpha  = 1}^N {\frac{{g_\alpha ^2
1576 + }}{{2m_\alpha  w_\alpha ^2 }}} x^2
1577 + \] and combining the last two terms in Equation
1578 + \ref{introEquation:bathGLE}, we may rewrite the Harmonic bath
1579 + Hamiltonian as
1580 + \[
1581 + H = \frac{{p^2 }}{{2m}} + W(x) + \sum\limits_{\alpha  = 1}^N
1582 + {\left\{ {\frac{{p_\alpha ^2 }}{{2m_\alpha  }} + \frac{1}{2}m_\alpha
1583 + w_\alpha ^2 \left( {x_\alpha   - \frac{{g_\alpha  }}{{m_\alpha
1584 + w_\alpha ^2 }}x} \right)^2 } \right\}}
1585 + \]
1586 + Since the first two terms of the new Hamiltonian depend only on the
1587 + system coordinates, we can get the equations of motion for
1588 + Generalized Langevin Dynamics by Hamilton's equations
1589 + \ref{introEquation:motionHamiltonianCoordinate,
1590 + introEquation:motionHamiltonianMomentum},
1591 + \begin{equation}
1592 + m\ddot x =  - \frac{{\partial W(x)}}{{\partial x}} -
1593 + \sum\limits_{\alpha  = 1}^N {g_\alpha  \left( {x_\alpha   -
1594 + \frac{{g_\alpha  }}{{m_\alpha  w_\alpha ^2 }}x} \right)},
1595 + \label{introEquation:coorMotionGLE}
1596 + \end{equation}
1597 + and
1598 + \begin{equation}
1599 + m\ddot x_\alpha   =  - m_\alpha  w_\alpha ^2 \left( {x_\alpha   -
1600 + \frac{{g_\alpha  }}{{m_\alpha  w_\alpha ^2 }}x} \right).
1601 + \label{introEquation:bathMotionGLE}
1602 + \end{equation}
1603 +
1604 + In order to derive an equation for $x$, the dynamics of the bath
1605 + variables $x_\alpha$ must be solved exactly first. As an integral
1606 + transform which is particularly useful in solving linear ordinary
1607 + differential equations, Laplace transform is the appropriate tool to
1608 + solve this problem. The basic idea is to transform the difficult
1609 + differential equations into simple algebra problems which can be
1610 + solved easily. Then applying inverse Laplace transform, also known
1611 + as the Bromwich integral, we can retrieve the solutions of the
1612 + original problems.
1613 +
1614 + Let $f(t)$ be a function defined on $ [0,\infty ) $. The Laplace
1615 + transform of f(t) is a new function defined as
1616 + \[
1617 + L(f(t)) \equiv F(p) = \int_0^\infty  {f(t)e^{ - pt} dt}
1618 + \]
1619 + where  $p$ is real and  $L$ is called the Laplace Transform
1620 + Operator. Below are some important properties of Laplace transform
1621 + \begin{equation}
1622 + \begin{array}{c}
1623 + L(x + y) = L(x) + L(y) \\
1624 + L(ax) = aL(x) \\
1625 + L(\dot x) = pL(x) - px(0) \\
1626 + L(\ddot x) = p^2 L(x) - px(0) - \dot x(0) \\
1627 + L\left( {\int_0^t {g(t - \tau )h(\tau )d\tau } } \right) = G(p)H(p) \\
1628 + \end{array}
1629 + \end{equation}
1630 +
1631 + Applying Laplace transform to the bath coordinates, we obtain
