ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/tengDissertation/Introduction.tex
Revision: 2898
Committed: Tue Jun 27 03:22:42 2006 UTC (18 years, 2 months ago) by tim
Content type: application/x-tex
File size: 75381 byte(s)
Log Message:
more corrections

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 tim 2819 behind classical mechanics. Firstly, one can determine the state of
10 tim 2692 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 tim 2819 Newton's first law defines a class of inertial frames. Inertial
21 tim 2694 frames are reference frames where a particle not interacting with
22     other bodies will move with constant speed in the same direction.
23 tim 2819 With respect to inertial frames, Newton's second law has the form
24 tim 2694 \begin{equation}
25 tim 2819 F = \frac {dp}{dt} = \frac {mdv}{dt}
26 tim 2694 \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 2819 Newton's third law states that
33 tim 2694 \begin{equation}
34 tim 2898 F_{ij} = -F_{ji}.
35 tim 2694 \label{introEquation:newtonThirdLaw}
36     \end{equation}
37     Conservation laws of Newtonian Mechanics play very important roles
38     in solving mechanics problems. The linear momentum of a particle is
39     conserved if it is free or it experiences no force. The second
40     conservation theorem concerns the angular momentum of a particle.
41     The angular momentum $L$ of a particle with respect to an origin
42     from which $r$ is measured is defined to be
43     \begin{equation}
44     L \equiv r \times p \label{introEquation:angularMomentumDefinition}
45     \end{equation}
46     The torque $\tau$ with respect to the same origin is defined to be
47     \begin{equation}
48 tim 2819 \tau \equiv r \times F \label{introEquation:torqueDefinition}
49 tim 2694 \end{equation}
50     Differentiating Eq.~\ref{introEquation:angularMomentumDefinition},
51     \[
52     \dot L = \frac{d}{{dt}}(r \times p) = (\dot r \times p) + (r \times
53     \dot p)
54     \]
55     since
56     \[
57     \dot r \times p = \dot r \times mv = m\dot r \times \dot r \equiv 0
58     \]
59     thus,
60     \begin{equation}
61 tim 2819 \dot L = r \times \dot p = \tau
62 tim 2694 \end{equation}
63     If there are no external torques acting on a body, the angular
64     momentum of it is conserved. The last conservation theorem state
65 tim 2696 that if all forces are conservative, Energy
66     \begin{equation}E = T + V \label{introEquation:energyConservation}
67     \end{equation}
68     is conserved. All of these conserved quantities are
69     important factors to determine the quality of numerical integration
70 tim 2819 schemes for rigid bodies \cite{Dullweber1997}.
71 tim 2694
72 tim 2693 \subsection{\label{introSection:lagrangian}Lagrangian Mechanics}
73 tim 2692
74 tim 2819 Newtonian Mechanics suffers from two important limitations: motions
75 tim 2895 can only be described in cartesian coordinate systems. Moreover, it
76     becomes impossible to predict analytically the properties of the
77 tim 2819 system even if we know all of the details of the interaction. In
78     order to overcome some of the practical difficulties which arise in
79     attempts to apply Newton's equation to complex system, approximate
80     numerical procedures may be developed.
81 tim 2692
82 tim 2819 \subsubsection{\label{introSection:halmiltonPrinciple}\textbf{Hamilton's
83     Principle}}
84 tim 2692
85     Hamilton introduced the dynamical principle upon which it is
86 tim 2819 possible to base all of mechanics and most of classical physics.
87 tim 2898 Hamilton's Principle may be stated as follows: the actual
88     trajectory, along which a dynamical system may move from one point
89     to another within a specified time, is derived by finding the path
90     which minimizes the time integral of the difference between the
91     kinetic, $K$, and potential energies, $U$.
92 tim 2692 \begin{equation}
93     \delta \int_{t_1 }^{t_2 } {(K - U)dt = 0} ,
94 tim 2693 \label{introEquation:halmitonianPrinciple1}
95 tim 2692 \end{equation}
96     For simple mechanical systems, where the forces acting on the
97 tim 2819 different parts are derivable from a potential, the Lagrangian
98     function $L$ can be defined as the difference between the kinetic
99     energy of the system and its potential energy,
100 tim 2692 \begin{equation}
101     L \equiv K - U = L(q_i ,\dot q_i ) ,
102     \label{introEquation:lagrangianDef}
103     \end{equation}
104     then Eq.~\ref{introEquation:halmitonianPrinciple1} becomes
105     \begin{equation}
106 tim 2693 \delta \int_{t_1 }^{t_2 } {L dt = 0} ,
107     \label{introEquation:halmitonianPrinciple2}
108 tim 2692 \end{equation}
109    
110 tim 2819 \subsubsection{\label{introSection:equationOfMotionLagrangian}\textbf{The
111     Equations of Motion in Lagrangian Mechanics}}
112 tim 2692
113 tim 2850 For a system of $f$ degrees of freedom, the equations of motion in
114     the Lagrangian form is
115 tim 2692 \begin{equation}
116     \frac{d}{{dt}}\frac{{\partial L}}{{\partial \dot q_i }} -
117     \frac{{\partial L}}{{\partial q_i }} = 0,{\rm{ }}i = 1, \ldots,f
118 tim 2693 \label{introEquation:eqMotionLagrangian}
119 tim 2692 \end{equation}
120     where $q_{i}$ is generalized coordinate and $\dot{q_{i}}$ is
121     generalized velocity.
122    
123 tim 2693 \subsection{\label{introSection:hamiltonian}Hamiltonian Mechanics}
124 tim 2692
125     Arising from Lagrangian Mechanics, Hamiltonian Mechanics was
126     introduced by William Rowan Hamilton in 1833 as a re-formulation of
127     classical mechanics. If the potential energy of a system is
128 tim 2819 independent of velocities, the momenta can be defined as
129 tim 2692 \begin{equation}
130     p_i = \frac{\partial L}{\partial \dot q_i}
131     \label{introEquation:generalizedMomenta}
132     \end{equation}
133 tim 2693 The Lagrange equations of motion are then expressed by
134 tim 2692 \begin{equation}
135 tim 2693 p_i = \frac{{\partial L}}{{\partial q_i }}
136     \label{introEquation:generalizedMomentaDot}
137     \end{equation}
138     With the help of the generalized momenta, we may now define a new
139     quantity $H$ by the equation
140     \begin{equation}
141     H = \sum\limits_k {p_k \dot q_k } - L ,
142 tim 2692 \label{introEquation:hamiltonianDefByLagrangian}
143     \end{equation}
144     where $ \dot q_1 \ldots \dot q_f $ are generalized velocities and
145 tim 2898 $L$ is the Lagrangian function for the system. Differentiating
146     Eq.~\ref{introEquation:hamiltonianDefByLagrangian}, one can obtain
147 tim 2693 \begin{equation}
148     dH = \sum\limits_k {\left( {p_k d\dot q_k + \dot q_k dp_k -
149     \frac{{\partial L}}{{\partial q_k }}dq_k - \frac{{\partial
150     L}}{{\partial \dot q_k }}d\dot q_k } \right)} - \frac{{\partial
151     L}}{{\partial t}}dt \label{introEquation:diffHamiltonian1}
152     \end{equation}
153     Making use of Eq.~\ref{introEquation:generalizedMomenta}, the
154     second and fourth terms in the parentheses cancel. Therefore,
155     Eq.~\ref{introEquation:diffHamiltonian1} can be rewritten as
156     \begin{equation}
157     dH = \sum\limits_k {\left( {\dot q_k dp_k - \dot p_k dq_k }
158     \right)} - \frac{{\partial L}}{{\partial t}}dt
159     \label{introEquation:diffHamiltonian2}
160     \end{equation}
161     By identifying the coefficients of $dq_k$, $dp_k$ and dt, we can
162     find
163     \begin{equation}
164 tim 2819 \frac{{\partial H}}{{\partial p_k }} = \dot {q_k}
165 tim 2693 \label{introEquation:motionHamiltonianCoordinate}
166     \end{equation}
167     \begin{equation}
168 tim 2819 \frac{{\partial H}}{{\partial q_k }} = - \dot {p_k}
169 tim 2693 \label{introEquation:motionHamiltonianMomentum}
170     \end{equation}
171     and
172     \begin{equation}
173     \frac{{\partial H}}{{\partial t}} = - \frac{{\partial L}}{{\partial
174     t}}
175     \label{introEquation:motionHamiltonianTime}
176     \end{equation}
177     Eq.~\ref{introEquation:motionHamiltonianCoordinate} and
178     Eq.~\ref{introEquation:motionHamiltonianMomentum} are Hamilton's
179     equation of motion. Due to their symmetrical formula, they are also
180 tim 2786 known as the canonical equations of motions \cite{Goldstein2001}.
