ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/tengDissertation/Introduction.tex
Revision: 2794
Committed: Tue Jun 6 01:58:27 2006 UTC (18 years ago) by tim
Content type: application/x-tex
File size: 87545 byte(s)
Log Message:
minor fixes

File Contents

# User Rev Content
1 tim 2685 \chapter{\label{chapt:introduction}INTRODUCTION AND THEORETICAL BACKGROUND}
2    
3 tim 2693 \section{\label{introSection:classicalMechanics}Classical
4     Mechanics}
5 tim 2685
6 tim 2692 Closely related to Classical Mechanics, Molecular Dynamics
7     simulations are carried out by integrating the equations of motion
8     for a given system of particles. There are three fundamental ideas
9     behind classical mechanics. Firstly, One can determine the state of
10     a mechanical system at any time of interest; Secondly, all the
11     mechanical properties of the system at that time can be determined
12     by combining the knowledge of the properties of the system with the
13     specification of this state; Finally, the specification of the state
14     when further combine with the laws of mechanics will also be
15     sufficient to predict the future behavior of the system.
16 tim 2685
17 tim 2693 \subsection{\label{introSection:newtonian}Newtonian Mechanics}
18 tim 2694 The discovery of Newton's three laws of mechanics which govern the
19     motion of particles is the foundation of the classical mechanics.
20     Newton¡¯s first law defines a class of inertial frames. Inertial
21     frames are reference frames where a particle not interacting with
22     other bodies will move with constant speed in the same direction.
23     With respect to inertial frames Newton¡¯s second law has the form
24     \begin{equation}
25     F = \frac {dp}{dt} = \frac {mv}{dt}
26     \label{introEquation:newtonSecondLaw}
27     \end{equation}
28     A point mass interacting with other bodies moves with the
29     acceleration along the direction of the force acting on it. Let
30 tim 2702 $F_{ij}$ 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 tim 2694 Newton¡¯s third law states that
33     \begin{equation}
34 tim 2702 F_{ij} = -F_{ji}
35 tim 2694 \label{introEquation:newtonThirdLaw}
36     \end{equation}
37 tim 2692
38 tim 2694 Conservation laws of Newtonian Mechanics play very important roles
39     in solving mechanics problems. The linear momentum of a particle is
40     conserved if it is free or it experiences no force. The second
41     conservation theorem concerns the angular momentum of a particle.
42     The angular momentum $L$ of a particle with respect to an origin
43     from which $r$ is measured is defined to be
44     \begin{equation}
45     L \equiv r \times p \label{introEquation:angularMomentumDefinition}
46     \end{equation}
47     The torque $\tau$ with respect to the same origin is defined to be
48     \begin{equation}
49     N \equiv r \times F \label{introEquation:torqueDefinition}
50     \end{equation}
51     Differentiating Eq.~\ref{introEquation:angularMomentumDefinition},
52     \[
53     \dot L = \frac{d}{{dt}}(r \times p) = (\dot r \times p) + (r \times
54     \dot p)
55     \]
56     since
57     \[
58     \dot r \times p = \dot r \times mv = m\dot r \times \dot r \equiv 0
59     \]
60     thus,
61     \begin{equation}
62     \dot L = r \times \dot p = N
63     \end{equation}
64     If there are no external torques acting on a body, the angular
65     momentum of it is conserved. The last conservation theorem state
66 tim 2696 that if all forces are conservative, Energy
67     \begin{equation}E = T + V \label{introEquation:energyConservation}
68     \end{equation}
69     is conserved. All of these conserved quantities are
70     important factors to determine the quality of numerical integration
71     scheme for rigid body \cite{Dullweber1997}.
72 tim 2694
73 tim 2693 \subsection{\label{introSection:lagrangian}Lagrangian Mechanics}
74 tim 2692
75     Newtonian Mechanics suffers from two important limitations: it
76     describes their motion in special cartesian coordinate systems.
77     Another limitation of Newtonian mechanics becomes obvious when we
78     try to describe systems with large numbers of particles. It becomes
79     very difficult to predict the properties of the system by carrying
80     out calculations involving the each individual interaction between
81     all the particles, even if we know all of the details of the
82     interaction. In order to overcome some of the practical difficulties
83     which arise in attempts to apply Newton's equation to complex
84     system, alternative procedures may be developed.
85    
86 tim 2694 \subsubsection{\label{introSection:halmiltonPrinciple}Hamilton's
87 tim 2692 Principle}
88    
89     Hamilton introduced the dynamical principle upon which it is
90     possible to base all of mechanics and, indeed, most of classical
91     physics. Hamilton's Principle may be stated as follow,
92    
93     The actual trajectory, along which a dynamical system may move from
94     one point to another within a specified time, is derived by finding
95     the path which minimizes the time integral of the difference between
96 tim 2786 the kinetic, $K$, and potential energies, $U$ \cite{Tolman1979}.
97 tim 2692 \begin{equation}
98     \delta \int_{t_1 }^{t_2 } {(K - U)dt = 0} ,
99 tim 2693 \label{introEquation:halmitonianPrinciple1}
100 tim 2692 \end{equation}
101    
102     For simple mechanical systems, where the forces acting on the
103     different part are derivable from a potential and the velocities are
104     small compared with that of light, the Lagrangian function $L$ can
105     be define as the difference between the kinetic energy of the system
106     and its potential energy,
107     \begin{equation}
108     L \equiv K - U = L(q_i ,\dot q_i ) ,
109     \label{introEquation:lagrangianDef}
110     \end{equation}
111     then Eq.~\ref{introEquation:halmitonianPrinciple1} becomes
112     \begin{equation}
113 tim 2693 \delta \int_{t_1 }^{t_2 } {L dt = 0} ,
114     \label{introEquation:halmitonianPrinciple2}
115 tim 2692 \end{equation}
116    
117 tim 2694 \subsubsection{\label{introSection:equationOfMotionLagrangian}The
118 tim 2692 Equations of Motion in Lagrangian Mechanics}
119    
120 tim 2700 For a holonomic system of $f$ degrees of freedom, the equations of
121 tim 2692 motion in the Lagrangian form is
122     \begin{equation}
123     \frac{d}{{dt}}\frac{{\partial L}}{{\partial \dot q_i }} -
124     \frac{{\partial L}}{{\partial q_i }} = 0,{\rm{ }}i = 1, \ldots,f
125 tim 2693 \label{introEquation:eqMotionLagrangian}
126 tim 2692 \end{equation}
127     where $q_{i}$ is generalized coordinate and $\dot{q_{i}}$ is
128     generalized velocity.
129    
130 tim 2693 \subsection{\label{introSection:hamiltonian}Hamiltonian Mechanics}
131 tim 2692
132     Arising from Lagrangian Mechanics, Hamiltonian Mechanics was
133     introduced by William Rowan Hamilton in 1833 as a re-formulation of
134     classical mechanics. If the potential energy of a system is
135     independent of generalized velocities, the generalized momenta can
136     be defined as
137     \begin{equation}
138     p_i = \frac{\partial L}{\partial \dot q_i}
139     \label{introEquation:generalizedMomenta}
140     \end{equation}
141 tim 2693 The Lagrange equations of motion are then expressed by
142 tim 2692 \begin{equation}
143 tim 2693 p_i = \frac{{\partial L}}{{\partial q_i }}
144     \label{introEquation:generalizedMomentaDot}
145     \end{equation}
146    
147     With the help of the generalized momenta, we may now define a new
148     quantity $H$ by the equation
149     \begin{equation}
150     H = \sum\limits_k {p_k \dot q_k } - L ,
151 tim 2692 \label{introEquation:hamiltonianDefByLagrangian}
152     \end{equation}
153     where $ \dot q_1 \ldots \dot q_f $ are generalized velocities and
154     $L$ is the Lagrangian function for the system.
155    
156 tim 2693 Differentiating Eq.~\ref{introEquation:hamiltonianDefByLagrangian},
157     one can obtain
158     \begin{equation}
159     dH = \sum\limits_k {\left( {p_k d\dot q_k + \dot q_k dp_k -
160     \frac{{\partial L}}{{\partial q_k }}dq_k - \frac{{\partial
161     L}}{{\partial \dot q_k }}d\dot q_k } \right)} - \frac{{\partial
162     L}}{{\partial t}}dt \label{introEquation:diffHamiltonian1}
163     \end{equation}
164     Making use of Eq.~\ref{introEquation:generalizedMomenta}, the
165     second and fourth terms in the parentheses cancel. Therefore,
166     Eq.~\ref{introEquation:diffHamiltonian1} can be rewritten as
167     \begin{equation}
168     dH = \sum\limits_k {\left( {\dot q_k dp_k - \dot p_k dq_k }
169     \right)} - \frac{{\partial L}}{{\partial t}}dt
170     \label{introEquation:diffHamiltonian2}
171     \end{equation}
172     By identifying the coefficients of $dq_k$, $dp_k$ and dt, we can
173     find
174     \begin{equation}
175     \frac{{\partial H}}{{\partial p_k }} = q_k
176     \label{introEquation:motionHamiltonianCoordinate}
177     \end{equation}
178     \begin{equation}
179     \frac{{\partial H}}{{\partial q_k }} = - p_k
180     \label{introEquation:motionHamiltonianMomentum}
181     \end{equation}
182     and
183     \begin{equation}
184     \frac{{\partial H}}{{\partial t}} = - \frac{{\partial L}}{{\partial
185     t}}
186     \label{introEquation:motionHamiltonianTime}
187     \end{equation}
188    
189     Eq.~\ref{introEquation:motionHamiltonianCoordinate} and
190     Eq.~\ref{introEquation:motionHamiltonianMomentum} are Hamilton's
191     equation of motion. Due to their symmetrical formula, they are also
192 tim 2786 known as the canonical equations of motions \cite{Goldstein2001}.
193 tim 2693
194 tim 2692 An important difference between Lagrangian approach and the
195     Hamiltonian approach is that the Lagrangian is considered to be a
196     function of the generalized velocities $\dot q_i$ and the
197     generalized coordinates $q_i$, while the Hamiltonian is considered
198     to be a function of the generalized momenta $p_i$ and the conjugate
199     generalized coordinate $q_i$. Hamiltonian Mechanics is more
200     appropriate for application to statistical mechanics and quantum
201     mechanics, since it treats the coordinate and its time derivative as
202     independent variables and it only works with 1st-order differential
203 tim 2786 equations\cite{Marion1990}.
204 tim 2692
205 tim 2696 In Newtonian Mechanics, a system described by conservative forces
206     conserves the total energy \ref{introEquation:energyConservation}.
207     It follows that Hamilton's equations of motion conserve the total
208     Hamiltonian.
209     \begin{equation}
210     \frac{{dH}}{{dt}} = \sum\limits_i {\left( {\frac{{\partial
211     H}}{{\partial q_i }}\dot q_i + \frac{{\partial H}}{{\partial p_i
212     }}\dot p_i } \right)} = \sum\limits_i {\left( {\frac{{\partial
213     H}}{{\partial q_i }}\frac{{\partial H}}{{\partial p_i }} -
214     \frac{{\partial H}}{{\partial p_i }}\frac{{\partial H}}{{\partial
215 tim 2698 q_i }}} \right) = 0} \label{introEquation:conserveHalmitonian}
216 tim 2696 \end{equation}
217    
218 tim 2693 \section{\label{introSection:statisticalMechanics}Statistical
219     Mechanics}
220 tim 2692
221 tim 2694 The thermodynamic behaviors and properties of Molecular Dynamics
222 tim 2692 simulation are governed by the principle of Statistical Mechanics.
223     The following section will give a brief introduction to some of the
224 tim 2700 Statistical Mechanics concepts and theorem presented in this
225     dissertation.
226 tim 2692
227 tim 2700 \subsection{\label{introSection:ensemble}Phase Space and Ensemble}
228 tim 2692
229 tim 2700 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 tim 2706 \Omega (N,V,E) = e^{\beta TS} \label{introEquation:NVEPartition}.
319 tim 2700 \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 tim 2702 \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 tim 2700 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 tim 2693 \subsection{\label{introSection:ergodic}The Ergodic Hypothesis}
462 tim 2692
463 tim 2695 Various thermodynamic properties can be calculated from Molecular
464     Dynamics simulation. By comparing experimental values with the
465     calculated properties, one can determine the accuracy of the
466     simulation and the quality of the underlying model. However, both of
467     experiment and computer simulation are usually performed during a
468     certain time interval and the measurements are averaged over a
469     period of them which is different from the average behavior of
470     many-body system in Statistical Mechanics. Fortunately, Ergodic
471     Hypothesis is proposed to make a connection between time average and
472     ensemble average. It states that time average and average over the
473 tim 2786 statistical ensemble are identical \cite{Frenkel1996, Leach2001}.