1632 + \[
1633 + \begin{array}{c}
1634 + p^2 L(x_\alpha  ) - px_\alpha  (0) - \dot x_\alpha  (0) =  - \omega _\alpha ^2 L(x_\alpha  ) + \frac{{g_\alpha  }}{{\omega _\alpha  }}L(x) \\
1635 + L(x_\alpha  ) = \frac{{\frac{{g_\alpha  }}{{\omega _\alpha  }}L(x) + px_\alpha  (0) + \dot x_\alpha  (0)}}{{p^2  + \omega _\alpha ^2 }} \\
1636 + \end{array}
1637 + \]
1638 + By the same way, the system coordinates become
1639 + \[
1640 + \begin{array}{c}
1641 + mL(\ddot x) =  - \frac{1}{p}\frac{{\partial W(x)}}{{\partial x}} \\
1642 +  - \sum\limits_{\alpha  = 1}^N {\left\{ { - \frac{{g_\alpha ^2 }}{{m_\alpha  \omega _\alpha ^2 }}\frac{p}{{p^2  + \omega _\alpha ^2 }}pL(x) - \frac{p}{{p^2  + \omega _\alpha ^2 }}g_\alpha  x_\alpha  (0) - \frac{1}{{p^2  + \omega _\alpha ^2 }}g_\alpha  \dot x_\alpha  (0)} \right\}}  \\
1643 + \end{array}
1644 + \]
1645 +
1646 + With the help of some relatively important inverse Laplace
1647 + transformations:
1648 + \[
1649 + \begin{array}{c}
1650 + L(\cos at) = \frac{p}{{p^2  + a^2 }} \\
1651 + L(\sin at) = \frac{a}{{p^2  + a^2 }} \\
1652 + L(1) = \frac{1}{p} \\
1653 + \end{array}
1654 + \]
1655 + , we obtain
1656   \begin{align}
1657   m\ddot x &=  - \frac{{\partial W(x)}}{{\partial x}} -
1658   \sum\limits_{\alpha  = 1}^N {\left\{ {\left( { - \frac{{g_\alpha ^2
# Line 615 | Line 1672 | t)\dot x(t - \tau )d} \tau }  + \sum\limits_{\alpha  =
1672   (\omega _\alpha  t)} \right\}}
1673   \end{align}
1674  
1675 + Introducing a \emph{dynamic friction kernel}
1676   \begin{equation}
1677 + \xi (t) = \sum\limits_{\alpha  = 1}^N {\left( { - \frac{{g_\alpha ^2
1678 + }}{{m_\alpha  \omega _\alpha ^2 }}} \right)\cos (\omega _\alpha  t)}
1679 + \label{introEquation:dynamicFrictionKernelDefinition}
1680 + \end{equation}
1681 + and \emph{a random force}
1682 + \begin{equation}
1683 + R(t) = \sum\limits_{\alpha  = 1}^N {\left( {g_\alpha  x_\alpha  (0)
1684 + - \frac{{g_\alpha ^2 }}{{m_\alpha  \omega _\alpha ^2 }}x(0)}
1685 + \right)\cos (\omega _\alpha  t)}  + \frac{{\dot x_\alpha
1686 + (0)}}{{\omega _\alpha  }}\sin (\omega _\alpha  t),
1687 + \label{introEquation:randomForceDefinition}
1688 + \end{equation}
1689 + the equation of motion can be rewritten as
1690 + \begin{equation}
1691   m\ddot x =  - \frac{{\partial W}}{{\partial x}} - \int_0^t {\xi
1692   (t)\dot x(t - \tau )d\tau }  + R(t)
1693   \label{introEuqation:GeneralizedLangevinDynamics}
1694   \end{equation}
1695 < %where $ {\xi (t)}$ is friction kernel, $R(t)$ is random force and
1696 < %$W$ is the potential of mean force. $W(x) =  - kT\ln p(x)$
1695 > which is known as the \emph{generalized Langevin equation}.