181 tim 2693
182 tim 2692 An important difference between Lagrangian approach and the
183     Hamiltonian approach is that the Lagrangian is considered to be a
184 tim 2819 function of the generalized velocities $\dot q_i$ and coordinates
185     $q_i$, while the Hamiltonian is considered to be a function of the
186     generalized momenta $p_i$ and the conjugate coordinates $q_i$.
187     Hamiltonian Mechanics is more appropriate for application to
188     statistical mechanics and quantum mechanics, since it treats the
189     coordinate and its time derivative as independent variables and it
190     only works with 1st-order differential equations\cite{Marion1990}.
191 tim 2692
192 tim 2696 In Newtonian Mechanics, a system described by conservative forces
193     conserves the total energy \ref{introEquation:energyConservation}.
194     It follows that Hamilton's equations of motion conserve the total
195     Hamiltonian.
196     \begin{equation}
197     \frac{{dH}}{{dt}} = \sum\limits_i {\left( {\frac{{\partial
198     H}}{{\partial q_i }}\dot q_i + \frac{{\partial H}}{{\partial p_i
199     }}\dot p_i } \right)} = \sum\limits_i {\left( {\frac{{\partial
200     H}}{{\partial q_i }}\frac{{\partial H}}{{\partial p_i }} -
201     \frac{{\partial H}}{{\partial p_i }}\frac{{\partial H}}{{\partial
202 tim 2698 q_i }}} \right) = 0} \label{introEquation:conserveHalmitonian}
203 tim 2696 \end{equation}
204    
205 tim 2693 \section{\label{introSection:statisticalMechanics}Statistical
206     Mechanics}
207 tim 2692
208 tim 2694 The thermodynamic behaviors and properties of Molecular Dynamics
209 tim 2692 simulation are governed by the principle of Statistical Mechanics.
210     The following section will give a brief introduction to some of the
211 tim 2700 Statistical Mechanics concepts and theorem presented in this
212     dissertation.
213 tim 2692
214 tim 2700 \subsection{\label{introSection:ensemble}Phase Space and Ensemble}
215 tim 2692
216 tim 2700 Mathematically, phase space is the space which represents all
217     possible states. Each possible state of the system corresponds to
218     one unique point in the phase space. For mechanical systems, the
219     phase space usually consists of all possible values of position and
220 tim 2819 momentum variables. Consider a dynamic system of $f$ particles in a
221     cartesian space, where each of the $6f$ coordinates and momenta is
222     assigned to one of $6f$ mutually orthogonal axes, the phase space of
223 tim 2888 this system is a $6f$ dimensional space. A point, $x = (\rightarrow
224     q_1 , \ldots ,\rightarrow q_f ,\rightarrow p_1 , \ldots ,\rightarrow
225     p_f )$, with a unique set of values of $6f$ coordinates and momenta
226     is a phase space vector.
227     %%%fix me
228 tim 2700
229 tim 2888 In statistical mechanics, the condition of an ensemble at any time
230 tim 2700 can be regarded as appropriately specified by the density $\rho$
231     with which representative points are distributed over the phase
232 tim 2819 space. The density distribution for an ensemble with $f$ degrees of
233     freedom is defined as,
234 tim 2700 \begin{equation}
235     \rho = \rho (q_1 , \ldots ,q_f ,p_1 , \ldots ,p_f ,t).
236     \label{introEquation:densityDistribution}
237     \end{equation}
238     Governed by the principles of mechanics, the phase points change
239 tim 2819 their locations which would change the density at any time at phase
240     space. Hence, the density distribution is also to be taken as a
241 tim 2700 function of the time.
242    
243     The number of systems $\delta N$ at time $t$ can be determined by,
244     \begin{equation}
245     \delta N = \rho (q,p,t)dq_1 \ldots dq_f dp_1 \ldots dp_f.
246     \label{introEquation:deltaN}
247     \end{equation}
248 tim 2819 Assuming a large enough population of systems, we can sufficiently
249     approximate $\delta N$ without introducing discontinuity when we go
250     from one region in the phase space to another. By integrating over
251     the whole phase space,
252 tim 2700 \begin{equation}
253     N = \int { \ldots \int {\rho (q,p,t)dq_1 } ...dq_f dp_1 } ...dp_f
254     \label{introEquation:totalNumberSystem}
255     \end{equation}
256     gives us an expression for the total number of the systems. Hence,
257     the probability per unit in the phase space can be obtained by,
258     \begin{equation}
259     \frac{{\rho (q,p,t)}}{N} = \frac{{\rho (q,p,t)}}{{\int { \ldots \int
260     {\rho (q,p,t)dq_1 } ...dq_f dp_1 } ...dp_f }}.