474 tim 2695 \begin{equation}
475 tim 2700 \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 tim 2695 \end{equation}
479 tim 2700 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 tim 2786 reasonable, the Monte Carlo techniques\cite{Metropolis1949} can be
488 tim 2700 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 tim 2694
493 tim 2697 \section{\label{introSection:geometricIntegratos}Geometric Integrators}
494     A variety of numerical integrators were proposed to simulate the
495     motions. They usually begin with an initial conditionals and move
496     the objects in the direction governed by the differential equations.
497     However, most of them ignore the hidden physical law contained
498     within the equations. Since 1990, geometric integrators, which
499     preserve various phase-flow invariants such as symplectic structure,
500     volume and time reversal symmetry, are developed to address this
501 tim 2789 issue\cite{Dullweber1997, McLachlan1998, Leimkuhler1999}. The
502     velocity verlet method, which happens to be a simple example of
503     symplectic integrator, continues to gain its popularity in molecular
504     dynamics community. This fact can be partly explained by its
505     geometric nature.
506 tim 2697
507     \subsection{\label{introSection:symplecticManifold}Symplectic Manifold}
508     A \emph{manifold} is an abstract mathematical space. It locally
509     looks like Euclidean space, but when viewed globally, it may have
510     more complicate structure. A good example of manifold is the surface
511     of Earth. It seems to be flat locally, but it is round if viewed as
512     a whole. A \emph{differentiable manifold} (also known as
513     \emph{smooth manifold}) is a manifold with an open cover in which
514     the covering neighborhoods are all smoothly isomorphic to one
515     another. In other words,it is possible to apply calculus on
516     \emph{differentiable manifold}. A \emph{symplectic manifold} is
517     defined as a pair $(M, \omega)$ which consisting of a
518     \emph{differentiable manifold} $M$ and a close, non-degenerated,
519     bilinear symplectic form, $\omega$. A symplectic form on a vector
520     space $V$ is a function $\omega(x, y)$ which satisfies
521     $\omega(\lambda_1x_1+\lambda_2x_2, y) = \lambda_1\omega(x_1, y)+
522     \lambda_2\omega(x_2, y)$, $\omega(x, y) = - \omega(y, x)$ and
523     $\omega(x, x) = 0$. Cross product operation in vector field is an
524     example of symplectic form.
525    
526     One of the motivations to study \emph{symplectic manifold} in
527     Hamiltonian Mechanics is that a symplectic manifold can represent
528     all possible configurations of the system and the phase space of the
529     system can be described by it's cotangent bundle. Every symplectic
530     manifold is even dimensional. For instance, in Hamilton equations,
531     coordinate and momentum always appear in pairs.
532    
533     Let $(M,\omega)$ and $(N, \eta)$ be symplectic manifolds. A map
534     \[
535     f : M \rightarrow N
536     \]
537     is a \emph{symplectomorphism} if it is a \emph{diffeomorphims} and
538     the \emph{pullback} of $\eta$ under f is equal to $\omega$.
539     Canonical transformation is an example of symplectomorphism in
540 tim 2698 classical mechanics.
541 tim 2697
542 tim 2698 \subsection{\label{introSection:ODE}Ordinary Differential Equations}
543 tim 2697
544 tim 2698 For a ordinary differential system defined as
545     \begin{equation}
546     \dot x = f(x)
547     \end{equation}
548     where $x = x(q,p)^T$, this system is canonical Hamiltonian, if
549     \begin{equation}
550 tim 2699 f(r) = J\nabla _x H(r).
551 tim 2698 \end{equation}
552     $H = H (q, p)$ is Hamiltonian function and $J$ is the skew-symmetric
553     matrix
554     \begin{equation}
555     J = \left( {\begin{array}{*{20}c}
556     0 & I \\
557     { - I} & 0 \\
558     \end{array}} \right)
559     \label{introEquation:canonicalMatrix}
560     \end{equation}
561     where $I$ is an identity matrix. Using this notation, Hamiltonian
562     system can be rewritten as,
563     \begin{equation}
564     \frac{d}{{dt}}x = J\nabla _x H(x)
565     \label{introEquation:compactHamiltonian}
566     \end{equation}In this case, $f$ is
567     called a \emph{Hamiltonian vector field}.
568 tim 2697
569 tim 2789 Another generalization of Hamiltonian dynamics is Poisson
570     Dynamics\cite{Olver1986},
571 tim 2698 \begin{equation}
572     \dot x = J(x)\nabla _x H \label{introEquation:poissonHamiltonian}
573     \end{equation}
574     The most obvious change being that matrix $J$ now depends on $x$.
575    
576 tim 2702 \subsection{\label{introSection:exactFlow}Exact Flow}
577    
578 tim 2698 Let $x(t)$ be the exact solution of the ODE system,
579     \begin{equation}
580     \frac{{dx}}{{dt}} = f(x) \label{introEquation:ODE}
581     \end{equation}
582     The exact flow(solution) $\varphi_\tau$ is defined by
583     \[
584     x(t+\tau) =\varphi_\tau(x(t))
585     \]
586     where $\tau$ is a fixed time step and $\varphi$ is a map from phase
587 tim 2702 space to itself. The flow has the continuous group property,
588 tim 2698 \begin{equation}
589 tim 2702 \varphi _{\tau _1 } \circ \varphi _{\tau _2 } = \varphi _{\tau _1
590     + \tau _2 } .
591     \end{equation}
592     In particular,
593     \begin{equation}
594     \varphi _\tau \circ \varphi _{ - \tau } = I
595     \end{equation}
596     Therefore, the exact flow is self-adjoint,
597     \begin{equation}
598     \varphi _\tau = \varphi _{ - \tau }^{ - 1}.
599     \end{equation}
600     The exact flow can also be written in terms of the of an operator,
601     \begin{equation}
602     \varphi _\tau (x) = e^{\tau \sum\limits_i {f_i (x)\frac{\partial
603     }{{\partial x_i }}} } (x) \equiv \exp (\tau f)(x).
604     \label{introEquation:exponentialOperator}
605     \end{equation}
606    
607     In most cases, it is not easy to find the exact flow $\varphi_\tau$.
608     Instead, we use a approximate map, $\psi_\tau$, which is usually
609     called integrator. The order of an integrator $\psi_\tau$ is $p$, if
610     the Taylor series of $\psi_\tau$ agree to order $p$,
611     \begin{equation}
612 tim 2698 \psi_tau(x) = x + \tau f(x) + O(\tau^{p+1})
613     \end{equation}
614    
615 tim 2702 \subsection{\label{introSection:geometricProperties}Geometric Properties}
616    
617 tim 2789 The hidden geometric properties\cite{Budd1999, Marsden1998} of ODE
618     and its flow play important roles in numerical studies. Many of them
619     can be found in systems which occur naturally in applications.
620 tim 2702
621     Let $\varphi$ be the flow of Hamiltonian vector field, $\varphi$ is
622     a \emph{symplectic} flow if it satisfies,
623 tim 2698 \begin{equation}
624 tim 2703 {\varphi '}^T J \varphi ' = J.
625 tim 2698 \end{equation}
626     According to Liouville's theorem, the symplectic volume is invariant
627     under a Hamiltonian flow, which is the basis for classical
628 tim 2699 statistical mechanics. Furthermore, the flow of a Hamiltonian vector
629     field on a symplectic manifold can be shown to be a
630     symplectomorphism. As to the Poisson system,
631 tim 2698 \begin{equation}
632 tim 2703 {\varphi '}^T J \varphi ' = J \circ \varphi
633 tim 2698 \end{equation}
634 tim 2702 is the property must be preserved by the integrator.
635    
636     It is possible to construct a \emph{volume-preserving} flow for a
637     source free($ \nabla \cdot f = 0 $) ODE, if the flow satisfies $
638     \det d\varphi = 1$. One can show easily that a symplectic flow will
639     be volume-preserving.
640    
641     Changing the variables $y = h(x)$ in a ODE\ref{introEquation:ODE}
642     will result in a new system,
643 tim 2698 \[
644     \dot y = \tilde f(y) = ((dh \cdot f)h^{ - 1} )(y).
645     \]
646     The vector filed $f$ has reversing symmetry $h$ if $f = - \tilde f$.
647     In other words, the flow of this vector field is reversible if and
648 tim 2702 only if $ h \circ \varphi ^{ - 1} = \varphi \circ h $.
649 tim 2698
650 tim 2705 A \emph{first integral}, or conserved quantity of a general
651     differential function is a function $ G:R^{2d} \to R^d $ which is
652     constant for all solutions of the ODE $\frac{{dx}}{{dt}} = f(x)$ ,
653     \[
654     \frac{{dG(x(t))}}{{dt}} = 0.
655     \]
656     Using chain rule, one may obtain,
657     \[
658     \sum\limits_i {\frac{{dG}}{{dx_i }}} f_i (x) = f \bullet \nabla G,
659     \]
660     which is the condition for conserving \emph{first integral}. For a
661     canonical Hamiltonian system, the time evolution of an arbitrary
662     smooth function $G$ is given by,
663 tim 2789
664     \begin{eqnarray}
665     \frac{{dG(x(t))}}{{dt}} & = & [\nabla _x G(x(t))]^T \dot x(t) \\
666     & = & [\nabla _x G(x(t))]^T J\nabla _x H(x(t)). \\
667 tim 2705 \label{introEquation:firstIntegral1}
668 tim 2789 \end{eqnarray}
669    
670    
671 tim 2705 Using poisson bracket notion, Equation
672     \ref{introEquation:firstIntegral1} can be rewritten as
673     \[
674     \frac{d}{{dt}}G(x(t)) = \left\{ {G,H} \right\}(x(t)).
675     \]
676     Therefore, the sufficient condition for $G$ to be the \emph{first
677     integral} of a Hamiltonian system is
678     \[
679     \left\{ {G,H} \right\} = 0.
680     \]
681     As well known, the Hamiltonian (or energy) H of a Hamiltonian system
682     is a \emph{first integral}, which is due to the fact $\{ H,H\} =
683     0$.
684    
685 tim 2789 When designing any numerical methods, one should always try to
686 tim 2702 preserve the structural properties of the original ODE and its flow.
687    
688 tim 2699 \subsection{\label{introSection:constructionSymplectic}Construction of Symplectic Methods}
689     A lot of well established and very effective numerical methods have
690     been successful precisely because of their symplecticities even
691     though this fact was not recognized when they were first
692     constructed. The most famous example is leapfrog methods in
693     molecular dynamics. In general, symplectic integrators can be
694     constructed using one of four different methods.
695     \begin{enumerate}
696     \item Generating functions
697     \item Variational methods
698     \item Runge-Kutta methods
699     \item Splitting methods
700     \end{enumerate}
701 tim 2698
702 tim 2789 Generating function\cite{Channell1990} tends to lead to methods
703     which are cumbersome and difficult to use. In dissipative systems,
704     variational methods can capture the decay of energy
705     accurately\cite{Kane2000}. Since their geometrically unstable nature
706     against non-Hamiltonian perturbations, ordinary implicit Runge-Kutta
707     methods are not suitable for Hamiltonian system. Recently, various
708     high-order explicit Runge-Kutta methods
709     \cite{Owren1992,Chen2003}have been developed to overcome this
710 tim 2703 instability. However, due to computational penalty involved in
711     implementing the Runge-Kutta methods, they do not attract too much
712     attention from Molecular Dynamics community. Instead, splitting have
713     been widely accepted since they exploit natural decompositions of
714 tim 2789 the system\cite{Tuckerman1992, McLachlan1998}.
715 tim 2702
716     \subsubsection{\label{introSection:splittingMethod}Splitting Method}
717    
718     The main idea behind splitting methods is to decompose the discrete
719     $\varphi_h$ as a composition of simpler flows,
720 tim 2699 \begin{equation}
721     \varphi _h = \varphi _{h_1 } \circ \varphi _{h_2 } \ldots \circ
722     \varphi _{h_n }
723     \label{introEquation:FlowDecomposition}
724     \end{equation}
725     where each of the sub-flow is chosen such that each represent a
726 tim 2702 simpler integration of the system.
727    
728     Suppose that a Hamiltonian system takes the form,
729     \[
730     H = H_1 + H_2.
731     \]
732     Here, $H_1$ and $H_2$ may represent different physical processes of
733     the system. For instance, they may relate to kinetic and potential
734     energy respectively, which is a natural decomposition of the
735     problem. If $H_1$ and $H_2$ can be integrated using exact flows
736     $\varphi_1(t)$ and $\varphi_2(t)$, respectively, a simple first
737     order is then given by the Lie-Trotter formula
738 tim 2699 \begin{equation}
739 tim 2702 \varphi _h = \varphi _{1,h} \circ \varphi _{2,h},
740     \label{introEquation:firstOrderSplitting}
741     \end{equation}
742     where $\varphi _h$ is the result of applying the corresponding
743     continuous $\varphi _i$ over a time $h$. By definition, as
744     $\varphi_i(t)$ is the exact solution of a Hamiltonian system, it
745     must follow that each operator $\varphi_i(t)$ is a symplectic map.