1696 >
1697 > \subsubsection{\label{introSection:randomForceDynamicFrictionKernel}Random Force and Dynamic Friction Kernel}
1698 >
1699 > One may notice that $R(t)$ depends only on initial conditions, which
1700 > implies it is completely deterministic within the context of a
1701 > harmonic bath. However, it is easy to verify that $R(t)$ is totally
1702 > uncorrelated to $x$ and $\dot x$,
1703   \[
1704 < \xi (t) = \sum\limits_{\alpha  = 1}^N {\left( { - \frac{{g_\alpha ^2
1705 < }}{{m_\alpha  \omega _\alpha ^2 }}} \right)\cos (\omega _\alpha  t)}
1704 > \begin{array}{l}
1705 > \left\langle {x(t)R(t)} \right\rangle  = 0, \\
1706 > \left\langle {\dot x(t)R(t)} \right\rangle  = 0. \\
1707 > \end{array}
1708   \]
1709 < For an infinite harmonic bath, we can use the spectral density and
1710 < an integral over frequencies.
1709 > This property is what we expect from a truly random process. As long
1710 > as the model, which is gaussian distribution in general, chosen for
1711 > $R(t)$ is a truly random process, the stochastic nature of the GLE
1712 > still remains.
1713  
1714 + %dynamic friction kernel
1715 + The convolution integral
1716   \[
1717 < R(t) = \sum\limits_{\alpha  = 1}^N {\left( {g_\alpha  x_\alpha  (0)
634 < - \frac{{g_\alpha ^2 }}{{m_\alpha  \omega _\alpha ^2 }}x(0)}
635 < \right)\cos (\omega _\alpha  t)}  + \frac{{\dot x_\alpha
636 < (0)}}{{\omega _\alpha  }}\sin (\omega _\alpha  t)
1717 > \int_0^t {\xi (t)\dot x(t - \tau )d\tau }
1718   \]
1719 < The random forces depend only on initial conditions.
1719 > depends on the entire history of the evolution of $x$, which implies
1720 > that the bath retains memory of previous motions. In other words,
1721 > the bath requires a finite time to respond to change in the motion
1722 > of the system. For a sluggish bath which responds slowly to changes
1723 > in the system coordinate, we may regard $\xi(t)$ as a constant
1724 > $\xi(t) = \Xi_0$. Hence, the convolution integral becomes
1725 > \[
1726 > \int_0^t {\xi (t)\dot x(t - \tau )d\tau }  = \xi _0 (x(t) - x(0))
1727 > \]
1728 > and Equation \ref{introEuqation:GeneralizedLangevinDynamics} becomes
1729 > \[
1730 > m\ddot x =  - \frac{\partial }{{\partial x}}\left( {W(x) +
1731 > \frac{1}{2}\xi _0 (x - x_0 )^2 } \right) + R(t),
1732 > \]
1733 > which can be used to describe dynamic caging effect. The other
1734 > extreme is the bath that responds infinitely quickly to motions in
1735 > the system. Thus, $\xi (t)$ can be taken as a $delta$ function in
1736 > time:
1737 > \[
1738 > \xi (t) = 2\xi _0 \delta (t)
1739 > \]
1740 > Hence, the convolution integral becomes
1741 > \[
1742 > \int_0^t {\xi (t)\dot x(t - \tau )d\tau }  = 2\xi _0 \int_0^t
1743 > {\delta (t)\dot x(t - \tau )d\tau }  = \xi _0 \dot x(t),
1744 > \]
1745 > and Equation \ref{introEuqation:GeneralizedLangevinDynamics} becomes
1746 > \begin{equation}
1747 > m\ddot x =  - \frac{{\partial W(x)}}{{\partial x}} - \xi _0 \dot
1748 > x(t) + R(t) \label{introEquation:LangevinEquation}
1749 > \end{equation}
1750 > which is known as the Langevin equation. The static friction
1751 > coefficient $\xi _0$ can either be calculated from spectral density
1752 > or be determined by Stokes' law for regular shaped particles.A
1753 > briefly review on calculating friction tensor for arbitrary shaped
1754 > particles is given in Sec.~\ref{introSection:frictionTensor}.