261     \label{introEquation:unitProbability}
262     \end{equation}
263 tim 2850 With the help of Eq.~\ref{introEquation:unitProbability} and the
264     knowledge of the system, it is possible to calculate the average
265 tim 2700 value of any desired quantity which depends on the coordinates and
266     momenta of the system. Even when the dynamics of the real system is
267     complex, or stochastic, or even discontinuous, the average
268 tim 2819 properties of the ensemble of possibilities as a whole remaining
269     well defined. For a classical system in thermal equilibrium with its
270     environment, the ensemble average of a mechanical quantity, $\langle
271     A(q , p) \rangle_t$, takes the form of an integral over the phase
272     space of the system,
273 tim 2700 \begin{equation}
274     \langle A(q , p) \rangle_t = \frac{{\int { \ldots \int {A(q,p)\rho
275     (q,p,t)dq_1 } ...dq_f dp_1 } ...dp_f }}{{\int { \ldots \int {\rho
276     (q,p,t)dq_1 } ...dq_f dp_1 } ...dp_f }}
277     \label{introEquation:ensembelAverage}
278     \end{equation}
279    
280     There are several different types of ensembles with different
281     statistical characteristics. As a function of macroscopic
282 tim 2819 parameters, such as temperature \textit{etc}, the partition function
283     can be used to describe the statistical properties of a system in
284 tim 2898 thermodynamic equilibrium. As an ensemble of systems, each of which
285     is known to be thermally isolated and conserve energy, the
286     Microcanonical ensemble (NVE) has a partition function like,
287 tim 2700 \begin{equation}
288 tim 2706 \Omega (N,V,E) = e^{\beta TS} \label{introEquation:NVEPartition}.
289 tim 2700 \end{equation}
290 tim 2850 A canonical ensemble (NVT)is an ensemble of systems, each of which
291 tim 2700 can share its energy with a large heat reservoir. The distribution
292     of the total energy amongst the possible dynamical states is given
293     by the partition function,
294     \begin{equation}
295     \Omega (N,V,T) = e^{ - \beta A}
296     \label{introEquation:NVTPartition}
297     \end{equation}
298     Here, $A$ is the Helmholtz free energy which is defined as $ A = U -
299 tim 2819 TS$. Since most experiments are carried out under constant pressure
300 tim 2850 condition, the isothermal-isobaric ensemble (NPT) plays a very
301 tim 2819 important role in molecular simulations. The isothermal-isobaric
302     ensemble allow the system to exchange energy with a heat bath of
303     temperature $T$ and to change the volume as well. Its partition
304     function is given as
305 tim 2700 \begin{equation}
306     \Delta (N,P,T) = - e^{\beta G}.
307     \label{introEquation:NPTPartition}
308     \end{equation}
309     Here, $G = U - TS + PV$, and $G$ is called Gibbs free energy.
310    
311     \subsection{\label{introSection:liouville}Liouville's theorem}
312    
313 tim 2819 Liouville's theorem is the foundation on which statistical mechanics
314     rests. It describes the time evolution of the phase space
315 tim 2700 distribution function. In order to calculate the rate of change of
316 tim 2850 $\rho$, we begin from Eq.~\ref{introEquation:deltaN}. If we consider
317     the two faces perpendicular to the $q_1$ axis, which are located at
318     $q_1$ and $q_1 + \delta q_1$, the number of phase points leaving the
319     opposite face is given by the expression,
320 tim 2700 \begin{equation}
321     \left( {\rho + \frac{{\partial \rho }}{{\partial q_1 }}\delta q_1 }
322     \right)\left( {\dot q_1 + \frac{{\partial \dot q_1 }}{{\partial q_1
323     }}\delta q_1 } \right)\delta q_2 \ldots \delta q_f \delta p_1
324     \ldots \delta p_f .
325     \end{equation}
326     Summing all over the phase space, we obtain
327     \begin{equation}
328     \frac{{d(\delta N)}}{{dt}} = - \sum\limits_{i = 1}^f {\left[ {\rho
329     \left( {\frac{{\partial \dot q_i }}{{\partial q_i }} +
330     \frac{{\partial \dot p_i }}{{\partial p_i }}} \right) + \left(
331     {\frac{{\partial \rho }}{{\partial q_i }}\dot q_i + \frac{{\partial
332     \rho }}{{\partial p_i }}\dot p_i } \right)} \right]} \delta q_1
333     \ldots \delta q_f \delta p_1 \ldots \delta p_f .
334     \end{equation}
335     Differentiating the equations of motion in Hamiltonian formalism
336     (\ref{introEquation:motionHamiltonianCoordinate},
337     \ref{introEquation:motionHamiltonianMomentum}), we can show,
338     \begin{equation}
339     \sum\limits_i {\left( {\frac{{\partial \dot q_i }}{{\partial q_i }}
340     + \frac{{\partial \dot p_i }}{{\partial p_i }}} \right)} = 0 ,
341     \end{equation}
342     which cancels the first terms of the right hand side. Furthermore,
343 tim 2819 dividing $ \delta q_1 \ldots \delta q_f \delta p_1 \ldots \delta
344 tim 2700 p_f $ in both sides, we can write out Liouville's theorem in a
345     simple form,
346     \begin{equation}
347     \frac{{\partial \rho }}{{\partial t}} + \sum\limits_{i = 1}^f
348     {\left( {\frac{{\partial \rho }}{{\partial q_i }}\dot q_i +
349     \frac{{\partial \rho }}{{\partial p_i }}\dot p_i } \right)} = 0 .
350     \label{introEquation:liouvilleTheorem}
351     \end{equation}
352    
353     Liouville's theorem states that the distribution function is
354     constant along any trajectory in phase space. In classical
355 tim 2850 statistical mechanics, since the number of members in an ensemble is
356     huge and constant, we can assume the local density has no reason
357     (other than classical mechanics) to change,
358 tim 2700 \begin{equation}
359     \frac{{\partial \rho }}{{\partial t}} = 0.
360     \label{introEquation:stationary}
361     \end{equation}
362     In such stationary system, the density of distribution $\rho$ can be
363     connected to the Hamiltonian $H$ through Maxwell-Boltzmann
364     distribution,
365     \begin{equation}
366     \rho \propto e^{ - \beta H}
367     \label{introEquation:densityAndHamiltonian}
368     \end{equation}
369    
370 tim 2819 \subsubsection{\label{introSection:phaseSpaceConservation}\textbf{Conservation of Phase Space}}
371 tim 2702 Lets consider a region in the phase space,
372     \begin{equation}
373     \delta v = \int { \ldots \int {dq_1 } ...dq_f dp_1 } ..dp_f .
374     \end{equation}
375     If this region is small enough, the density $\rho$ can be regarded
376 tim 2819 as uniform over the whole integral. Thus, the number of phase points
377     inside this region is given by,
378 tim 2702 \begin{equation}
379     \delta N = \rho \delta v = \rho \int { \ldots \int {dq_1 } ...dq_f
380     dp_1 } ..dp_f.
381     \end{equation}
382    
383     \begin{equation}
384     \frac{{d(\delta N)}}{{dt}} = \frac{{d\rho }}{{dt}}\delta v + \rho
385     \frac{d}{{dt}}(\delta v) = 0.
386     \end{equation}
387     With the help of stationary assumption
388     (\ref{introEquation:stationary}), we obtain the principle of the
389 tim 2819 \emph{conservation of volume in phase space},
390 tim 2702 \begin{equation}
391     \frac{d}{{dt}}(\delta v) = \frac{d}{{dt}}\int { \ldots \int {dq_1 }
392     ...dq_f dp_1 } ..dp_f = 0.