746     It is easy to show that any composition of symplectic flows yields a
747     symplectic map,
748     \begin{equation}
749 tim 2699 (\varphi '\phi ')^T J\varphi '\phi ' = \phi '^T \varphi '^T J\varphi
750 tim 2702 '\phi ' = \phi '^T J\phi ' = J,
751 tim 2699 \label{introEquation:SymplecticFlowComposition}
752     \end{equation}
753 tim 2702 where $\phi$ and $\psi$ both are symplectic maps. Thus operator
754     splitting in this context automatically generates a symplectic map.
755 tim 2699
756 tim 2702 The Lie-Trotter splitting(\ref{introEquation:firstOrderSplitting})
757     introduces local errors proportional to $h^2$, while Strang
758     splitting gives a second-order decomposition,
759     \begin{equation}
760     \varphi _h = \varphi _{1,h/2} \circ \varphi _{2,h} \circ \varphi
761 tim 2706 _{1,h/2} , \label{introEquation:secondOrderSplitting}
762 tim 2702 \end{equation}
763     which has a local error proportional to $h^3$. Sprang splitting's
764     popularity in molecular simulation community attribute to its
765     symmetric property,
766     \begin{equation}
767     \varphi _h^{ - 1} = \varphi _{ - h}.
768 tim 2703 \label{introEquation:timeReversible}
769 tim 2702 \end{equation}
770    
771     \subsubsection{\label{introSection:exampleSplittingMethod}Example of Splitting Method}
772     The classical equation for a system consisting of interacting
773     particles can be written in Hamiltonian form,
774     \[
775     H = T + V
776     \]
777     where $T$ is the kinetic energy and $V$ is the potential energy.
778     Setting $H_1 = T, H_2 = V$ and applying Strang splitting, one
779     obtains the following:
780     \begin{align}
781     q(\Delta t) &= q(0) + \dot{q}(0)\Delta t +
782     \frac{F[q(0)]}{m}\frac{\Delta t^2}{2}, %
783     \label{introEquation:Lp10a} \\%
784     %
785     \dot{q}(\Delta t) &= \dot{q}(0) + \frac{\Delta t}{2m}
786     \biggl [F[q(0)] + F[q(\Delta t)] \biggr]. %
787     \label{introEquation:Lp10b}
788     \end{align}
789     where $F(t)$ is the force at time $t$. This integration scheme is
790     known as \emph{velocity verlet} which is
791     symplectic(\ref{introEquation:SymplecticFlowComposition}),
792     time-reversible(\ref{introEquation:timeReversible}) and
793     volume-preserving (\ref{introEquation:volumePreserving}). These
794     geometric properties attribute to its long-time stability and its
795     popularity in the community. However, the most commonly used
796     velocity verlet integration scheme is written as below,
797     \begin{align}
798     \dot{q}\biggl (\frac{\Delta t}{2}\biggr ) &=
799     \dot{q}(0) + \frac{\Delta t}{2m}\, F[q(0)], \label{introEquation:Lp9a}\\%
800     %
801     q(\Delta t) &= q(0) + \Delta t\, \dot{q}\biggl (\frac{\Delta t}{2}\biggr ),%
802     \label{introEquation:Lp9b}\\%
803     %
804     \dot{q}(\Delta t) &= \dot{q}\biggl (\frac{\Delta t}{2}\biggr ) +
805     \frac{\Delta t}{2m}\, F[q(0)]. \label{introEquation:Lp9c}
806     \end{align}
807     From the preceding splitting, one can see that the integration of
808     the equations of motion would follow:
809     \begin{enumerate}
810     \item calculate the velocities at the half step, $\frac{\Delta t}{2}$, from the forces calculated at the initial position.
811    
812     \item Use the half step velocities to move positions one whole step, $\Delta t$.
813    
814     \item Evaluate the forces at the new positions, $\mathbf{r}(\Delta t)$, and use the new forces to complete the velocity move.
815    
816     \item Repeat from step 1 with the new position, velocities, and forces assuming the roles of the initial values.
817     \end{enumerate}
818    
819     Simply switching the order of splitting and composing, a new
820     integrator, the \emph{position verlet} integrator, can be generated,
821     \begin{align}
822     \dot q(\Delta t) &= \dot q(0) + \Delta tF(q(0))\left[ {q(0) +
823     \frac{{\Delta t}}{{2m}}\dot q(0)} \right], %
824     \label{introEquation:positionVerlet1} \\%
825     %
826 tim 2703 q(\Delta t) &= q(0) + \frac{{\Delta t}}{2}\left[ {\dot q(0) + \dot
827 tim 2702 q(\Delta t)} \right]. %
828 tim 2719 \label{introEquation:positionVerlet2}
829 tim 2702 \end{align}
830    
831     \subsubsection{\label{introSection:errorAnalysis}Error Analysis and Higher Order Methods}
832    
833     Baker-Campbell-Hausdorff formula can be used to determine the local
834     error of splitting method in terms of commutator of the
835     operators(\ref{introEquation:exponentialOperator}) associated with
836     the sub-flow. For operators $hX$ and $hY$ which are associate to
837 tim 2726 $\varphi_1(t)$ and $\varphi_2(t)$ respectively , we have
838 tim 2702 \begin{equation}
839     \exp (hX + hY) = \exp (hZ)
840     \end{equation}
841     where
842     \begin{equation}
843     hZ = hX + hY + \frac{{h^2 }}{2}[X,Y] + \frac{{h^3 }}{2}\left(
844     {[X,[X,Y]] + [Y,[Y,X]]} \right) + \ldots .
845     \end{equation}
846     Here, $[X,Y]$ is the commutators of operator $X$ and $Y$ given by
847     \[
848     [X,Y] = XY - YX .
849     \]
850 tim 2789 Applying Baker-Campbell-Hausdorff formula\cite{Varadarajan1974} to
851     Sprang splitting, we can obtain
852 tim 2779 \begin{eqnarray*}
853 tim 2778 \exp (h X/2)\exp (h Y)\exp (h X/2) & = & \exp (h X + h Y + h^2 [X,Y]/4 + h^2 [Y,X]/4 \\
854     & & \mbox{} + h^2 [X,X]/8 + h^2 [Y,Y]/8 \\
855 tim 2779 & & \mbox{} + h^3 [Y,[Y,X]]/12 - h^3[X,[X,Y]]/24 + \ldots )
856     \end{eqnarray*}
857 tim 2702 Since \[ [X,Y] + [Y,X] = 0\] and \[ [X,X] = 0\], the dominant local
858     error of Spring splitting is proportional to $h^3$. The same
859     procedure can be applied to general splitting, of the form
860     \begin{equation}
861     \varphi _{b_m h}^2 \circ \varphi _{a_m h}^1 \circ \varphi _{b_{m -
862     1} h}^2 \circ \ldots \circ \varphi _{a_1 h}^1 .
863     \end{equation}
864 tim 2779 Careful choice of coefficient $a_1 \ldot b_m$ will lead to higher
865 tim 2702 order method. Yoshida proposed an elegant way to compose higher
866 tim 2789 order methods based on symmetric splitting\cite{Yoshida1990}. Given
867     a symmetric second order base method $ \varphi _h^{(2)} $, a
868     fourth-order symmetric method can be constructed by composing,
869 tim 2702 \[
870     \varphi _h^{(4)} = \varphi _{\alpha h}^{(2)} \circ \varphi _{\beta
871     h}^{(2)} \circ \varphi _{\alpha h}^{(2)}
872     \]
873     where $ \alpha = - \frac{{2^{1/3} }}{{2 - 2^{1/3} }}$ and $ \beta
874     = \frac{{2^{1/3} }}{{2 - 2^{1/3} }}$. Moreover, a symmetric
875     integrator $ \varphi _h^{(2n + 2)}$ can be composed by
876     \begin{equation}
877     \varphi _h^{(2n + 2)} = \varphi _{\alpha h}^{(2n)} \circ \varphi
878     _{\beta h}^{(2n)} \circ \varphi _{\alpha h}^{(2n)}
879     \end{equation}
880     , if the weights are chosen as
881     \[
882     \alpha = - \frac{{2^{1/(2n + 1)} }}{{2 - 2^{1/(2n + 1)} }},\beta =
883     \frac{{2^{1/(2n + 1)} }}{{2 - 2^{1/(2n + 1)} }} .
884     \]
885    
886 tim 2694 \section{\label{introSection:molecularDynamics}Molecular Dynamics}
887    
888 tim 2720 As one of the principal tools of molecular modeling, Molecular
889     dynamics has proven to be a powerful tool for studying the functions
890     of biological systems, providing structural, thermodynamic and
891     dynamical information. The basic idea of molecular dynamics is that
892     macroscopic properties are related to microscopic behavior and
893     microscopic behavior can be calculated from the trajectories in
894     simulations. For instance, instantaneous temperature of an
895     Hamiltonian system of $N$ particle can be measured by
896     \[
897 tim 2725 T = \sum\limits_{i = 1}^N {\frac{{m_i v_i^2 }}{{fk_B }}}
898 tim 2720 \]
899     where $m_i$ and $v_i$ are the mass and velocity of $i$th particle
900     respectively, $f$ is the number of degrees of freedom, and $k_B$ is
901     the boltzman constant.
902 tim 2694
903 tim 2720 A typical molecular dynamics run consists of three essential steps:
904     \begin{enumerate}
905     \item Initialization
906     \begin{enumerate}
907     \item Preliminary preparation
908     \item Minimization
909     \item Heating
910     \item Equilibration
911     \end{enumerate}
912     \item Production
913     \item Analysis
914     \end{enumerate}
915     These three individual steps will be covered in the following
916     sections. Sec.~\ref{introSec:initialSystemSettings} deals with the
917     initialization of a simulation. Sec.~\ref{introSec:production} will
918 tim 2725 discusses issues in production run. Sec.~\ref{introSection:Analysis}
919     provides the theoretical tools for trajectory analysis.
920 tim 2719
921 tim 2720 \subsection{\label{introSec:initialSystemSettings}Initialization}
922 tim 2719
923 tim 2720 \subsubsection{Preliminary preparation}
924 tim 2719
925 tim 2720 When selecting the starting structure of a molecule for molecular
926     simulation, one may retrieve its Cartesian coordinates from public
927     databases, such as RCSB Protein Data Bank \textit{etc}. Although
928     thousands of crystal structures of molecules are discovered every
929     year, many more remain unknown due to the difficulties of
930     purification and crystallization. Even for the molecule with known
931     structure, some important information is missing. For example, the
932     missing hydrogen atom which acts as donor in hydrogen bonding must
933     be added. Moreover, in order to include electrostatic interaction,
934     one may need to specify the partial charges for individual atoms.
935     Under some circumstances, we may even need to prepare the system in
936     a special setup. For instance, when studying transport phenomenon in
937     membrane system, we may prepare the lipids in bilayer structure
938     instead of placing lipids randomly in solvent, since we are not
939     interested in self-aggregation and it takes a long time to happen.
940 tim 2694
941 tim 2720 \subsubsection{Minimization}
942 tim 2705
943 tim 2720 It is quite possible that some of molecules in the system from
944     preliminary preparation may be overlapped with each other. This
945     close proximity leads to high potential energy which consequently
946     jeopardizes any molecular dynamics simulations. To remove these
947     steric overlaps, one typically performs energy minimization to find
948     a more reasonable conformation. Several energy minimization methods
949     have been developed to exploit the energy surface and to locate the
950     local minimum. While converging slowly near the minimum, steepest
951     descent method is extremely robust when systems are far from
952     harmonic. Thus, it is often used to refine structure from
953     crystallographic data. Relied on the gradient or hessian, advanced
954     methods like conjugate gradient and Newton-Raphson converge rapidly
955     to a local minimum, while become unstable if the energy surface is
956     far from quadratic. Another factor must be taken into account, when
957     choosing energy minimization method, is the size of the system.
958     Steepest descent and conjugate gradient can deal with models of any
959     size. Because of the limit of computation power to calculate hessian
960     matrix and insufficient storage capacity to store them, most
961     Newton-Raphson methods can not be used with very large models.
962 tim 2694
963 tim 2720 \subsubsection{Heating}
964    
965     Typically, Heating is performed by assigning random velocities
966     according to a Gaussian distribution for a temperature. Beginning at
967     a lower temperature and gradually increasing the temperature by
968     assigning greater random velocities, we end up with setting the
969     temperature of the system to a final temperature at which the
970     simulation will be conducted. In heating phase, we should also keep
971     the system from drifting or rotating as a whole. Equivalently, the
972     net linear momentum and angular momentum of the system should be
973     shifted to zero.