1755  
1756   \subsubsection{\label{introSection:secondFluctuationDissipation}The Second Fluctuation Dissipation Theorem}
1757 < So we can define a new set of coordinates,
1757 >
1758 > Defining a new set of coordinates,
1759   \[
1760   q_\alpha  (t) = x_\alpha  (t) - \frac{1}{{m_\alpha  \omega _\alpha
1761   ^2 }}x(0)
1762 < \]
1763 < This makes
1762 > \],
1763 > we can rewrite $R(T)$ as
1764   \[
1765 < R(t) = \sum\limits_{\alpha  = 1}^N {g_\alpha  q_\alpha  (t)}
1765 > R(t) = \sum\limits_{\alpha  = 1}^N {g_\alpha  q_\alpha  (t)}.
1766   \]
1767   And since the $q$ coordinates are harmonic oscillators,
1768   \[
1769 < \begin{array}{l}
1769 > \begin{array}{c}
1770 > \left\langle {q_\alpha ^2 } \right\rangle  = \frac{{kT}}{{m_\alpha  \omega _\alpha ^2 }} \\
1771   \left\langle {q_\alpha  (t)q_\alpha  (0)} \right\rangle  = \left\langle {q_\alpha ^2 (0)} \right\rangle \cos (\omega _\alpha  t) \\
1772   \left\langle {q_\alpha  (t)q_\beta  (0)} \right\rangle  = \delta _{\alpha \beta } \left\langle {q_\alpha  (t)q_\alpha  (0)} \right\rangle  \\
1773 + \left\langle {R(t)R(0)} \right\rangle  = \sum\limits_\alpha  {\sum\limits_\beta  {g_\alpha  g_\beta  \left\langle {q_\alpha  (t)q_\beta  (0)} \right\rangle } }  \\
1774 +  = \sum\limits_\alpha  {g_\alpha ^2 \left\langle {q_\alpha ^2 (0)} \right\rangle \cos (\omega _\alpha  t)}  \\
1775 +  = kT\xi (t) \\
1776   \end{array}
1777   \]
1778 <
658 < \begin{align}
659 < \left\langle {R(t)R(0)} \right\rangle  &= \sum\limits_\alpha
660 < {\sum\limits_\beta  {g_\alpha  g_\beta  \left\langle {q_\alpha
661 < (t)q_\beta  (0)} \right\rangle } }
662 < %
663 < &= \sum\limits_\alpha  {g_\alpha ^2 \left\langle {q_\alpha ^2 (0)}
664 < \right\rangle \cos (\omega _\alpha  t)}
665 < %
666 < &= kT\xi (t)
667 < \end{align}
668 <
1778 > Thus, we recover the \emph{second fluctuation dissipation theorem}
1779   \begin{equation}
1780   \xi (t) = \left\langle {R(t)R(0)} \right\rangle
1781 < \label{introEquation:secondFluctuationDissipation}
1781 > \label{introEquation:secondFluctuationDissipation}.
1782   \end{equation}
1783 + In effect, it acts as a constraint on the possible ways in which one
1784 + can model the random force and friction kernel.
1785  
674 \section{\label{introSection:hydroynamics}Hydrodynamics}
675
1786   \subsection{\label{introSection:frictionTensor} Friction Tensor}
1787 < \subsection{\label{introSection:analyticalApproach}Analytical
1788 < Approach}
1787 > Theoretically, the friction kernel can be determined using velocity
1788 > autocorrelation function. However, this approach become impractical
1789 > when the system become more and more complicate. Instead, various
1790 > approaches based on hydrodynamics have been developed to calculate
1791 > the friction coefficients. The friction effect is isotropic in
1792 > Equation, $\zeta$ can be taken as a scalar. In general, friction
1793 > tensor $\Xi$ is a $6\times 6$ matrix given by
1794 > \[
1795 > \Xi  = \left( {\begin{array}{*{20}c}
1796 >   {\Xi _{}^{tt} } & {\Xi _{}^{rt} }  \\
1797 >   {\Xi _{}^{tr} } & {\Xi _{}^{rr} }  \\
1798 > \end{array}} \right).