393     \label{introEquation:volumePreserving}
394     \end{equation}
395    
396 tim 2819 \subsubsection{\label{introSection:liouvilleInOtherForms}\textbf{Liouville's Theorem in Other Forms}}
397 tim 2702
398 tim 2700 Liouville's theorem can be expresses in a variety of different forms
399     which are convenient within different contexts. For any two function
400     $F$ and $G$ of the coordinates and momenta of a system, the Poisson
401     bracket ${F, G}$ is defined as
402     \begin{equation}
403     \left\{ {F,G} \right\} = \sum\limits_i {\left( {\frac{{\partial
404     F}}{{\partial q_i }}\frac{{\partial G}}{{\partial p_i }} -
405     \frac{{\partial F}}{{\partial p_i }}\frac{{\partial G}}{{\partial
406     q_i }}} \right)}.
407     \label{introEquation:poissonBracket}
408     \end{equation}
409     Substituting equations of motion in Hamiltonian formalism(
410 tim 2850 Eq.~\ref{introEquation:motionHamiltonianCoordinate} ,
411     Eq.~\ref{introEquation:motionHamiltonianMomentum} ) into
412     (Eq.~\ref{introEquation:liouvilleTheorem}), we can rewrite
413     Liouville's theorem using Poisson bracket notion,
414 tim 2700 \begin{equation}
415     \left( {\frac{{\partial \rho }}{{\partial t}}} \right) = - \left\{
416     {\rho ,H} \right\}.
417     \label{introEquation:liouvilleTheromInPoissin}
418     \end{equation}
419     Moreover, the Liouville operator is defined as
420     \begin{equation}
421     iL = \sum\limits_{i = 1}^f {\left( {\frac{{\partial H}}{{\partial
422     p_i }}\frac{\partial }{{\partial q_i }} - \frac{{\partial
423     H}}{{\partial q_i }}\frac{\partial }{{\partial p_i }}} \right)}
424     \label{introEquation:liouvilleOperator}
425     \end{equation}
426     In terms of Liouville operator, Liouville's equation can also be
427     expressed as
428     \begin{equation}
429     \left( {\frac{{\partial \rho }}{{\partial t}}} \right) = - iL\rho
430     \label{introEquation:liouvilleTheoremInOperator}
431     \end{equation}
432    
433 tim 2693 \subsection{\label{introSection:ergodic}The Ergodic Hypothesis}
434 tim 2692
435 tim 2695 Various thermodynamic properties can be calculated from Molecular
436     Dynamics simulation. By comparing experimental values with the
437     calculated properties, one can determine the accuracy of the
438 tim 2819 simulation and the quality of the underlying model. However, both
439     experiments and computer simulations are usually performed during a
440 tim 2695 certain time interval and the measurements are averaged over a
441     period of them which is different from the average behavior of
442 tim 2819 many-body system in Statistical Mechanics. Fortunately, the Ergodic
443     Hypothesis makes a connection between time average and the ensemble
444     average. It states that the time average and average over the
445 tim 2786 statistical ensemble are identical \cite{Frenkel1996, Leach2001}.
446 tim 2695 \begin{equation}
447 tim 2700 \langle A(q , p) \rangle_t = \mathop {\lim }\limits_{t \to \infty }
448     \frac{1}{t}\int\limits_0^t {A(q(t),p(t))dt = \int\limits_\Gamma
449     {A(q(t),p(t))} } \rho (q(t), p(t)) dqdp
450 tim 2695 \end{equation}
451 tim 2700 where $\langle A(q , p) \rangle_t$ is an equilibrium value of a
452     physical quantity and $\rho (p(t), q(t))$ is the equilibrium
453     distribution function. If an observation is averaged over a
454     sufficiently long time (longer than relaxation time), all accessible
455     microstates in phase space are assumed to be equally probed, giving
456     a properly weighted statistical average. This allows the researcher
457     freedom of choice when deciding how best to measure a given
458     observable. In case an ensemble averaged approach sounds most
459 tim 2786 reasonable, the Monte Carlo techniques\cite{Metropolis1949} can be
460 tim 2700 utilized. Or if the system lends itself to a time averaging
461     approach, the Molecular Dynamics techniques in
462     Sec.~\ref{introSection:molecularDynamics} will be the best
463     choice\cite{Frenkel1996}.
464 tim 2694
465 tim 2697 \section{\label{introSection:geometricIntegratos}Geometric Integrators}
466 tim 2819 A variety of numerical integrators have been proposed to simulate
467     the motions of atoms in MD simulation. They usually begin with
468     initial conditionals and move the objects in the direction governed
469     by the differential equations. However, most of them ignore the
470     hidden physical laws contained within the equations. Since 1990,
471     geometric integrators, which preserve various phase-flow invariants
472     such as symplectic structure, volume and time reversal symmetry, are
473     developed to address this issue\cite{Dullweber1997, McLachlan1998,
474 tim 2872 Leimkuhler1999}. The velocity Verlet method, which happens to be a
475 tim 2819 simple example of symplectic integrator, continues to gain
476     popularity in the molecular dynamics community. This fact can be
477     partly explained by its geometric nature.
478 tim 2697
479 tim 2819 \subsection{\label{introSection:symplecticManifold}Symplectic Manifolds}
480     A \emph{manifold} is an abstract mathematical space. It looks
481     locally like Euclidean space, but when viewed globally, it may have
482     more complicated structure. A good example of manifold is the
483     surface of Earth. It seems to be flat locally, but it is round if
484     viewed as a whole. A \emph{differentiable manifold} (also known as
485     \emph{smooth manifold}) is a manifold on which it is possible to
486     apply calculus on \emph{differentiable manifold}. A \emph{symplectic
487     manifold} is defined as a pair $(M, \omega)$ which consists of a
488 tim 2697 \emph{differentiable manifold} $M$ and a close, non-degenerated,
489     bilinear symplectic form, $\omega$. A symplectic form on a vector
490     space $V$ is a function $\omega(x, y)$ which satisfies
491     $\omega(\lambda_1x_1+\lambda_2x_2, y) = \lambda_1\omega(x_1, y)+
492     \lambda_2\omega(x_2, y)$, $\omega(x, y) = - \omega(y, x)$ and
493 tim 2819 $\omega(x, x) = 0$. The cross product operation in vector field is
494     an example of symplectic form.
495 tim 2697
496 tim 2819 One of the motivations to study \emph{symplectic manifolds} in
497 tim 2697 Hamiltonian Mechanics is that a symplectic manifold can represent
498     all possible configurations of the system and the phase space of the
499     system can be described by it's cotangent bundle. Every symplectic
500     manifold is even dimensional. For instance, in Hamilton equations,
501     coordinate and momentum always appear in pairs.
502    
503 tim 2698 \subsection{\label{introSection:ODE}Ordinary Differential Equations}
504 tim 2697
505 tim 2819 For an ordinary differential system defined as
506 tim 2698 \begin{equation}
507     \dot x = f(x)
508     \end{equation}
509 tim 2819 where $x = x(q,p)^T$, this system is a canonical Hamiltonian, if
510 tim 2698 \begin{equation}
511 tim 2699 f(r) = J\nabla _x H(r).