974    
975     \subsubsection{Equilibration}
976    
977     The purpose of equilibration is to allow the system to evolve
978     spontaneously for a period of time and reach equilibrium. The
979     procedure is continued until various statistical properties, such as
980     temperature, pressure, energy, volume and other structural
981     properties \textit{etc}, become independent of time. Strictly
982     speaking, minimization and heating are not necessary, provided the
983     equilibration process is long enough. However, these steps can serve
984     as a means to arrive at an equilibrated structure in an effective
985     way.
986    
987     \subsection{\label{introSection:production}Production}
988    
989 tim 2789 Production run is the most important step of the simulation, in
990 tim 2725 which the equilibrated structure is used as a starting point and the
991     motions of the molecules are collected for later analysis. In order
992     to capture the macroscopic properties of the system, the molecular
993     dynamics simulation must be performed in correct and efficient way.
994 tim 2720
995 tim 2725 The most expensive part of a molecular dynamics simulation is the
996     calculation of non-bonded forces, such as van der Waals force and
997     Coulombic forces \textit{etc}. For a system of $N$ particles, the
998     complexity of the algorithm for pair-wise interactions is $O(N^2 )$,
999     which making large simulations prohibitive in the absence of any
1000     computation saving techniques.
1001 tim 2720
1002 tim 2725 A natural approach to avoid system size issue is to represent the
1003     bulk behavior by a finite number of the particles. However, this
1004     approach will suffer from the surface effect. To offset this,
1005 tim 2789 \textit{Periodic boundary condition} (see Fig.~\ref{introFig:pbc})
1006     is developed to simulate bulk properties with a relatively small
1007     number of particles. In this method, the simulation box is
1008     replicated throughout space to form an infinite lattice. During the
1009     simulation, when a particle moves in the primary cell, its image in
1010     other cells move in exactly the same direction with exactly the same
1011     orientation. Thus, as a particle leaves the primary cell, one of its
1012     images will enter through the opposite face.
1013     \begin{figure}
1014     \centering
1015     \includegraphics[width=\linewidth]{pbc.eps}
1016     \caption[An illustration of periodic boundary conditions]{A 2-D
1017     illustration of periodic boundary conditions. As one particle leaves
1018     the left of the simulation box, an image of it enters the right.}
1019     \label{introFig:pbc}
1020     \end{figure}
1021 tim 2725
1022     %cutoff and minimum image convention
1023     Another important technique to improve the efficiency of force
1024     evaluation is to apply cutoff where particles farther than a
1025     predetermined distance, are not included in the calculation
1026     \cite{Frenkel1996}. The use of a cutoff radius will cause a
1027 tim 2730 discontinuity in the potential energy curve. Fortunately, one can
1028     shift the potential to ensure the potential curve go smoothly to
1029     zero at the cutoff radius. Cutoff strategy works pretty well for
1030     Lennard-Jones interaction because of its short range nature.
1031     However, simply truncating the electrostatic interaction with the
1032     use of cutoff has been shown to lead to severe artifacts in
1033     simulations. Ewald summation, in which the slowly conditionally
1034     convergent Coulomb potential is transformed into direct and
1035     reciprocal sums with rapid and absolute convergence, has proved to
1036     minimize the periodicity artifacts in liquid simulations. Taking the
1037     advantages of the fast Fourier transform (FFT) for calculating
1038 tim 2789 discrete Fourier transforms, the particle mesh-based
1039     methods\cite{Hockney1981,Shimada1993, Luty1994} are accelerated from
1040     $O(N^{3/2})$ to $O(N logN)$. An alternative approach is \emph{fast
1041     multipole method}\cite{Greengard1987, Greengard1994}, which treats
1042     Coulombic interaction exactly at short range, and approximate the
1043     potential at long range through multipolar expansion. In spite of
1044     their wide acceptances at the molecular simulation community, these
1045     two methods are hard to be implemented correctly and efficiently.
1046     Instead, we use a damped and charge-neutralized Coulomb potential
1047     method developed by Wolf and his coworkers\cite{Wolf1999}. The
1048     shifted Coulomb potential for particle $i$ and particle $j$ at
1049     distance $r_{rj}$ is given by:
1050 tim 2725 \begin{equation}
1051     V(r_{ij})= \frac{q_i q_j \textrm{erfc}(\alpha
1052     r_{ij})}{r_{ij}}-\lim_{r_{ij}\rightarrow
1053     R_\textrm{c}}\left\{\frac{q_iq_j \textrm{erfc}(\alpha
1054     r_{ij})}{r_{ij}}\right\}. \label{introEquation:shiftedCoulomb}
1055     \end{equation}
1056     where $\alpha$ is the convergence parameter. Due to the lack of
1057     inherent periodicity and rapid convergence,this method is extremely
1058     efficient and easy to implement.
1059 tim 2789 \begin{figure}
1060     \centering
1061     \includegraphics[width=\linewidth]{shifted_coulomb.eps}
1062     \caption[An illustration of shifted Coulomb potential]{An
1063     illustration of shifted Coulomb potential.}
1064     \label{introFigure:shiftedCoulomb}
1065     \end{figure}
1066 tim 2725
1067     %multiple time step
1068    
1069 tim 2720 \subsection{\label{introSection:Analysis} Analysis}
1070    
1071 tim 2721 Recently, advanced visualization technique are widely applied to
1072     monitor the motions of molecules. Although the dynamics of the
1073     system can be described qualitatively from animation, quantitative
1074     trajectory analysis are more appreciable. According to the
1075     principles of Statistical Mechanics,
1076     Sec.~\ref{introSection:statisticalMechanics}, one can compute
1077     thermodynamics properties, analyze fluctuations of structural
1078     parameters, and investigate time-dependent processes of the molecule
1079     from the trajectories.
1080    
1081     \subsubsection{\label{introSection:thermodynamicsProperties}Thermodynamics Properties}
1082    
1083 tim 2725 Thermodynamics properties, which can be expressed in terms of some
1084     function of the coordinates and momenta of all particles in the
1085     system, can be directly computed from molecular dynamics. The usual
1086     way to measure the pressure is based on virial theorem of Clausius
1087     which states that the virial is equal to $-3Nk_BT$. For a system
1088     with forces between particles, the total virial, $W$, contains the
1089     contribution from external pressure and interaction between the
1090     particles:
1091     \[
1092     W = - 3PV + \left\langle {\sum\limits_{i < j} {r{}_{ij} \cdot
1093     f_{ij} } } \right\rangle
1094     \]
1095     where $f_{ij}$ is the force between particle $i$ and $j$ at a
1096     distance $r_{ij}$. Thus, the expression for the pressure is given
1097     by:
1098     \begin{equation}
1099     P = \frac{{Nk_B T}}{V} - \frac{1}{{3V}}\left\langle {\sum\limits_{i
1100     < j} {r{}_{ij} \cdot f_{ij} } } \right\rangle
1101     \end{equation}
1102    
1103 tim 2721 \subsubsection{\label{introSection:structuralProperties}Structural Properties}
1104    
1105     Structural Properties of a simple fluid can be described by a set of
1106     distribution functions. Among these functions,\emph{pair
1107     distribution function}, also known as \emph{radial distribution
1108 tim 2725 function}, is of most fundamental importance to liquid-state theory.
1109     Pair distribution function can be gathered by Fourier transforming
1110     raw data from a series of neutron diffraction experiments and
1111 tim 2786 integrating over the surface factor \cite{Powles1973}. The
1112     experiment result can serve as a criterion to justify the
1113     correctness of the theory. Moreover, various equilibrium
1114     thermodynamic and structural properties can also be expressed in
1115     terms of radial distribution function \cite{Allen1987}.
1116 tim 2721
1117     A pair distribution functions $g(r)$ gives the probability that a
1118     particle $i$ will be located at a distance $r$ from a another
1119     particle $j$ in the system
1120     \[
1121     g(r) = \frac{V}{{N^2 }}\left\langle {\sum\limits_i {\sum\limits_{j
1122     \ne i} {\delta (r - r_{ij} )} } } \right\rangle.
1123     \]
1124     Note that the delta function can be replaced by a histogram in
1125     computer simulation. Figure
1126     \ref{introFigure:pairDistributionFunction} shows a typical pair
1127     distribution function for the liquid argon system. The occurrence of
1128     several peaks in the plot of $g(r)$ suggests that it is more likely
1129     to find particles at certain radial values than at others. This is a
1130     result of the attractive interaction at such distances. Because of
1131     the strong repulsive forces at short distance, the probability of
1132     locating particles at distances less than about 2.5{\AA} from each
1133     other is essentially zero.
1134    
1135     %\begin{figure}
1136     %\centering
1137     %\includegraphics[width=\linewidth]{pdf.eps}
1138     %\caption[Pair distribution function for the liquid argon
1139     %]{Pair distribution function for the liquid argon}
1140     %\label{introFigure:pairDistributionFunction}
1141     %\end{figure}
1142    
1143     \subsubsection{\label{introSection:timeDependentProperties}Time-dependent
1144     Properties}
1145    
1146     Time-dependent properties are usually calculated using \emph{time
1147     correlation function}, which correlates random variables $A$ and $B$
1148     at two different time
1149     \begin{equation}
1150     C_{AB} (t) = \left\langle {A(t)B(0)} \right\rangle.
1151     \label{introEquation:timeCorrelationFunction}
1152     \end{equation}
1153     If $A$ and $B$ refer to same variable, this kind of correlation
1154     function is called \emph{auto correlation function}. One example of
1155     auto correlation function is velocity auto-correlation function
1156     which is directly related to transport properties of molecular
1157 tim 2725 liquids:
1158     \[
1159     D = \frac{1}{3}\int\limits_0^\infty {\left\langle {v(t) \cdot v(0)}
1160     \right\rangle } dt
1161     \]
1162     where $D$ is diffusion constant. Unlike velocity autocorrelation
1163     function which is averaging over time origins and over all the
1164     atoms, dipole autocorrelation are calculated for the entire system.
1165     The dipole autocorrelation function is given by:
1166     \[
1167     c_{dipole} = \left\langle {u_{tot} (t) \cdot u_{tot} (t)}
1168     \right\rangle
1169     \]
1170     Here $u_{tot}$ is the net dipole of the entire system and is given
1171     by
1172     \[
1173     u_{tot} (t) = \sum\limits_i {u_i (t)}
1174     \]
1175     In principle, many time correlation functions can be related with
1176     Fourier transforms of the infrared, Raman, and inelastic neutron
1177     scattering spectra of molecular liquids. In practice, one can
1178     extract the IR spectrum from the intensity of dipole fluctuation at
1179     each frequency using the following relationship:
1180     \[
1181     \hat c_{dipole} (v) = \int_{ - \infty }^\infty {c_{dipole} (t)e^{ -
1182     i2\pi vt} dt}
1183     \]
1184 tim 2721
1185 tim 2693 \section{\label{introSection:rigidBody}Dynamics of Rigid Bodies}
1186 tim 2692
1187 tim 2705 Rigid bodies are frequently involved in the modeling of different
1188     areas, from engineering, physics, to chemistry. For example,
1189     missiles and vehicle are usually modeled by rigid bodies. The
1190     movement of the objects in 3D gaming engine or other physics
1191     simulator is governed by the rigid body dynamics. In molecular
1192     simulation, rigid body is used to simplify the model in
1193 tim 2789 protein-protein docking study\cite{Gray2003}.