1799 > \]
1800 > Here, $ {\Xi^{tt} }$ and $ {\Xi^{rr} }$ are translational friction
1801 > tensor and rotational resistance (friction) tensor respectively,
1802 > while ${\Xi^{tr} }$ is translation-rotation coupling tensor and $
1803 > {\Xi^{rt} }$ is rotation-translation coupling tensor. When a
1804 > particle moves in a fluid, it may experience friction force or
1805 > torque along the opposite direction of the velocity or angular
1806 > velocity,
1807 > \[
1808 > \left( \begin{array}{l}
1809 > F_R  \\
1810 > \tau _R  \\
1811 > \end{array} \right) =  - \left( {\begin{array}{*{20}c}
1812 >   {\Xi ^{tt} } & {\Xi ^{rt} }  \\
1813 >   {\Xi ^{tr} } & {\Xi ^{rr} }  \\
1814 > \end{array}} \right)\left( \begin{array}{l}
1815 > v \\
1816 > w \\
1817 > \end{array} \right)
1818 > \]
1819 > where $F_r$ is the friction force and $\tau _R$ is the friction
1820 > toque.
1821  
1822 < \subsection{\label{introSection:approximationApproach}Approximation
681 < Approach}
1822 > \subsubsection{\label{introSection:resistanceTensorRegular}The Resistance Tensor for Regular Shape}
1823  
1824 < \subsection{\label{introSection:centersRigidBody}Centers of Rigid
1825 < Body}
1824 > For a spherical particle, the translational and rotational friction
1825 > constant can be calculated from Stoke's law,
1826 > \[
1827 > \Xi ^{tt}  = \left( {\begin{array}{*{20}c}
1828 >   {6\pi \eta R} & 0 & 0  \\
1829 >   0 & {6\pi \eta R} & 0  \\
1830 >   0 & 0 & {6\pi \eta R}  \\
1831 > \end{array}} \right)
1832 > \]
1833 > and
1834 > \[
1835 > \Xi ^{rr}  = \left( {\begin{array}{*{20}c}
1836 >   {8\pi \eta R^3 } & 0 & 0  \\
1837 >   0 & {8\pi \eta R^3 } & 0  \\
1838 >   0 & 0 & {8\pi \eta R^3 }  \\
1839 > \end{array}} \right)
1840 > \]
1841 > where $\eta$ is the viscosity of the solvent and $R$ is the
1842 > hydrodynamics radius.
1843 >
1844 > Other non-spherical shape, such as cylinder and ellipsoid
1845 > \textit{etc}, are widely used as reference for developing new
1846 > hydrodynamics theory, because their properties can be calculated
1847 > exactly. In 1936, Perrin extended Stokes's law to general ellipsoid,
1848 > also called a triaxial ellipsoid, which is given in Cartesian
1849 > coordinates by
1850 > \[
1851 > \frac{{x^2 }}{{a^2 }} + \frac{{y^2 }}{{b^2 }} + \frac{{z^2 }}{{c^2
1852 > }} = 1
1853 > \]
1854 > where the semi-axes are of lengths $a$, $b$, and $c$. Unfortunately,
1855 > due to the complexity of the elliptic integral, only the ellipsoid
1856 > with the restriction of two axes having to be equal, \textit{i.e.}
1857 > prolate($ a \ge b = c$) and oblate ($ a < b = c $), can be solved
1858 > exactly. Introducing an elliptic integral parameter $S$ for prolate,
1859 > \[
1860 > S = \frac{2}{{\sqrt {a^2  - b^2 } }}\ln \frac{{a + \sqrt {a^2  - b^2
1861 > } }}{b},
1862 > \]
1863 > and oblate,
1864 > \[
1865 > S = \frac{2}{{\sqrt {b^2  - a^2 } }}arctg\frac{{\sqrt {b^2  - a^2 }
1866 > }}{a}
1867 > \],
1868 > one can write down the translational and rotational resistance
1869 > tensors
1870 > \[
1871 > \begin{array}{l}
1872 > \Xi _a^{tt}  = 16\pi \eta \frac{{a^2  - b^2 }}{{(2a^2  - b^2 )S - 2a}} \\
1873 > \Xi _b^{tt}  = \Xi _c^{tt}  = 32\pi \eta \frac{{a^2  - b^2 }}{{(2a^2  - 3b^2 )S + 2a}} \\
1874 > \end{array},
1875 > \]
1876 > and
1877 > \[
1878 > \begin{array}{l}
1879 > \Xi _a^{rr}  = \frac{{32\pi }}{3}\eta \frac{{(a^2  - b^2 )b^2 }}{{2a - b^2 S}} \\
1880 > \Xi _b^{rr}  = \Xi _c^{rr}  = \frac{{32\pi }}{3}\eta \frac{{(a^4  - b^4 )}}{{(2a^2  - b^2 )S - 2a}} \\
1881 > \end{array}.