512 tim 2698 \end{equation}
513     $H = H (q, p)$ is Hamiltonian function and $J$ is the skew-symmetric
514     matrix
515     \begin{equation}
516     J = \left( {\begin{array}{*{20}c}
517     0 & I \\
518     { - I} & 0 \\
519     \end{array}} \right)
520     \label{introEquation:canonicalMatrix}
521     \end{equation}
522     where $I$ is an identity matrix. Using this notation, Hamiltonian
523     system can be rewritten as,
524     \begin{equation}
525     \frac{d}{{dt}}x = J\nabla _x H(x)
526     \label{introEquation:compactHamiltonian}
527     \end{equation}In this case, $f$ is
528     called a \emph{Hamiltonian vector field}.
529 tim 2697
530 tim 2789 Another generalization of Hamiltonian dynamics is Poisson
531     Dynamics\cite{Olver1986},
532 tim 2698 \begin{equation}
533     \dot x = J(x)\nabla _x H \label{introEquation:poissonHamiltonian}
534     \end{equation}
535     The most obvious change being that matrix $J$ now depends on $x$.
536    
537 tim 2702 \subsection{\label{introSection:exactFlow}Exact Flow}
538    
539 tim 2698 Let $x(t)$ be the exact solution of the ODE system,
540     \begin{equation}
541     \frac{{dx}}{{dt}} = f(x) \label{introEquation:ODE}
542     \end{equation}
543     The exact flow(solution) $\varphi_\tau$ is defined by
544     \[
545     x(t+\tau) =\varphi_\tau(x(t))
546     \]
547     where $\tau$ is a fixed time step and $\varphi$ is a map from phase
548 tim 2702 space to itself. The flow has the continuous group property,
549 tim 2698 \begin{equation}
550 tim 2702 \varphi _{\tau _1 } \circ \varphi _{\tau _2 } = \varphi _{\tau _1
551     + \tau _2 } .
552     \end{equation}
553     In particular,
554     \begin{equation}
555     \varphi _\tau \circ \varphi _{ - \tau } = I
556     \end{equation}
557     Therefore, the exact flow is self-adjoint,
558     \begin{equation}
559     \varphi _\tau = \varphi _{ - \tau }^{ - 1}.
560     \end{equation}
561     The exact flow can also be written in terms of the of an operator,
562     \begin{equation}
563     \varphi _\tau (x) = e^{\tau \sum\limits_i {f_i (x)\frac{\partial
564     }{{\partial x_i }}} } (x) \equiv \exp (\tau f)(x).
565     \label{introEquation:exponentialOperator}
566     \end{equation}
567    
568     In most cases, it is not easy to find the exact flow $\varphi_\tau$.
569 tim 2872 Instead, we use an approximate map, $\psi_\tau$, which is usually
570 tim 2702 called integrator. The order of an integrator $\psi_\tau$ is $p$, if
571     the Taylor series of $\psi_\tau$ agree to order $p$,
572     \begin{equation}
573 tim 2872 \psi_\tau(x) = x + \tau f(x) + O(\tau^{p+1})
574 tim 2698 \end{equation}
575    
576 tim 2702 \subsection{\label{introSection:geometricProperties}Geometric Properties}
577    
578 tim 2872 The hidden geometric properties\cite{Budd1999, Marsden1998} of an
579     ODE and its flow play important roles in numerical studies. Many of
580     them can be found in systems which occur naturally in applications.
581 tim 2702 Let $\varphi$ be the flow of Hamiltonian vector field, $\varphi$ is
582     a \emph{symplectic} flow if it satisfies,
583 tim 2698 \begin{equation}
584 tim 2703 {\varphi '}^T J \varphi ' = J.
585 tim 2698 \end{equation}
586     According to Liouville's theorem, the symplectic volume is invariant
587     under a Hamiltonian flow, which is the basis for classical
588 tim 2699 statistical mechanics. Furthermore, the flow of a Hamiltonian vector
589     field on a symplectic manifold can be shown to be a
590     symplectomorphism. As to the Poisson system,
591 tim 2698 \begin{equation}
592 tim 2703 {\varphi '}^T J \varphi ' = J \circ \varphi
593 tim 2698 \end{equation}
594 tim 2898 is the property that must be preserved by the integrator. It is
595     possible to construct a \emph{volume-preserving} flow for a source
596     free ODE ($ \nabla \cdot f = 0 $), if the flow satisfies $ \det
597     d\varphi = 1$. One can show easily that a symplectic flow will be
598     volume-preserving. Changing the variables $y = h(x)$ in an ODE
599 tim 2872 (Eq.~\ref{introEquation:ODE}) will result in a new system,
600 tim 2698 \[
601     \dot y = \tilde f(y) = ((dh \cdot f)h^{ - 1} )(y).
602     \]
603     The vector filed $f$ has reversing symmetry $h$ if $f = - \tilde f$.
604     In other words, the flow of this vector field is reversible if and
605 tim 2898 only if $ h \circ \varphi ^{ - 1} = \varphi \circ h $. A
606     \emph{first integral}, or conserved quantity of a general
607 tim 2705 differential function is a function $ G:R^{2d} \to R^d $ which is
608     constant for all solutions of the ODE $\frac{{dx}}{{dt}} = f(x)$ ,
609     \[
610     \frac{{dG(x(t))}}{{dt}} = 0.
611     \]
612     Using chain rule, one may obtain,
613     \[
614     \sum\limits_i {\frac{{dG}}{{dx_i }}} f_i (x) = f \bullet \nabla G,
615     \]
616     which is the condition for conserving \emph{first integral}. For a
617     canonical Hamiltonian system, the time evolution of an arbitrary
618     smooth function $G$ is given by,
619 tim 2789 \begin{eqnarray}
620     \frac{{dG(x(t))}}{{dt}} & = & [\nabla _x G(x(t))]^T \dot x(t) \\
621     & = & [\nabla _x G(x(t))]^T J\nabla _x H(x(t)). \\
622 tim 2705 \label{introEquation:firstIntegral1}
623 tim 2789 \end{eqnarray}
624 tim 2705 Using poisson bracket notion, Equation
625     \ref{introEquation:firstIntegral1} can be rewritten as
626     \[
627     \frac{d}{{dt}}G(x(t)) = \left\{ {G,H} \right\}(x(t)).
628     \]
629     Therefore, the sufficient condition for $G$ to be the \emph{first
630     integral} of a Hamiltonian system is
631     \[
632     \left\{ {G,H} \right\} = 0.
633     \]
634     As well known, the Hamiltonian (or energy) H of a Hamiltonian system
635     is a \emph{first integral}, which is due to the fact $\{ H,H\} =
636 tim 2898 0$. When designing any numerical methods, one should always try to
637 tim 2702 preserve the structural properties of the original ODE and its flow.
638    
639 tim 2699 \subsection{\label{introSection:constructionSymplectic}Construction of Symplectic Methods}
640     A lot of well established and very effective numerical methods have
641     been successful precisely because of their symplecticities even
642     though this fact was not recognized when they were first
643 tim 2872 constructed. The most famous example is the Verlet-leapfrog method
644 tim 2819 in molecular dynamics. In general, symplectic integrators can be
645 tim 2699 constructed using one of four different methods.