1194 tim 2694
1195 tim 2705 It is very important to develop stable and efficient methods to
1196     integrate the equations of motion of orientational degrees of
1197     freedom. Euler angles are the nature choice to describe the
1198     rotational degrees of freedom. However, due to its singularity, the
1199     numerical integration of corresponding equations of motion is very
1200     inefficient and inaccurate. Although an alternative integrator using
1201 tim 2789 different sets of Euler angles can overcome this
1202     difficulty\cite{Barojas1973}, the computational penalty and the lost
1203     of angular momentum conservation still remain. A singularity free
1204     representation utilizing quaternions was developed by Evans in
1205     1977\cite{Evans1977}. Unfortunately, this approach suffer from the
1206     nonseparable Hamiltonian resulted from quaternion representation,
1207     which prevents the symplectic algorithm to be utilized. Another
1208     different approach is to apply holonomic constraints to the atoms
1209     belonging to the rigid body. Each atom moves independently under the
1210     normal forces deriving from potential energy and constraint forces
1211     which are used to guarantee the rigidness. However, due to their
1212     iterative nature, SHAKE and Rattle algorithm converge very slowly
1213     when the number of constraint increases\cite{Ryckaert1977,
1214     Andersen1983}.
1215 tim 2694
1216 tim 2705 The break through in geometric literature suggests that, in order to
1217     develop a long-term integration scheme, one should preserve the
1218     symplectic structure of the flow. Introducing conjugate momentum to
1219 tim 2719 rotation matrix $Q$ and re-formulating Hamiltonian's equation, a
1220 tim 2789 symplectic integrator, RSHAKE\cite{Kol1997}, was proposed to evolve
1221     the Hamiltonian system in a constraint manifold by iteratively
1222 tim 2719 satisfying the orthogonality constraint $Q_T Q = 1$. An alternative
1223 tim 2789 method using quaternion representation was developed by
1224     Omelyan\cite{Omelyan1998}. However, both of these methods are
1225     iterative and inefficient. In this section, we will present a
1226     symplectic Lie-Poisson integrator for rigid body developed by
1227     Dullweber and his coworkers\cite{Dullweber1997} in depth.
1228 tim 2705
1229 tim 2706 \subsection{\label{introSection:constrainedHamiltonianRB}Constrained Hamiltonian for Rigid Body}
1230 tim 2713 The motion of the rigid body is Hamiltonian with the Hamiltonian
1231     function
1232 tim 2706 \begin{equation}
1233     H = \frac{1}{2}(p^T m^{ - 1} p) + \frac{1}{2}tr(PJ^{ - 1} P) +
1234     V(q,Q) + \frac{1}{2}tr[(QQ^T - 1)\Lambda ].
1235     \label{introEquation:RBHamiltonian}
1236     \end{equation}
1237     Here, $q$ and $Q$ are the position and rotation matrix for the
1238     rigid-body, $p$ and $P$ are conjugate momenta to $q$ and $Q$ , and
1239     $J$, a diagonal matrix, is defined by
1240     \[
1241     I_{ii}^{ - 1} = \frac{1}{2}\sum\limits_{i \ne j} {J_{jj}^{ - 1} }
1242     \]
1243     where $I_{ii}$ is the diagonal element of the inertia tensor. This
1244     constrained Hamiltonian equation subjects to a holonomic constraint,
1245     \begin{equation}
1246 tim 2726 Q^T Q = 1, \label{introEquation:orthogonalConstraint}
1247 tim 2706 \end{equation}
1248     which is used to ensure rotation matrix's orthogonality.
1249     Differentiating \ref{introEquation:orthogonalConstraint} and using
1250     Equation \ref{introEquation:RBMotionMomentum}, one may obtain,
1251     \begin{equation}
1252 tim 2707 Q^T PJ^{ - 1} + J^{ - 1} P^T Q = 0 . \\
1253 tim 2706 \label{introEquation:RBFirstOrderConstraint}
1254     \end{equation}
1255    
1256     Using Equation (\ref{introEquation:motionHamiltonianCoordinate},
1257     \ref{introEquation:motionHamiltonianMomentum}), one can write down
1258     the equations of motion,
1259     \[
1260     \begin{array}{c}
1261     \frac{{dq}}{{dt}} = \frac{p}{m} \label{introEquation:RBMotionPosition}\\
1262     \frac{{dp}}{{dt}} = - \nabla _q V(q,Q) \label{introEquation:RBMotionMomentum}\\
1263     \frac{{dQ}}{{dt}} = PJ^{ - 1} \label{introEquation:RBMotionRotation}\\
1264 tim 2707 \frac{{dP}}{{dt}} = - \nabla _Q V(q,Q) - 2Q\Lambda . \label{introEquation:RBMotionP}\\
1265 tim 2706 \end{array}
1266     \]
1267    
1268 tim 2707 In general, there are two ways to satisfy the holonomic constraints.
1269     We can use constraint force provided by lagrange multiplier on the
1270     normal manifold to keep the motion on constraint space. Or we can
1271 tim 2776 simply evolve the system in constraint manifold. These two methods
1272     are proved to be equivalent. The holonomic constraint and equations
1273     of motions define a constraint manifold for rigid body
1274 tim 2707 \[
1275     M = \left\{ {(Q,P):Q^T Q = 1,Q^T PJ^{ - 1} + J^{ - 1} P^T Q = 0}
1276     \right\}.
1277     \]
1278 tim 2706
1279 tim 2707 Unfortunately, this constraint manifold is not the cotangent bundle
1280     $T_{\star}SO(3)$. However, it turns out that under symplectic
1281     transformation, the cotangent space and the phase space are
1282     diffeomorphic. Introducing
1283 tim 2706 \[
1284 tim 2707 \tilde Q = Q,\tilde P = \frac{1}{2}\left( {P - QP^T Q} \right),
1285 tim 2706 \]
1286 tim 2707 the mechanical system subject to a holonomic constraint manifold $M$
1287     can be re-formulated as a Hamiltonian system on the cotangent space
1288     \[
1289     T^* SO(3) = \left\{ {(\tilde Q,\tilde P):\tilde Q^T \tilde Q =
1290     1,\tilde Q^T \tilde PJ^{ - 1} + J^{ - 1} P^T \tilde Q = 0} \right\}
1291     \]
1292 tim 2706
1293 tim 2707 For a body fixed vector $X_i$ with respect to the center of mass of
1294     the rigid body, its corresponding lab fixed vector $X_0^{lab}$ is
1295     given as
1296     \begin{equation}
1297     X_i^{lab} = Q X_i + q.
1298     \end{equation}
1299     Therefore, potential energy $V(q,Q)$ is defined by
1300     \[
1301     V(q,Q) = V(Q X_0 + q).
1302     \]
1303 tim 2713 Hence, the force and torque are given by
1304 tim 2707 \[
1305 tim 2713 \nabla _q V(q,Q) = F(q,Q) = \sum\limits_i {F_i (q,Q)},
1306 tim 2707 \]
1307 tim 2713 and
1308 tim 2707 \[
1309     \nabla _Q V(q,Q) = F(q,Q)X_i^t
1310     \]
1311 tim 2713 respectively.
1312 tim 2695
1313 tim 2707 As a common choice to describe the rotation dynamics of the rigid
1314     body, angular momentum on body frame $\Pi = Q^t P$ is introduced to
1315     rewrite the equations of motion,
1316     \begin{equation}
1317     \begin{array}{l}
1318     \mathop \Pi \limits^ \bullet = J^{ - 1} \Pi ^T \Pi + Q^T \sum\limits_i {F_i (q,Q)X_i^T } - \Lambda \\
1319     \mathop Q\limits^{{\rm{ }} \bullet } = Q\Pi {\rm{ }}J^{ - 1} \\
1320     \end{array}
1321     \label{introEqaution:RBMotionPI}
1322     \end{equation}
1323     , as well as holonomic constraints,
1324     \[
1325     \begin{array}{l}
1326     \Pi J^{ - 1} + J^{ - 1} \Pi ^t = 0 \\
1327     Q^T Q = 1 \\
1328     \end{array}
1329     \]
1330 tim 2692
1331 tim 2707 For a vector $v(v_1 ,v_2 ,v_3 ) \in R^3$ and a matrix $\hat v \in
1332     so(3)^ \star$, the hat-map isomorphism,
1333     \begin{equation}
1334     v(v_1 ,v_2 ,v_3 ) \Leftrightarrow \hat v = \left(
1335     {\begin{array}{*{20}c}
1336     0 & { - v_3 } & {v_2 } \\
1337     {v_3 } & 0 & { - v_1 } \\
1338     { - v_2 } & {v_1 } & 0 \\
1339     \end{array}} \right),
1340     \label{introEquation:hatmapIsomorphism}
1341     \end{equation}
1342     will let us associate the matrix products with traditional vector
1343     operations
1344     \[
1345     \hat vu = v \times u
1346     \]
1347    
1348     Using \ref{introEqaution:RBMotionPI}, one can construct a skew
1349     matrix,
1350     \begin{equation}
1351     (\mathop \Pi \limits^ \bullet - \mathop \Pi \limits^ \bullet ^T
1352     ){\rm{ }} = {\rm{ }}(\Pi - \Pi ^T ){\rm{ }}(J^{ - 1} \Pi + \Pi J^{
1353     - 1} ) + \sum\limits_i {[Q^T F_i (r,Q)X_i^T - X_i F_i (r,Q)^T Q]} -
1354     (\Lambda - \Lambda ^T ) . \label{introEquation:skewMatrixPI}
1355     \end{equation}
1356     Since $\Lambda$ is symmetric, the last term of Equation
1357 tim 2713 \ref{introEquation:skewMatrixPI} is zero, which implies the Lagrange
1358     multiplier $\Lambda$ is absent from the equations of motion. This
1359     unique property eliminate the requirement of iterations which can
1360 tim 2789 not be avoided in other methods\cite{Kol1997, Omelyan1998}.
1361 tim 2707
1362 tim 2713 Applying hat-map isomorphism, we obtain the equation of motion for
1363     angular momentum on body frame
1364     \begin{equation}
1365     \dot \pi = \pi \times I^{ - 1} \pi + \sum\limits_i {\left( {Q^T
1366     F_i (r,Q)} \right) \times X_i }.
1367     \label{introEquation:bodyAngularMotion}
1368     \end{equation}
1369 tim 2707 In the same manner, the equation of motion for rotation matrix is
1370     given by
1371     \[
1372 tim 2713 \dot Q = Qskew(I^{ - 1} \pi )
1373 tim 2707 \]
1374    
1375 tim 2713 \subsection{\label{introSection:SymplecticFreeRB}Symplectic
1376     Lie-Poisson Integrator for Free Rigid Body}
1377 tim 2707
1378 tim 2713 If there is not external forces exerted on the rigid body, the only
1379     contribution to the rotational is from the kinetic potential (the
1380     first term of \ref{ introEquation:bodyAngularMotion}). The free
1381     rigid body is an example of Lie-Poisson system with Hamiltonian
1382     function
1383     \begin{equation}
1384     T^r (\pi ) = T_1 ^r (\pi _1 ) + T_2^r (\pi _2 ) + T_3^r (\pi _3 )
1385     \label{introEquation:rotationalKineticRB}
1386     \end{equation}
1387     where $T_i^r (\pi _i ) = \frac{{\pi _i ^2 }}{{2I_i }}$ and
1388     Lie-Poisson structure matrix,
1389     \begin{equation}
1390     J(\pi ) = \left( {\begin{array}{*{20}c}
1391     0 & {\pi _3 } & { - \pi _2 } \\
1392     { - \pi _3 } & 0 & {\pi _1 } \\
1393     {\pi _2 } & { - \pi _1 } & 0 \\
1394     \end{array}} \right)
1395     \end{equation}
1396     Thus, the dynamics of free rigid body is governed by
1397     \begin{equation}
1398     \frac{d}{{dt}}\pi = J(\pi )\nabla _\pi T^r (\pi )
1399     \end{equation}
1400 tim 2707
1401 tim 2713 One may notice that each $T_i^r$ in Equation
1402     \ref{introEquation:rotationalKineticRB} can be solved exactly. For
1403     instance, the equations of motion due to $T_1^r$ are given by
1404     \begin{equation}
1405     \frac{d}{{dt}}\pi = R_1 \pi ,\frac{d}{{dt}}Q = QR_1
1406     \label{introEqaution:RBMotionSingleTerm}
1407     \end{equation}
1408     where
1409     \[ R_1 = \left( {\begin{array}{*{20}c}
1410     0 & 0 & 0 \\
1411     0 & 0 & {\pi _1 } \\
1412     0 & { - \pi _1 } & 0 \\
1413     \end{array}} \right).
1414     \]
1415     The solutions of Equation \ref{introEqaution:RBMotionSingleTerm} is
1416 tim 2707 \[
1417 tim 2713 \pi (\Delta t) = e^{\Delta tR_1 } \pi (0),Q(\Delta t) =
1418     Q(0)e^{\Delta tR_1 }
1419 tim 2707 \]
1420 tim 2713 with
1421 tim 2707 \[
1422 tim 2713 e^{\Delta tR_1 } = \left( {\begin{array}{*{20}c}
1423     0 & 0 & 0 \\
1424     0 & {\cos \theta _1 } & {\sin \theta _1 } \\
1425     0 & { - \sin \theta _1 } & {\cos \theta _1 } \\
1426     \end{array}} \right),\theta _1 = \frac{{\pi _1 }}{{I_1 }}\Delta t.
1427 tim 2707 \]
1428 tim 2719 To reduce the cost of computing expensive functions in $e^{\Delta
1429     tR_1 }$, we can use Cayley transformation,
1430 tim 2713 \[
1431     e^{\Delta tR_1 } \approx (1 - \Delta tR_1 )^{ - 1} (1 + \Delta tR_1
1432     )
1433     \]
1434 tim 2720 The flow maps for $T_2^r$ and $T_3^r$ can be found in the same
1435 tim 2713 manner.
1436    
1437     In order to construct a second-order symplectic method, we split the
1438     angular kinetic Hamiltonian function can into five terms
1439 tim 2707 \[
1440 tim 2713 T^r (\pi ) = \frac{1}{2}T_1 ^r (\pi _1 ) + \frac{1}{2}T_2^r (\pi _2
1441     ) + T_3^r (\pi _3 ) + \frac{1}{2}T_2^r (\pi _2 ) + \frac{1}{2}T_1 ^r
1442     (\pi _1 )
1443     \].