1882 > \]
1883 >
1884 > \subsubsection{\label{introSection:resistanceTensorRegularArbitrary}The Resistance Tensor for Arbitrary Shape}
1885 >
1886 > Unlike spherical and other regular shaped molecules, there is not
1887 > analytical solution for friction tensor of any arbitrary shaped
1888 > rigid molecules. The ellipsoid of revolution model and general
1889 > triaxial ellipsoid model have been used to approximate the
1890 > hydrodynamic properties of rigid bodies. However, since the mapping
1891 > from all possible ellipsoidal space, $r$-space, to all possible
1892 > combination of rotational diffusion coefficients, $D$-space is not
1893 > unique\cite{Wegener79} as well as the intrinsic coupling between
1894 > translational and rotational motion of rigid body\cite{}, general
1895 > ellipsoid is not always suitable for modeling arbitrarily shaped
1896 > rigid molecule. A number of studies have been devoted to determine
1897 > the friction tensor for irregularly shaped rigid bodies using more
1898 > advanced method\cite{} where the molecule of interest was modeled by
1899 > combinations of spheres(beads)\cite{} and the hydrodynamics
1900 > properties of the molecule can be calculated using the hydrodynamic
1901 > interaction tensor. Let us consider a rigid assembly of $N$ beads
1902 > immersed in a continuous medium. Due to hydrodynamics interaction,
1903 > the ``net'' velocity of $i$th bead, $v'_i$ is different than its
1904 > unperturbed velocity $v_i$,
1905 > \[
1906 > v'_i  = v_i  - \sum\limits_{j \ne i} {T_{ij} F_j }
1907 > \]
1908 > where $F_i$ is the frictional force, and $T_{ij}$ is the
1909 > hydrodynamic interaction tensor. The friction force of $i$th bead is
1910 > proportional to its ``net'' velocity
1911 > \begin{equation}
1912 > F_i  = \zeta _i v_i  - \zeta _i \sum\limits_{j \ne i} {T_{ij} F_j }.
1913 > \label{introEquation:tensorExpression}
1914 > \end{equation}
1915 > This equation is the basis for deriving the hydrodynamic tensor. In
1916 > 1930, Oseen and Burgers gave a simple solution to Equation
1917 > \ref{introEquation:tensorExpression}
1918 > \begin{equation}
1919 > T_{ij}  = \frac{1}{{8\pi \eta r_{ij} }}\left( {I + \frac{{R_{ij}
1920 > R_{ij}^T }}{{R_{ij}^2 }}} \right).
1921 > \label{introEquation:oseenTensor}
1922 > \end{equation}
1923 > Here $R_{ij}$ is the distance vector between bead $i$ and bead $j$.
1924 > A second order expression for element of different size was
1925 > introduced by Rotne and Prager\cite{} and improved by Garc\'{i}a de
1926 > la Torre and Bloomfield,
1927 > \begin{equation}
1928 > T_{ij}  = \frac{1}{{8\pi \eta R_{ij} }}\left[ {\left( {I +
1929 > \frac{{R_{ij} R_{ij}^T }}{{R_{ij}^2 }}} \right) + R\frac{{\sigma
1930 > _i^2  + \sigma _j^2 }}{{r_{ij}^2 }}\left( {\frac{I}{3} -
1931 > \frac{{R_{ij} R_{ij}^T }}{{R_{ij}^2 }}} \right)} \right].