646     \begin{enumerate}
647     \item Generating functions
648     \item Variational methods
649     \item Runge-Kutta methods
650     \item Splitting methods
651     \end{enumerate}
652 tim 2698
653 tim 2789 Generating function\cite{Channell1990} tends to lead to methods
654     which are cumbersome and difficult to use. In dissipative systems,
655     variational methods can capture the decay of energy
656     accurately\cite{Kane2000}. Since their geometrically unstable nature
657     against non-Hamiltonian perturbations, ordinary implicit Runge-Kutta
658     methods are not suitable for Hamiltonian system. Recently, various
659     high-order explicit Runge-Kutta methods
660     \cite{Owren1992,Chen2003}have been developed to overcome this
661 tim 2703 instability. However, due to computational penalty involved in
662 tim 2819 implementing the Runge-Kutta methods, they have not attracted much
663     attention from the Molecular Dynamics community. Instead, splitting
664     methods have been widely accepted since they exploit natural
665     decompositions of the system\cite{Tuckerman1992, McLachlan1998}.
666 tim 2702
667 tim 2819 \subsubsection{\label{introSection:splittingMethod}\textbf{Splitting Methods}}
668 tim 2702
669     The main idea behind splitting methods is to decompose the discrete
670     $\varphi_h$ as a composition of simpler flows,
671 tim 2699 \begin{equation}
672     \varphi _h = \varphi _{h_1 } \circ \varphi _{h_2 } \ldots \circ
673     \varphi _{h_n }
674     \label{introEquation:FlowDecomposition}
675     \end{equation}
676     where each of the sub-flow is chosen such that each represent a
677 tim 2898 simpler integration of the system. Suppose that a Hamiltonian system
678     takes the form,
679 tim 2702 \[
680     H = H_1 + H_2.
681     \]
682     Here, $H_1$ and $H_2$ may represent different physical processes of
683     the system. For instance, they may relate to kinetic and potential
684     energy respectively, which is a natural decomposition of the
685     problem. If $H_1$ and $H_2$ can be integrated using exact flows
686     $\varphi_1(t)$ and $\varphi_2(t)$, respectively, a simple first
687 tim 2819 order expression is then given by the Lie-Trotter formula
688 tim 2699 \begin{equation}
689 tim 2702 \varphi _h = \varphi _{1,h} \circ \varphi _{2,h},
690     \label{introEquation:firstOrderSplitting}
691     \end{equation}
692     where $\varphi _h$ is the result of applying the corresponding
693     continuous $\varphi _i$ over a time $h$. By definition, as
694     $\varphi_i(t)$ is the exact solution of a Hamiltonian system, it
695     must follow that each operator $\varphi_i(t)$ is a symplectic map.
696     It is easy to show that any composition of symplectic flows yields a
697     symplectic map,
698     \begin{equation}
699 tim 2699 (\varphi '\phi ')^T J\varphi '\phi ' = \phi '^T \varphi '^T J\varphi
700 tim 2702 '\phi ' = \phi '^T J\phi ' = J,
701 tim 2699 \label{introEquation:SymplecticFlowComposition}
702     \end{equation}
703 tim 2702 where $\phi$ and $\psi$ both are symplectic maps. Thus operator
704     splitting in this context automatically generates a symplectic map.
705 tim 2699
706 tim 2702 The Lie-Trotter splitting(\ref{introEquation:firstOrderSplitting})
707     introduces local errors proportional to $h^2$, while Strang
708     splitting gives a second-order decomposition,
709     \begin{equation}
710     \varphi _h = \varphi _{1,h/2} \circ \varphi _{2,h} \circ \varphi
711 tim 2706 _{1,h/2} , \label{introEquation:secondOrderSplitting}
712 tim 2702 \end{equation}
713 tim 2819 which has a local error proportional to $h^3$. The Sprang
714     splitting's popularity in molecular simulation community attribute
715     to its symmetric property,
716 tim 2702 \begin{equation}
717     \varphi _h^{ - 1} = \varphi _{ - h}.
718 tim 2703 \label{introEquation:timeReversible}
719 tim 2882 \end{equation}
720 tim 2702
721 tim 2872 \subsubsection{\label{introSection:exampleSplittingMethod}\textbf{Examples of the Splitting Method}}
722 tim 2702 The classical equation for a system consisting of interacting
723     particles can be written in Hamiltonian form,
724     \[
725     H = T + V
726     \]
727     where $T$ is the kinetic energy and $V$ is the potential energy.
728 tim 2872 Setting $H_1 = T, H_2 = V$ and applying the Strang splitting, one
729 tim 2702 obtains the following:
730     \begin{align}
731     q(\Delta t) &= q(0) + \dot{q}(0)\Delta t +
732     \frac{F[q(0)]}{m}\frac{\Delta t^2}{2}, %
733     \label{introEquation:Lp10a} \\%
734     %
735     \dot{q}(\Delta t) &= \dot{q}(0) + \frac{\Delta t}{2m}
736     \biggl [F[q(0)] + F[q(\Delta t)] \biggr]. %
737     \label{introEquation:Lp10b}
738     \end{align}
739     where $F(t)$ is the force at time $t$. This integration scheme is
740     known as \emph{velocity verlet} which is
741     symplectic(\ref{introEquation:SymplecticFlowComposition}),
742     time-reversible(\ref{introEquation:timeReversible}) and
743     volume-preserving (\ref{introEquation:volumePreserving}). These
744     geometric properties attribute to its long-time stability and its
745     popularity in the community. However, the most commonly used
746     velocity verlet integration scheme is written as below,
747     \begin{align}
748     \dot{q}\biggl (\frac{\Delta t}{2}\biggr ) &=
749     \dot{q}(0) + \frac{\Delta t}{2m}\, F[q(0)], \label{introEquation:Lp9a}\\%
750     %
751     q(\Delta t) &= q(0) + \Delta t\, \dot{q}\biggl (\frac{\Delta t}{2}\biggr ),%
752     \label{introEquation:Lp9b}\\%
753     %
754     \dot{q}(\Delta t) &= \dot{q}\biggl (\frac{\Delta t}{2}\biggr ) +
755 tim 2872 \frac{\Delta t}{2m}\, F[q(t)]. \label{introEquation:Lp9c}
756 tim 2702 \end{align}
757     From the preceding splitting, one can see that the integration of
758     the equations of motion would follow:
759     \begin{enumerate}
760     \item calculate the velocities at the half step, $\frac{\Delta t}{2}$, from the forces calculated at the initial position.
761    
762     \item Use the half step velocities to move positions one whole step, $\Delta t$.
763    
764 tim 2872 \item Evaluate the forces at the new positions, $\mathbf{q}(\Delta t)$, and use the new forces to complete the velocity move.
765 tim 2702
766     \item Repeat from step 1 with the new position, velocities, and forces assuming the roles of the initial values.