1444     Concatenating flows corresponding to these five terms, we can obtain
1445     an symplectic integrator,
1446     \[
1447     \varphi _{\Delta t,T^r } = \varphi _{\Delta t/2,\pi _1 } \circ
1448 tim 2707 \varphi _{\Delta t/2,\pi _2 } \circ \varphi _{\Delta t,\pi _3 }
1449     \circ \varphi _{\Delta t/2,\pi _2 } \circ \varphi _{\Delta t/2,\pi
1450 tim 2713 _1 }.
1451 tim 2707 \]
1452    
1453 tim 2713 The non-canonical Lie-Poisson bracket ${F, G}$ of two function
1454     $F(\pi )$ and $G(\pi )$ is defined by
1455 tim 2707 \[
1456 tim 2713 \{ F,G\} (\pi ) = [\nabla _\pi F(\pi )]^T J(\pi )\nabla _\pi G(\pi
1457     )
1458     \]
1459     If the Poisson bracket of a function $F$ with an arbitrary smooth
1460     function $G$ is zero, $F$ is a \emph{Casimir}, which is the
1461     conserved quantity in Poisson system. We can easily verify that the
1462     norm of the angular momentum, $\parallel \pi
1463     \parallel$, is a \emph{Casimir}. Let$ F(\pi ) = S(\frac{{\parallel
1464     \pi \parallel ^2 }}{2})$ for an arbitrary function $ S:R \to R$ ,
1465     then by the chain rule
1466     \[
1467     \nabla _\pi F(\pi ) = S'(\frac{{\parallel \pi \parallel ^2
1468     }}{2})\pi
1469     \]
1470     Thus $ [\nabla _\pi F(\pi )]^T J(\pi ) = - S'(\frac{{\parallel \pi
1471     \parallel ^2 }}{2})\pi \times \pi = 0 $. This explicit
1472     Lie-Poisson integrator is found to be extremely efficient and stable
1473     which can be explained by the fact the small angle approximation is
1474     used and the norm of the angular momentum is conserved.
1475    
1476     \subsection{\label{introSection:RBHamiltonianSplitting} Hamiltonian
1477     Splitting for Rigid Body}
1478    
1479     The Hamiltonian of rigid body can be separated in terms of kinetic
1480     energy and potential energy,
1481     \[
1482     H = T(p,\pi ) + V(q,Q)
1483     \]
1484     The equations of motion corresponding to potential energy and
1485     kinetic energy are listed in the below table,
1486 tim 2776 \begin{table}
1487     \caption{Equations of motion due to Potential and Kinetic Energies}
1488 tim 2713 \begin{center}
1489     \begin{tabular}{|l|l|}
1490     \hline
1491     % after \\: \hline or \cline{col1-col2} \cline{col3-col4} ...
1492     Potential & Kinetic \\
1493     $\frac{{dq}}{{dt}} = \frac{p}{m}$ & $\frac{d}{{dt}}q = p$ \\
1494     $\frac{d}{{dt}}p = - \frac{{\partial V}}{{\partial q}}$ & $ \frac{d}{{dt}}p = 0$ \\
1495     $\frac{d}{{dt}}Q = 0$ & $ \frac{d}{{dt}}Q = Qskew(I^{ - 1} j)$ \\
1496     $ \frac{d}{{dt}}\pi = \sum\limits_i {\left( {Q^T F_i (r,Q)} \right) \times X_i }$ & $\frac{d}{{dt}}\pi = \pi \times I^{ - 1} \pi$\\
1497     \hline
1498     \end{tabular}
1499     \end{center}
1500 tim 2776 \end{table}
1501     A second-order symplectic method is now obtained by the
1502     composition of the flow maps,
1503 tim 2713 \[
1504     \varphi _{\Delta t} = \varphi _{\Delta t/2,V} \circ \varphi
1505     _{\Delta t,T} \circ \varphi _{\Delta t/2,V}.
1506     \]
1507 tim 2719 Moreover, $\varphi _{\Delta t/2,V}$ can be divided into two
1508     sub-flows which corresponding to force and torque respectively,
1509 tim 2713 \[
1510 tim 2707 \varphi _{\Delta t/2,V} = \varphi _{\Delta t/2,F} \circ \varphi
1511 tim 2713 _{\Delta t/2,\tau }.
1512 tim 2707 \]
1513 tim 2713 Since the associated operators of $\varphi _{\Delta t/2,F} $ and
1514     $\circ \varphi _{\Delta t/2,\tau }$ are commuted, the composition
1515 tim 2719 order inside $\varphi _{\Delta t/2,V}$ does not matter.
1516 tim 2707
1517 tim 2713 Furthermore, kinetic potential can be separated to translational
1518     kinetic term, $T^t (p)$, and rotational kinetic term, $T^r (\pi )$,
1519     \begin{equation}
1520     T(p,\pi ) =T^t (p) + T^r (\pi ).
1521     \end{equation}
1522     where $ T^t (p) = \frac{1}{2}p^T m^{ - 1} p $ and $T^r (\pi )$ is
1523     defined by \ref{introEquation:rotationalKineticRB}. Therefore, the
1524     corresponding flow maps are given by
1525     \[
1526     \varphi _{\Delta t,T} = \varphi _{\Delta t,T^t } \circ \varphi
1527     _{\Delta t,T^r }.
1528     \]
1529     Finally, we obtain the overall symplectic flow maps for free moving
1530     rigid body
1531     \begin{equation}
1532     \begin{array}{c}
1533     \varphi _{\Delta t} = \varphi _{\Delta t/2,F} \circ \varphi _{\Delta t/2,\tau } \\
1534     \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 } \\
1535     \circ \varphi _{\Delta t/2,\tau } \circ \varphi _{\Delta t/2,F} .\\
1536     \end{array}
1537     \label{introEquation:overallRBFlowMaps}
1538     \end{equation}
1539 tim 2707
1540 tim 2685 \section{\label{introSection:langevinDynamics}Langevin Dynamics}
1541 tim 2716 As an alternative to newtonian dynamics, Langevin dynamics, which
1542     mimics a simple heat bath with stochastic and dissipative forces,
1543     has been applied in a variety of studies. This section will review
1544     the theory of Langevin dynamics simulation. A brief derivation of
1545 tim 2719 generalized Langevin equation will be given first. Follow that, we
1546 tim 2716 will discuss the physical meaning of the terms appearing in the
1547     equation as well as the calculation of friction tensor from
1548     hydrodynamics theory.
1549 tim 2685
1550 tim 2719 \subsection{\label{introSection:generalizedLangevinDynamics}Derivation of Generalized Langevin Equation}
1551 tim 2685
1552 tim 2719 Harmonic bath model, in which an effective set of harmonic
1553     oscillators are used to mimic the effect of a linearly responding
1554     environment, has been widely used in quantum chemistry and
1555     statistical mechanics. One of the successful applications of
1556     Harmonic bath model is the derivation of Deriving Generalized
1557     Langevin Dynamics. Lets consider a system, in which the degree of
1558     freedom $x$ is assumed to couple to the bath linearly, giving a
1559     Hamiltonian of the form
1560 tim 2696 \begin{equation}
1561     H = \frac{{p^2 }}{{2m}} + U(x) + H_B + \Delta U(x,x_1 , \ldots x_N)
1562 tim 2719 \label{introEquation:bathGLE}.
1563 tim 2696 \end{equation}
1564 tim 2719 Here $p$ is a momentum conjugate to $q$, $m$ is the mass associated
1565     with this degree of freedom, $H_B$ is harmonic bath Hamiltonian,
1566 tim 2696 \[
1567 tim 2719 H_B = \sum\limits_{\alpha = 1}^N {\left\{ {\frac{{p_\alpha ^2
1568     }}{{2m_\alpha }} + \frac{1}{2}m_\alpha \omega _\alpha ^2 }
1569     \right\}}
1570 tim 2696 \]
1571 tim 2719 where the index $\alpha$ runs over all the bath degrees of freedom,
1572     $\omega _\alpha$ are the harmonic bath frequencies, $m_\alpha$ are
1573     the harmonic bath masses, and $\Delta U$ is bilinear system-bath
1574     coupling,
1575 tim 2696 \[
1576     \Delta U = - \sum\limits_{\alpha = 1}^N {g_\alpha x_\alpha x}
1577     \]
1578 tim 2719 where $g_\alpha$ are the coupling constants between the bath and the
1579     coordinate $x$. Introducing
1580 tim 2696 \[
1581 tim 2719 W(x) = U(x) - \sum\limits_{\alpha = 1}^N {\frac{{g_\alpha ^2
1582     }}{{2m_\alpha w_\alpha ^2 }}} x^2
1583     \] and combining the last two terms in Equation
1584     \ref{introEquation:bathGLE}, we may rewrite the Harmonic bath
1585     Hamiltonian as
1586 tim 2696 \[
1587     H = \frac{{p^2 }}{{2m}} + W(x) + \sum\limits_{\alpha = 1}^N
1588     {\left\{ {\frac{{p_\alpha ^2 }}{{2m_\alpha }} + \frac{1}{2}m_\alpha
1589     w_\alpha ^2 \left( {x_\alpha - \frac{{g_\alpha }}{{m_\alpha
1590     w_\alpha ^2 }}x} \right)^2 } \right\}}
1591     \]
1592     Since the first two terms of the new Hamiltonian depend only on the
1593     system coordinates, we can get the equations of motion for
1594     Generalized Langevin Dynamics by Hamilton's equations
1595     \ref{introEquation:motionHamiltonianCoordinate,
1596     introEquation:motionHamiltonianMomentum},
1597 tim 2719 \begin{equation}
1598     m\ddot x = - \frac{{\partial W(x)}}{{\partial x}} -
1599     \sum\limits_{\alpha = 1}^N {g_\alpha \left( {x_\alpha -
1600     \frac{{g_\alpha }}{{m_\alpha w_\alpha ^2 }}x} \right)},
1601     \label{introEquation:coorMotionGLE}
1602     \end{equation}
1603     and
1604     \begin{equation}
1605     m\ddot x_\alpha = - m_\alpha w_\alpha ^2 \left( {x_\alpha -
1606     \frac{{g_\alpha }}{{m_\alpha w_\alpha ^2 }}x} \right).
1607     \label{introEquation:bathMotionGLE}
1608     \end{equation}
1609 tim 2696
1610 tim 2719 In order to derive an equation for $x$, the dynamics of the bath
1611     variables $x_\alpha$ must be solved exactly first. As an integral
1612     transform which is particularly useful in solving linear ordinary
1613     differential equations, Laplace transform is the appropriate tool to
1614     solve this problem. The basic idea is to transform the difficult
1615     differential equations into simple algebra problems which can be
1616     solved easily. Then applying inverse Laplace transform, also known
1617     as the Bromwich integral, we can retrieve the solutions of the
1618     original problems.