1932 > \label{introEquation:RPTensorNonOverlapped}
1933 > \end{equation}
1934 > Both of the Equation \ref{introEquation:oseenTensor} and Equation
1935 > \ref{introEquation:RPTensorNonOverlapped} have an assumption $R_{ij}
1936 > \ge \sigma _i  + \sigma _j$. An alternative expression for
1937 > overlapping beads with the same radius, $\sigma$, is given by
1938 > \begin{equation}
1939 > T_{ij}  = \frac{1}{{6\pi \eta R_{ij} }}\left[ {\left( {1 -
1940 > \frac{2}{{32}}\frac{{R_{ij} }}{\sigma }} \right)I +
1941 > \frac{2}{{32}}\frac{{R_{ij} R_{ij}^T }}{{R_{ij} \sigma }}} \right]
1942 > \label{introEquation:RPTensorOverlapped}
1943 > \end{equation}
1944 >
1945 > To calculate the resistance tensor at an arbitrary origin $O$, we
1946 > construct a $3N \times 3N$ matrix consisting of $N \times N$
1947 > $B_{ij}$ blocks
1948 > \begin{equation}
1949 > B = \left( {\begin{array}{*{20}c}
1950 >   {B_{11} } &  \ldots  & {B_{1N} }  \\
1951 >    \vdots  &  \ddots  &  \vdots   \\
1952 >   {B_{N1} } &  \cdots  & {B_{NN} }  \\
1953 > \end{array}} \right),
1954 > \end{equation}
1955 > where $B_{ij}$ is given by
1956 > \[
1957 > B_{ij}  = \delta _{ij} \frac{I}{{6\pi \eta R}} + (1 - \delta _{ij}
1958 > )T_{ij}
1959 > \]
1960 > where $\delta _{ij}$ is Kronecker delta function. Inverting matrix
1961 > $B$, we obtain
1962 >
1963 > \[
1964 > C = B^{ - 1}  = \left( {\begin{array}{*{20}c}
1965 >   {C_{11} } &  \ldots  & {C_{1N} }  \\
1966 >    \vdots  &  \ddots  &  \vdots   \\
1967 >   {C_{N1} } &  \cdots  & {C_{NN} }  \\
1968 > \end{array}} \right)
1969 > \]
1970 > , which can be partitioned into $N \times N$ $3 \times 3$ block
1971 > $C_{ij}$. With the help of $C_{ij}$ and skew matrix $U_i$
1972 > \[
1973 > U_i  = \left( {\begin{array}{*{20}c}
1974 >   0 & { - z_i } & {y_i }  \\
1975 >   {z_i } & 0 & { - x_i }  \\
1976 >   { - y_i } & {x_i } & 0  \\
1977 > \end{array}} \right)
1978 > \]
1979 > where $x_i$, $y_i$, $z_i$ are the components of the vector joining
1980 > bead $i$ and origin $O$. Hence, the elements of resistance tensor at
1981 > arbitrary origin $O$ can be written as
1982 > \begin{equation}
1983 > \begin{array}{l}
1984 > \Xi _{}^{tt}  = \sum\limits_i {\sum\limits_j {C_{ij} } } , \\
1985 > \Xi _{}^{tr}  = \Xi _{}^{rt}  = \sum\limits_i {\sum\limits_j {U_i C_{ij} } } , \\
1986 > \Xi _{}^{rr}  =  - \sum\limits_i {\sum\limits_j {U_i C_{ij} } } U_j  \\
1987 > \end{array}
1988 > \label{introEquation:ResistanceTensorArbitraryOrigin}
1989 > \end{equation}
1990 >
1991 > The resistance tensor depends on the origin to which they refer. The
1992 > proper location for applying friction force is the center of