767     \end{enumerate}
768    
769 tim 2872 By simply switching the order of the propagators in the splitting
770     and composing a new integrator, the \emph{position verlet}
771     integrator, can be generated,
772 tim 2702 \begin{align}
773     \dot q(\Delta t) &= \dot q(0) + \Delta tF(q(0))\left[ {q(0) +
774     \frac{{\Delta t}}{{2m}}\dot q(0)} \right], %
775     \label{introEquation:positionVerlet1} \\%
776     %
777 tim 2703 q(\Delta t) &= q(0) + \frac{{\Delta t}}{2}\left[ {\dot q(0) + \dot
778 tim 2702 q(\Delta t)} \right]. %
779 tim 2719 \label{introEquation:positionVerlet2}
780 tim 2702 \end{align}
781    
782 tim 2819 \subsubsection{\label{introSection:errorAnalysis}\textbf{Error Analysis and Higher Order Methods}}
783 tim 2702
784 tim 2872 The Baker-Campbell-Hausdorff formula can be used to determine the
785     local error of splitting method in terms of the commutator of the
786 tim 2702 operators(\ref{introEquation:exponentialOperator}) associated with
787 tim 2872 the sub-flow. For operators $hX$ and $hY$ which are associated with
788 tim 2726 $\varphi_1(t)$ and $\varphi_2(t)$ respectively , we have
789 tim 2702 \begin{equation}
790     \exp (hX + hY) = \exp (hZ)
791     \end{equation}
792     where
793     \begin{equation}
794     hZ = hX + hY + \frac{{h^2 }}{2}[X,Y] + \frac{{h^3 }}{2}\left(
795     {[X,[X,Y]] + [Y,[Y,X]]} \right) + \ldots .
796     \end{equation}
797     Here, $[X,Y]$ is the commutators of operator $X$ and $Y$ given by
798     \[
799     [X,Y] = XY - YX .
800     \]
801 tim 2872 Applying the Baker-Campbell-Hausdorff formula\cite{Varadarajan1974}
802     to the Sprang splitting, we can obtain
803 tim 2779 \begin{eqnarray*}
804 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 \\
805     & & \mbox{} + h^2 [X,X]/8 + h^2 [Y,Y]/8 \\
806 tim 2779 & & \mbox{} + h^3 [Y,[Y,X]]/12 - h^3[X,[X,Y]]/24 + \ldots )
807     \end{eqnarray*}
808 tim 2872 Since \[ [X,Y] + [Y,X] = 0\] and \[ [X,X] = 0,\] the dominant local
809 tim 2702 error of Spring splitting is proportional to $h^3$. The same
810 tim 2872 procedure can be applied to a general splitting, of the form
811 tim 2702 \begin{equation}
812     \varphi _{b_m h}^2 \circ \varphi _{a_m h}^1 \circ \varphi _{b_{m -
813     1} h}^2 \circ \ldots \circ \varphi _{a_1 h}^1 .
814     \end{equation}
815 tim 2872 A careful choice of coefficient $a_1 \ldots b_m$ will lead to higher
816     order methods. Yoshida proposed an elegant way to compose higher
817 tim 2789 order methods based on symmetric splitting\cite{Yoshida1990}. Given
818     a symmetric second order base method $ \varphi _h^{(2)} $, a
819     fourth-order symmetric method can be constructed by composing,
820 tim 2702 \[
821     \varphi _h^{(4)} = \varphi _{\alpha h}^{(2)} \circ \varphi _{\beta
822     h}^{(2)} \circ \varphi _{\alpha h}^{(2)}
823     \]
824     where $ \alpha = - \frac{{2^{1/3} }}{{2 - 2^{1/3} }}$ and $ \beta
825     = \frac{{2^{1/3} }}{{2 - 2^{1/3} }}$. Moreover, a symmetric
826     integrator $ \varphi _h^{(2n + 2)}$ can be composed by
827     \begin{equation}
828     \varphi _h^{(2n + 2)} = \varphi _{\alpha h}^{(2n)} \circ \varphi
829 tim 2872 _{\beta h}^{(2n)} \circ \varphi _{\alpha h}^{(2n)},
830 tim 2702 \end{equation}
831 tim 2872 if the weights are chosen as
832 tim 2702 \[
833     \alpha = - \frac{{2^{1/(2n + 1)} }}{{2 - 2^{1/(2n + 1)} }},\beta =
834     \frac{{2^{1/(2n + 1)} }}{{2 - 2^{1/(2n + 1)} }} .
835     \]
836    
837 tim 2694 \section{\label{introSection:molecularDynamics}Molecular Dynamics}
838    
839 tim 2720 As one of the principal tools of molecular modeling, Molecular
840     dynamics has proven to be a powerful tool for studying the functions
841     of biological systems, providing structural, thermodynamic and
842     dynamical information. The basic idea of molecular dynamics is that
843     macroscopic properties are related to microscopic behavior and
844     microscopic behavior can be calculated from the trajectories in
845     simulations. For instance, instantaneous temperature of an
846     Hamiltonian system of $N$ particle can be measured by
847     \[
848 tim 2725 T = \sum\limits_{i = 1}^N {\frac{{m_i v_i^2 }}{{fk_B }}}
849 tim 2720 \]
850     where $m_i$ and $v_i$ are the mass and velocity of $i$th particle
851     respectively, $f$ is the number of degrees of freedom, and $k_B$ is
852     the boltzman constant.
853 tim 2694
854 tim 2720 A typical molecular dynamics run consists of three essential steps:
855     \begin{enumerate}
856     \item Initialization
857     \begin{enumerate}
858     \item Preliminary preparation
859     \item Minimization
860     \item Heating
861     \item Equilibration
862     \end{enumerate}
863     \item Production
864     \item Analysis
865     \end{enumerate}
866     These three individual steps will be covered in the following
867     sections. Sec.~\ref{introSec:initialSystemSettings} deals with the
868 tim 2801 initialization of a simulation. Sec.~\ref{introSection:production}
869 tim 2872 will discusse issues in production run.
870 tim 2801 Sec.~\ref{introSection:Analysis} provides the theoretical tools for
871     trajectory analysis.
872 tim 2719
873 tim 2720 \subsection{\label{introSec:initialSystemSettings}Initialization}
874 tim 2719
875 tim 2819 \subsubsection{\textbf{Preliminary preparation}}
876 tim 2719
877 tim 2720 When selecting the starting structure of a molecule for molecular
878     simulation, one may retrieve its Cartesian coordinates from public
879     databases, such as RCSB Protein Data Bank \textit{etc}. Although
880     thousands of crystal structures of molecules are discovered every
881     year, many more remain unknown due to the difficulties of
882 tim 2872 purification and crystallization. Even for molecules with known
883     structure, some important information is missing. For example, a
884 tim 2720 missing hydrogen atom which acts as donor in hydrogen bonding must
885     be added. Moreover, in order to include electrostatic interaction,
886     one may need to specify the partial charges for individual atoms.
887     Under some circumstances, we may even need to prepare the system in
888 tim 2872 a special configuration. For instance, when studying transport
889     phenomenon in membrane systems, we may prepare the lipids in a
890     bilayer structure instead of placing lipids randomly in solvent,
891     since we are not interested in the slow self-aggregation process.