1619 tim 2696
1620 tim 2719 Let $f(t)$ be a function defined on $ [0,\infty ) $. The Laplace
1621     transform of f(t) is a new function defined as
1622 tim 2696 \[
1623 tim 2719 L(f(t)) \equiv F(p) = \int_0^\infty {f(t)e^{ - pt} dt}
1624 tim 2696 \]
1625 tim 2719 where $p$ is real and $L$ is called the Laplace Transform
1626     Operator. Below are some important properties of Laplace transform
1627 tim 2696
1628 tim 2789 \begin{eqnarray*}
1629     L(x + y) & = & L(x) + L(y) \\
1630     L(ax) & = & aL(x) \\
1631     L(\dot x) & = & pL(x) - px(0) \\
1632     L(\ddot x)& = & p^2 L(x) - px(0) - \dot x(0) \\
1633     L\left( {\int_0^t {g(t - \tau )h(\tau )d\tau } } \right)& = & G(p)H(p) \\
1634     \end{eqnarray*}
1635    
1636    
1637 tim 2719 Applying Laplace transform to the bath coordinates, we obtain
1638 tim 2789 \begin{eqnarray*}
1639     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) \\
1640     L(x_\alpha ) & = & \frac{{\frac{{g_\alpha }}{{\omega _\alpha }}L(x) + px_\alpha (0) + \dot x_\alpha (0)}}{{p^2 + \omega _\alpha ^2 }} \\
1641     \end{eqnarray*}
1642    
1643 tim 2719 By the same way, the system coordinates become
1644 tim 2789 \begin{eqnarray*}
1645     mL(\ddot x) & = & - \frac{1}{p}\frac{{\partial W(x)}}{{\partial x}} \\
1646     & & \mbox{} - \sum\limits_{\alpha = 1}^N {\left\{ { - \frac{{g_\alpha ^2 }}{{m_\alpha \omega _\alpha ^2 }}\frac{p}{{p^2 + \omega _\alpha ^2 }}pL(x) - \frac{p}{{p^2 + \omega _\alpha ^2 }}g_\alpha x_\alpha (0) - \frac{1}{{p^2 + \omega _\alpha ^2 }}g_\alpha \dot x_\alpha (0)} \right\}} \\
1647     \end{eqnarray*}
1648 tim 2696
1649 tim 2719 With the help of some relatively important inverse Laplace
1650     transformations:
1651 tim 2696 \[
1652 tim 2719 \begin{array}{c}
1653     L(\cos at) = \frac{p}{{p^2 + a^2 }} \\
1654     L(\sin at) = \frac{a}{{p^2 + a^2 }} \\
1655     L(1) = \frac{1}{p} \\
1656     \end{array}
1657 tim 2696 \]
1658 tim 2719 , we obtain
1659 tim 2794 \begin{eqnarray*}
1660     m\ddot x & = & - \frac{{\partial W(x)}}{{\partial x}} -
1661 tim 2696 \sum\limits_{\alpha = 1}^N {\left\{ {\left( { - \frac{{g_\alpha ^2
1662     }}{{m_\alpha \omega _\alpha ^2 }}} \right)\int_0^t {\cos (\omega
1663 tim 2794 _\alpha t)\dot x(t - \tau )d\tau } } \right\}} \\
1664     & & + \sum\limits_{\alpha = 1}^N {\left\{ {\left[ {g_\alpha
1665     x_\alpha (0) - \frac{{g_\alpha }}{{m_\alpha \omega _\alpha }}}
1666     \right]\cos (\omega _\alpha t) + \frac{{g_\alpha \dot x_\alpha
1667     (0)}}{{\omega _\alpha }}\sin (\omega _\alpha t)} \right\}}
1668     \end{eqnarray*}
1669     \begin{eqnarray*}
1670     m\ddot x & = & - \frac{{\partial W(x)}}{{\partial x}} - \int_0^t
1671 tim 2696 {\sum\limits_{\alpha = 1}^N {\left( { - \frac{{g_\alpha ^2
1672     }}{{m_\alpha \omega _\alpha ^2 }}} \right)\cos (\omega _\alpha
1673 tim 2794 t)\dot x(t - \tau )d} \tau } \\
1674     & & + \sum\limits_{\alpha = 1}^N {\left\{ {\left[ {g_\alpha
1675     x_\alpha (0) - \frac{{g_\alpha }}{{m_\alpha \omega _\alpha }}}
1676     \right]\cos (\omega _\alpha t) + \frac{{g_\alpha \dot x_\alpha
1677     (0)}}{{\omega _\alpha }}\sin (\omega _\alpha t)} \right\}}
1678     \end{eqnarray*}
1679 tim 2719 Introducing a \emph{dynamic friction kernel}
1680 tim 2696 \begin{equation}
1681 tim 2719 \xi (t) = \sum\limits_{\alpha = 1}^N {\left( { - \frac{{g_\alpha ^2
1682     }}{{m_\alpha \omega _\alpha ^2 }}} \right)\cos (\omega _\alpha t)}
1683     \label{introEquation:dynamicFrictionKernelDefinition}
1684     \end{equation}
1685     and \emph{a random force}
1686     \begin{equation}
1687     R(t) = \sum\limits_{\alpha = 1}^N {\left( {g_\alpha x_\alpha (0)
1688     - \frac{{g_\alpha ^2 }}{{m_\alpha \omega _\alpha ^2 }}x(0)}
1689     \right)\cos (\omega _\alpha t)} + \frac{{\dot x_\alpha
1690     (0)}}{{\omega _\alpha }}\sin (\omega _\alpha t),
1691     \label{introEquation:randomForceDefinition}
1692     \end{equation}
1693     the equation of motion can be rewritten as
1694     \begin{equation}
1695 tim 2696 m\ddot x = - \frac{{\partial W}}{{\partial x}} - \int_0^t {\xi
1696     (t)\dot x(t - \tau )d\tau } + R(t)
1697     \label{introEuqation:GeneralizedLangevinDynamics}
1698     \end{equation}
1699 tim 2719 which is known as the \emph{generalized Langevin equation}.
1700    
1701     \subsubsection{\label{introSection:randomForceDynamicFrictionKernel}Random Force and Dynamic Friction Kernel}
1702    
1703     One may notice that $R(t)$ depends only on initial conditions, which
1704     implies it is completely deterministic within the context of a
1705     harmonic bath. However, it is easy to verify that $R(t)$ is totally
1706     uncorrelated to $x$ and $\dot x$,
1707 tim 2696 \[
1708 tim 2719 \begin{array}{l}
1709     \left\langle {x(t)R(t)} \right\rangle = 0, \\
1710     \left\langle {\dot x(t)R(t)} \right\rangle = 0. \\
1711     \end{array}
1712 tim 2696 \]
1713 tim 2719 This property is what we expect from a truly random process. As long
1714     as the model, which is gaussian distribution in general, chosen for
1715     $R(t)$ is a truly random process, the stochastic nature of the GLE
1716     still remains.
1717 tim 2696
1718 tim 2719 %dynamic friction kernel
1719     The convolution integral
1720 tim 2696 \[
1721 tim 2719 \int_0^t {\xi (t)\dot x(t - \tau )d\tau }
1722 tim 2696 \]
1723 tim 2719 depends on the entire history of the evolution of $x$, which implies
1724     that the bath retains memory of previous motions. In other words,
1725     the bath requires a finite time to respond to change in the motion
1726     of the system. For a sluggish bath which responds slowly to changes
1727     in the system coordinate, we may regard $\xi(t)$ as a constant
1728     $\xi(t) = \Xi_0$. Hence, the convolution integral becomes
1729     \[
1730     \int_0^t {\xi (t)\dot x(t - \tau )d\tau } = \xi _0 (x(t) - x(0))
1731     \]
1732     and Equation \ref{introEuqation:GeneralizedLangevinDynamics} becomes
1733     \[
1734     m\ddot x = - \frac{\partial }{{\partial x}}\left( {W(x) +
1735     \frac{1}{2}\xi _0 (x - x_0 )^2 } \right) + R(t),
1736     \]
1737     which can be used to describe dynamic caging effect. The other
1738     extreme is the bath that responds infinitely quickly to motions in
1739     the system. Thus, $\xi (t)$ can be taken as a $delta$ function in
1740     time:
1741     \[
1742     \xi (t) = 2\xi _0 \delta (t)
1743     \]
1744     Hence, the convolution integral becomes
1745     \[
1746     \int_0^t {\xi (t)\dot x(t - \tau )d\tau } = 2\xi _0 \int_0^t
1747     {\delta (t)\dot x(t - \tau )d\tau } = \xi _0 \dot x(t),
1748     \]
1749     and Equation \ref{introEuqation:GeneralizedLangevinDynamics} becomes
1750     \begin{equation}
1751     m\ddot x = - \frac{{\partial W(x)}}{{\partial x}} - \xi _0 \dot
1752     x(t) + R(t) \label{introEquation:LangevinEquation}
1753     \end{equation}
1754     which is known as the Langevin equation. The static friction
1755     coefficient $\xi _0$ can either be calculated from spectral density
1756     or be determined by Stokes' law for regular shaped particles.A
1757     briefly review on calculating friction tensor for arbitrary shaped
1758 tim 2720 particles is given in Sec.~\ref{introSection:frictionTensor}.
1759 tim 2696
1760     \subsubsection{\label{introSection:secondFluctuationDissipation}The Second Fluctuation Dissipation Theorem}
1761 tim 2719
1762     Defining a new set of coordinates,
1763 tim 2696 \[
1764     q_\alpha (t) = x_\alpha (t) - \frac{1}{{m_\alpha \omega _\alpha
1765     ^2 }}x(0)
1766 tim 2719 \],
1767     we can rewrite $R(T)$ as
1768 tim 2696 \[
1769 tim 2719 R(t) = \sum\limits_{\alpha = 1}^N {g_\alpha q_\alpha (t)}.
1770 tim 2696 \]
1771     And since the $q$ coordinates are harmonic oscillators,
1772 tim 2789
1773     \begin{eqnarray*}
1774     \left\langle {q_\alpha ^2 } \right\rangle & = & \frac{{kT}}{{m_\alpha \omega _\alpha ^2 }} \\
1775     \left\langle {q_\alpha (t)q_\alpha (0)} \right\rangle & = & \left\langle {q_\alpha ^2 (0)} \right\rangle \cos (\omega _\alpha t) \\
1776     \left\langle {q_\alpha (t)q_\beta (0)} \right\rangle & = &\delta _{\alpha \beta } \left\langle {q_\alpha (t)q_\alpha (0)} \right\rangle \\
1777     \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 } } \\
1778     & = &\sum\limits_\alpha {g_\alpha ^2 \left\langle {q_\alpha ^2 (0)} \right\rangle \cos (\omega _\alpha t)} \\
1779     & = &kT\xi (t) \\
1780     \end{eqnarray*}
1781    
1782 tim 2719 Thus, we recover the \emph{second fluctuation dissipation theorem}
1783 tim 2696 \begin{equation}
1784     \xi (t) = \left\langle {R(t)R(0)} \right\rangle
1785 tim 2719 \label{introEquation:secondFluctuationDissipation}.
1786 tim 2696 \end{equation}
1787 tim 2719 In effect, it acts as a constraint on the possible ways in which one
1788     can model the random force and friction kernel.
1789 tim 2696
1790     \subsection{\label{introSection:frictionTensor} Friction Tensor}
1791 tim 2716 Theoretically, the friction kernel can be determined using velocity
1792     autocorrelation function. However, this approach become impractical
1793     when the system become more and more complicate. Instead, various
1794     approaches based on hydrodynamics have been developed to calculate
1795     the friction coefficients. The friction effect is isotropic in
1796 tim 2776 Equation, $\zeta$ can be taken as a scalar. In general, friction
1797     tensor $\Xi$ is a $6\times 6$ matrix given by
1798 tim 2716 \[
1799     \Xi = \left( {\begin{array}{*{20}c}
1800     {\Xi _{}^{tt} } & {\Xi _{}^{rt} } \\
1801     {\Xi _{}^{tr} } & {\Xi _{}^{rr} } \\
1802     \end{array}} \right).
1803     \]
1804     Here, $ {\Xi^{tt} }$ and $ {\Xi^{rr} }$ are translational friction
1805 tim 2718 tensor and rotational resistance (friction) tensor respectively,
1806     while ${\Xi^{tr} }$ is translation-rotation coupling tensor and $
1807     {\Xi^{rt} }$ is rotation-translation coupling tensor. When a
1808     particle moves in a fluid, it may experience friction force or
1809     torque along the opposite direction of the velocity or angular
1810     velocity,
1811 tim 2716 \[
1812     \left( \begin{array}{l}
1813 tim 2718 F_R \\
1814     \tau _R \\
1815 tim 2716 \end{array} \right) = - \left( {\begin{array}{*{20}c}
1816     {\Xi ^{tt} } & {\Xi ^{rt} } \\
1817     {\Xi ^{tr} } & {\Xi ^{rr} } \\
1818     \end{array}} \right)\left( \begin{array}{l}
1819     v \\
1820     w \\
1821     \end{array} \right)
1822     \]
1823 tim 2718 where $F_r$ is the friction force and $\tau _R$ is the friction
1824     toque.
1825 tim 2696
1826 tim 2718 \subsubsection{\label{introSection:resistanceTensorRegular}The Resistance Tensor for Regular Shape}
1827    
1828 tim 2716 For a spherical particle, the translational and rotational friction
1829     constant can be calculated from Stoke's law,
1830     \[
1831     \Xi ^{tt} = \left( {\begin{array}{*{20}c}
1832     {6\pi \eta R} & 0 & 0 \\
1833     0 & {6\pi \eta R} & 0 \\
1834     0 & 0 & {6\pi \eta R} \\
1835     \end{array}} \right)
1836     \]
1837     and
1838     \[
1839     \Xi ^{rr} = \left( {\begin{array}{*{20}c}
1840     {8\pi \eta R^3 } & 0 & 0 \\
1841     0 & {8\pi \eta R^3 } & 0 \\
1842     0 & 0 & {8\pi \eta R^3 } \\
1843     \end{array}} \right)
1844     \]
1845     where $\eta$ is the viscosity of the solvent and $R$ is the
1846     hydrodynamics radius.