1993 > resistance (reaction), at which the trace of rotational resistance
1994 > tensor, $ \Xi ^{rr}$ reaches minimum. Mathematically, the center of
1995 > resistance is defined as an unique point of the rigid body at which
1996 > the translation-rotation coupling tensor are symmetric,
1997 > \begin{equation}
1998 > \Xi^{tr}  = \left( {\Xi^{tr} } \right)^T
1999 > \label{introEquation:definitionCR}
2000 > \end{equation}
2001 > Form Equation \ref{introEquation:ResistanceTensorArbitraryOrigin},
2002 > we can easily find out that the translational resistance tensor is
2003 > origin independent, while the rotational resistance tensor and
2004 > translation-rotation coupling resistance tensor depend on the
2005 > origin. Given resistance tensor at an arbitrary origin $O$, and a
2006 > vector ,$r_{OP}(x_{OP}, y_{OP}, z_{OP})$, from $O$ to $P$, we can
2007 > obtain the resistance tensor at $P$ by
2008 > \begin{equation}
2009 > \begin{array}{l}
2010 > \Xi _P^{tt}  = \Xi _O^{tt}  \\
2011 > \Xi _P^{tr}  = \Xi _P^{rt}  = \Xi _O^{tr}  - U_{OP} \Xi _O^{tt}  \\
2012 > \Xi _P^{rr}  = \Xi _O^{rr}  - U_{OP} \Xi _O^{tt} U_{OP}  + \Xi _O^{tr} U_{OP}  - U_{OP} \Xi _O^{tr} ^{^T }  \\
2013 > \end{array}
2014 > \label{introEquation:resistanceTensorTransformation}
2015 > \end{equation}
2016 > where
2017 > \[
2018 > U_{OP}  = \left( {\begin{array}{*{20}c}
2019 >   0 & { - z_{OP} } & {y_{OP} }  \\
2020 >   {z_i } & 0 & { - x_{OP} }  \\
2021 >   { - y_{OP} } & {x_{OP} } & 0  \\
2022 > \end{array}} \right)
2023 > \]
2024 > Using Equations \ref{introEquation:definitionCR} and
2025 > \ref{introEquation:resistanceTensorTransformation}, one can locate
2026 > the position of center of resistance,
2027 > \[
2028 > \left( \begin{array}{l}
2029 > x_{OR}  \\
2030 > y_{OR}  \\
2031 > z_{OR}  \\
2032 > \end{array} \right) = \left( {\begin{array}{*{20}c}
2033 >   {(\Xi _O^{rr} )_{yy}  + (\Xi _O^{rr} )_{zz} } & { - (\Xi _O^{rr} )_{xy} } & { - (\Xi _O^{rr} )_{xz} }  \\
2034 >   { - (\Xi _O^{rr} )_{xy} } & {(\Xi _O^{rr} )_{zz}  + (\Xi _O^{rr} )_{xx} } & { - (\Xi _O^{rr} )_{yz} }  \\
2035 >   { - (\Xi _O^{rr} )_{xz} } & { - (\Xi _O^{rr} )_{yz} } & {(\Xi _O^{rr} )_{xx}  + (\Xi _O^{rr} )_{yy} }  \\
2036 > \end{array}} \right)^{ - 1} \left( \begin{array}{l}
2037 > (\Xi _O^{tr} )_{yz}  - (\Xi _O^{tr} )_{zy}  \\
2038 > (\Xi _O^{tr} )_{zx}  - (\Xi _O^{tr} )_{xz}  \\
2039 > (\Xi _O^{tr} )_{xy}  - (\Xi _O^{tr} )_{yx}  \\
2040 > \end{array} \right).
2041 > \]
2042 > where $x_OR$, $y_OR$, $z_OR$ are the components of the vector
2043 > joining center of resistance $R$ and origin $O$.

Diff Legend

Removed lines
+ Added lines
< Changed lines
> Changed lines