1847 tim 2706
1848 tim 2718 Other non-spherical shape, such as cylinder and ellipsoid
1849     \textit{etc}, are widely used as reference for developing new
1850     hydrodynamics theory, because their properties can be calculated
1851     exactly. In 1936, Perrin extended Stokes's law to general ellipsoid,
1852     also called a triaxial ellipsoid, which is given in Cartesian
1853 tim 2789 coordinates by\cite{Perrin1934, Perrin1936}
1854 tim 2716 \[
1855 tim 2718 \frac{{x^2 }}{{a^2 }} + \frac{{y^2 }}{{b^2 }} + \frac{{z^2 }}{{c^2
1856     }} = 1
1857     \]
1858     where the semi-axes are of lengths $a$, $b$, and $c$. Unfortunately,
1859     due to the complexity of the elliptic integral, only the ellipsoid
1860     with the restriction of two axes having to be equal, \textit{i.e.}
1861     prolate($ a \ge b = c$) and oblate ($ a < b = c $), can be solved
1862     exactly. Introducing an elliptic integral parameter $S$ for prolate,
1863     \[
1864 tim 2716 S = \frac{2}{{\sqrt {a^2 - b^2 } }}\ln \frac{{a + \sqrt {a^2 - b^2
1865 tim 2718 } }}{b},
1866 tim 2716 \]
1867 tim 2718 and oblate,
1868 tim 2716 \[
1869     S = \frac{2}{{\sqrt {b^2 - a^2 } }}arctg\frac{{\sqrt {b^2 - a^2 }
1870     }}{a}
1871 tim 2718 \],
1872     one can write down the translational and rotational resistance
1873     tensors
1874 tim 2716 \[
1875     \begin{array}{l}
1876     \Xi _a^{tt} = 16\pi \eta \frac{{a^2 - b^2 }}{{(2a^2 - b^2 )S - 2a}} \\
1877     \Xi _b^{tt} = \Xi _c^{tt} = 32\pi \eta \frac{{a^2 - b^2 }}{{(2a^2 - 3b^2 )S + 2a}} \\
1878 tim 2718 \end{array},
1879 tim 2716 \]
1880 tim 2718 and
1881 tim 2716 \[
1882     \begin{array}{l}
1883     \Xi _a^{rr} = \frac{{32\pi }}{3}\eta \frac{{(a^2 - b^2 )b^2 }}{{2a - b^2 S}} \\
1884     \Xi _b^{rr} = \Xi _c^{rr} = \frac{{32\pi }}{3}\eta \frac{{(a^4 - b^4 )}}{{(2a^2 - b^2 )S - 2a}} \\
1885 tim 2718 \end{array}.
1886 tim 2716 \]
1887    
1888 tim 2718 \subsubsection{\label{introSection:resistanceTensorRegularArbitrary}The Resistance Tensor for Arbitrary Shape}
1889 tim 2716
1890     Unlike spherical and other regular shaped molecules, there is not
1891     analytical solution for friction tensor of any arbitrary shaped
1892     rigid molecules. The ellipsoid of revolution model and general
1893     triaxial ellipsoid model have been used to approximate the
1894     hydrodynamic properties of rigid bodies. However, since the mapping
1895     from all possible ellipsoidal space, $r$-space, to all possible
1896     combination of rotational diffusion coefficients, $D$-space is not
1897 tim 2786 unique\cite{Wegener1979} as well as the intrinsic coupling between
1898 tim 2789 translational and rotational motion of rigid body, general ellipsoid
1899     is not always suitable for modeling arbitrarily shaped rigid
1900     molecule. A number of studies have been devoted to determine the
1901     friction tensor for irregularly shaped rigid bodies using more
1902     advanced method where the molecule of interest was modeled by
1903     combinations of spheres(beads)\cite{Carrasco1999} and the
1904     hydrodynamics properties of the molecule can be calculated using the
1905     hydrodynamic interaction tensor. Let us consider a rigid assembly of
1906     $N$ beads immersed in a continuous medium. Due to hydrodynamics
1907     interaction, the ``net'' velocity of $i$th bead, $v'_i$ is different
1908     than its unperturbed velocity $v_i$,
1909 tim 2716 \[
1910     v'_i = v_i - \sum\limits_{j \ne i} {T_{ij} F_j }
1911     \]
1912     where $F_i$ is the frictional force, and $T_{ij}$ is the
1913     hydrodynamic interaction tensor. The friction force of $i$th bead is
1914     proportional to its ``net'' velocity
1915     \begin{equation}
1916     F_i = \zeta _i v_i - \zeta _i \sum\limits_{j \ne i} {T_{ij} F_j }.
1917     \label{introEquation:tensorExpression}
1918     \end{equation}
1919     This equation is the basis for deriving the hydrodynamic tensor. In
1920     1930, Oseen and Burgers gave a simple solution to Equation
1921     \ref{introEquation:tensorExpression}
1922     \begin{equation}
1923     T_{ij} = \frac{1}{{8\pi \eta r_{ij} }}\left( {I + \frac{{R_{ij}
1924     R_{ij}^T }}{{R_{ij}^2 }}} \right).
1925     \label{introEquation:oseenTensor}
1926     \end{equation}
1927     Here $R_{ij}$ is the distance vector between bead $i$ and bead $j$.
1928     A second order expression for element of different size was
1929 tim 2789 introduced by Rotne and Prager\cite{Rotne1969} and improved by
1930     Garc\'{i}a de la Torre and Bloomfield\cite{Torre1977},
1931 tim 2716 \begin{equation}
1932     T_{ij} = \frac{1}{{8\pi \eta R_{ij} }}\left[ {\left( {I +
1933     \frac{{R_{ij} R_{ij}^T }}{{R_{ij}^2 }}} \right) + R\frac{{\sigma
1934     _i^2 + \sigma _j^2 }}{{r_{ij}^2 }}\left( {\frac{I}{3} -
1935     \frac{{R_{ij} R_{ij}^T }}{{R_{ij}^2 }}} \right)} \right].
1936     \label{introEquation:RPTensorNonOverlapped}
1937     \end{equation}
1938     Both of the Equation \ref{introEquation:oseenTensor} and Equation
1939     \ref{introEquation:RPTensorNonOverlapped} have an assumption $R_{ij}
1940     \ge \sigma _i + \sigma _j$. An alternative expression for
1941     overlapping beads with the same radius, $\sigma$, is given by
1942     \begin{equation}
1943     T_{ij} = \frac{1}{{6\pi \eta R_{ij} }}\left[ {\left( {1 -
1944     \frac{2}{{32}}\frac{{R_{ij} }}{\sigma }} \right)I +
1945     \frac{2}{{32}}\frac{{R_{ij} R_{ij}^T }}{{R_{ij} \sigma }}} \right]
1946     \label{introEquation:RPTensorOverlapped}
1947     \end{equation}
1948    
1949 tim 2718 To calculate the resistance tensor at an arbitrary origin $O$, we
1950     construct a $3N \times 3N$ matrix consisting of $N \times N$
1951     $B_{ij}$ blocks
1952     \begin{equation}
1953 tim 2716 B = \left( {\begin{array}{*{20}c}
1954 tim 2718 {B_{11} } & \ldots & {B_{1N} } \\
1955 tim 2716 \vdots & \ddots & \vdots \\
1956 tim 2718 {B_{N1} } & \cdots & {B_{NN} } \\
1957     \end{array}} \right),
1958     \end{equation}
1959     where $B_{ij}$ is given by
1960     \[
1961     B_{ij} = \delta _{ij} \frac{I}{{6\pi \eta R}} + (1 - \delta _{ij}
1962     )T_{ij}
1963 tim 2716 \]
1964 tim 2719 where $\delta _{ij}$ is Kronecker delta function. Inverting matrix
1965 tim 2718 $B$, we obtain
1966 tim 2716
1967     \[
1968     C = B^{ - 1} = \left( {\begin{array}{*{20}c}
1969     {C_{11} } & \ldots & {C_{1N} } \\
1970     \vdots & \ddots & \vdots \\
1971     {C_{N1} } & \cdots & {C_{NN} } \\
1972     \end{array}} \right)
1973     \]
1974 tim 2718 , which can be partitioned into $N \times N$ $3 \times 3$ block
1975     $C_{ij}$. With the help of $C_{ij}$ and skew matrix $U_i$
1976     \[
1977     U_i = \left( {\begin{array}{*{20}c}
1978     0 & { - z_i } & {y_i } \\
1979     {z_i } & 0 & { - x_i } \\
1980     { - y_i } & {x_i } & 0 \\
1981     \end{array}} \right)
1982     \]
1983     where $x_i$, $y_i$, $z_i$ are the components of the vector joining
1984     bead $i$ and origin $O$. Hence, the elements of resistance tensor at
1985     arbitrary origin $O$ can be written as
1986 tim 2716 \begin{equation}
1987     \begin{array}{l}
1988     \Xi _{}^{tt} = \sum\limits_i {\sum\limits_j {C_{ij} } } , \\
1989     \Xi _{}^{tr} = \Xi _{}^{rt} = \sum\limits_i {\sum\limits_j {U_i C_{ij} } } , \\
1990     \Xi _{}^{rr} = - \sum\limits_i {\sum\limits_j {U_i C_{ij} } } U_j \\
1991     \end{array}
1992 tim 2718 \label{introEquation:ResistanceTensorArbitraryOrigin}
1993 tim 2716 \end{equation}
1994 tim 2718
1995     The resistance tensor depends on the origin to which they refer. The
1996     proper location for applying friction force is the center of
1997     resistance (reaction), at which the trace of rotational resistance
1998     tensor, $ \Xi ^{rr}$ reaches minimum. Mathematically, the center of
1999     resistance is defined as an unique point of the rigid body at which
2000     the translation-rotation coupling tensor are symmetric,
2001     \begin{equation}
2002     \Xi^{tr} = \left( {\Xi^{tr} } \right)^T
2003     \label{introEquation:definitionCR}
2004     \end{equation}
2005     Form Equation \ref{introEquation:ResistanceTensorArbitraryOrigin},
2006     we can easily find out that the translational resistance tensor is
2007     origin independent, while the rotational resistance tensor and
2008 tim 2719 translation-rotation coupling resistance tensor depend on the
2009 tim 2718 origin. Given resistance tensor at an arbitrary origin $O$, and a
2010     vector ,$r_{OP}(x_{OP}, y_{OP}, z_{OP})$, from $O$ to $P$, we can
2011     obtain the resistance tensor at $P$ by
2012     \begin{equation}
2013     \begin{array}{l}
2014     \Xi _P^{tt} = \Xi _O^{tt} \\
2015     \Xi _P^{tr} = \Xi _P^{rt} = \Xi _O^{tr} - U_{OP} \Xi _O^{tt} \\
2016     \Xi _P^{rr} = \Xi _O^{rr} - U_{OP} \Xi _O^{tt} U_{OP} + \Xi _O^{tr} U_{OP} - U_{OP} \Xi _O^{tr} ^{^T } \\
2017     \end{array}
2018     \label{introEquation:resistanceTensorTransformation}
2019     \end{equation}
2020 tim 2716 where
2021     \[
2022 tim 2718 U_{OP} = \left( {\begin{array}{*{20}c}
2023     0 & { - z_{OP} } & {y_{OP} } \\
2024     {z_i } & 0 & { - x_{OP} } \\
2025     { - y_{OP} } & {x_{OP} } & 0 \\
2026 tim 2716 \end{array}} \right)
2027     \]
2028 tim 2718 Using Equations \ref{introEquation:definitionCR} and
2029     \ref{introEquation:resistanceTensorTransformation}, one can locate
2030     the position of center of resistance,
2031 tim 2789 \begin{eqnarray*}
2032     \left( \begin{array}{l}
2033     x_{OR} \\
2034     y_{OR} \\
2035     z_{OR} \\
2036     \end{array} \right) & = &\left( {\begin{array}{*{20}c}
2037     {(\Xi _O^{rr} )_{yy} + (\Xi _O^{rr} )_{zz} } & { - (\Xi _O^{rr} )_{xy} } & { - (\Xi _O^{rr} )_{xz} } \\
2038     { - (\Xi _O^{rr} )_{xy} } & {(\Xi _O^{rr} )_{zz} + (\Xi _O^{rr} )_{xx} } & { - (\Xi _O^{rr} )_{yz} } \\
2039     { - (\Xi _O^{rr} )_{xz} } & { - (\Xi _O^{rr} )_{yz} } & {(\Xi _O^{rr} )_{xx} + (\Xi _O^{rr} )_{yy} } \\
2040     \end{array}} \right)^{ - 1} \\
2041     & & \left( \begin{array}{l}
2042     (\Xi _O^{tr} )_{yz} - (\Xi _O^{tr} )_{zy} \\
2043     (\Xi _O^{tr} )_{zx} - (\Xi _O^{tr} )_{xz} \\
2044     (\Xi _O^{tr} )_{xy} - (\Xi _O^{tr} )_{yx} \\
2045     \end{array} \right) \\
2046     \end{eqnarray*}
2047    
2048    
2049    
2050 tim 2718 where $x_OR$, $y_OR$, $z_OR$ are the components of the vector
2051     joining center of resistance $R$ and origin $O$.