892 tim 2694
893 tim 2819 \subsubsection{\textbf{Minimization}}
894 tim 2705
895 tim 2720 It is quite possible that some of molecules in the system from
896 tim 2872 preliminary preparation may be overlapping with each other. This
897     close proximity leads to high initial potential energy which
898     consequently jeopardizes any molecular dynamics simulations. To
899     remove these steric overlaps, one typically performs energy
900     minimization to find a more reasonable conformation. Several energy
901     minimization methods have been developed to exploit the energy
902     surface and to locate the local minimum. While converging slowly
903     near the minimum, steepest descent method is extremely robust when
904     systems are strongly anharmonic. Thus, it is often used to refine
905     structure from crystallographic data. Relied on the gradient or
906     hessian, advanced methods like Newton-Raphson converge rapidly to a
907     local minimum, but become unstable if the energy surface is far from
908     quadratic. Another factor that must be taken into account, when
909 tim 2720 choosing energy minimization method, is the size of the system.
910     Steepest descent and conjugate gradient can deal with models of any
911 tim 2872 size. Because of the limits on computer memory to store the hessian
912     matrix and the computing power needed to diagonalized these
913     matrices, most Newton-Raphson methods can not be used with very
914     large systems.
915 tim 2694
916 tim 2819 \subsubsection{\textbf{Heating}}
917 tim 2720
918     Typically, Heating is performed by assigning random velocities
919 tim 2872 according to a Maxwell-Boltzman distribution for a desired
920     temperature. Beginning at a lower temperature and gradually
921     increasing the temperature by assigning larger random velocities, we
922     end up with setting the temperature of the system to a final
923     temperature at which the simulation will be conducted. In heating
924     phase, we should also keep the system from drifting or rotating as a
925     whole. To do this, the net linear momentum and angular momentum of
926     the system is shifted to zero after each resampling from the Maxwell
927     -Boltzman distribution.
928 tim 2720
929 tim 2819 \subsubsection{\textbf{Equilibration}}
930 tim 2720
931     The purpose of equilibration is to allow the system to evolve
932     spontaneously for a period of time and reach equilibrium. The
933     procedure is continued until various statistical properties, such as
934     temperature, pressure, energy, volume and other structural
935     properties \textit{etc}, become independent of time. Strictly
936     speaking, minimization and heating are not necessary, provided the
937     equilibration process is long enough. However, these steps can serve
938     as a means to arrive at an equilibrated structure in an effective
939     way.
940    
941     \subsection{\label{introSection:production}Production}
942    
943 tim 2872 The production run is the most important step of the simulation, in
944 tim 2725 which the equilibrated structure is used as a starting point and the
945     motions of the molecules are collected for later analysis. In order
946     to capture the macroscopic properties of the system, the molecular
947 tim 2872 dynamics simulation must be performed by sampling correctly and
948     efficiently from the relevant thermodynamic ensemble.
949 tim 2720
950 tim 2725 The most expensive part of a molecular dynamics simulation is the
951     calculation of non-bonded forces, such as van der Waals force and
952     Coulombic forces \textit{etc}. For a system of $N$ particles, the
953     complexity of the algorithm for pair-wise interactions is $O(N^2 )$,
954     which making large simulations prohibitive in the absence of any
955 tim 2872 algorithmic tricks.
956 tim 2720
957 tim 2872 A natural approach to avoid system size issues is to represent the
958 tim 2725 bulk behavior by a finite number of the particles. However, this
959 tim 2872 approach will suffer from the surface effect at the edges of the
960     simulation. To offset this, \textit{Periodic boundary conditions}
961     (see Fig.~\ref{introFig:pbc}) is developed to simulate bulk
962     properties with a relatively small number of particles. In this
963     method, the simulation box is replicated throughout space to form an
964     infinite lattice. During the simulation, when a particle moves in
965     the primary cell, its image in other cells move in exactly the same
966     direction with exactly the same orientation. Thus, as a particle
967     leaves the primary cell, one of its images will enter through the
968     opposite face.
969 tim 2789 \begin{figure}
970     \centering
971     \includegraphics[width=\linewidth]{pbc.eps}
972     \caption[An illustration of periodic boundary conditions]{A 2-D
973     illustration of periodic boundary conditions. As one particle leaves
974     the left of the simulation box, an image of it enters the right.}
975     \label{introFig:pbc}
976     \end{figure}
977 tim 2725
978     %cutoff and minimum image convention
979     Another important technique to improve the efficiency of force
980 tim 2872 evaluation is to apply spherical cutoff where particles farther than
981     a predetermined distance are not included in the calculation
982 tim 2725 \cite{Frenkel1996}. The use of a cutoff radius will cause a
983 tim 2730 discontinuity in the potential energy curve. Fortunately, one can
984 tim 2872 shift simple radial potential to ensure the potential curve go
985     smoothly to zero at the cutoff radius. The cutoff strategy works
986     well for Lennard-Jones interaction because of its short range
987     nature. However, simply truncating the electrostatic interaction
988     with the use of cutoffs has been shown to lead to severe artifacts
989     in simulations. The Ewald summation, in which the slowly decaying
990     Coulomb potential is transformed into direct and reciprocal sums
991     with rapid and absolute convergence, has proved to minimize the
992     periodicity artifacts in liquid simulations. Taking the advantages
993     of the fast Fourier transform (FFT) for calculating discrete Fourier
994     transforms, the particle mesh-based
995 tim 2789 methods\cite{Hockney1981,Shimada1993, Luty1994} are accelerated from
996 tim 2872 $O(N^{3/2})$ to $O(N logN)$. An alternative approach is the
997     \emph{fast multipole method}\cite{Greengard1987, Greengard1994},
998     which treats Coulombic interactions exactly at short range, and
999     approximate the potential at long range through multipolar
1000     expansion. In spite of their wide acceptance at the molecular
1001     simulation community, these two methods are difficult to implement
1002     correctly and efficiently. Instead, we use a damped and
1003     charge-neutralized Coulomb potential method developed by Wolf and
1004     his coworkers\cite{Wolf1999}. The shifted Coulomb potential for
1005     particle $i$ and particle $j$ at distance $r_{rj}$ is given by:
1006 tim 2725 \begin{equation}
1007     V(r_{ij})= \frac{q_i q_j \textrm{erfc}(\alpha
1008     r_{ij})}{r_{ij}}-\lim_{r_{ij}\rightarrow
1009     R_\textrm{c}}\left\{\frac{q_iq_j \textrm{erfc}(\alpha
1010     r_{ij})}{r_{ij}}\right\}. \label{introEquation:shiftedCoulomb}
1011     \end{equation}
1012     where $\alpha$ is the convergence parameter. Due to the lack of
1013     inherent periodicity and rapid convergence,this method is extremely
1014     efficient and easy to implement.
1015 tim 2789 \begin{figure}
1016     \centering
1017     \includegraphics[width=\linewidth]{shifted_coulomb.eps}
1018     \caption[An illustration of shifted Coulomb potential]{An
1019     illustration of shifted Coulomb potential.}
1020     \label{introFigure:shiftedCoulomb}
1021     \end{figure}
1022 tim 2725
1023     %multiple time step
1024    
1025 tim 2720 \subsection{\label{introSection:Analysis} Analysis}
1026    
1027 tim 2872 Recently, advanced visualization technique have become applied to
1028 tim 2721 monitor the motions of molecules. Although the dynamics of the
1029     system can be described qualitatively from animation, quantitative
1030 tim 2872 trajectory analysis are more useful. According to the principles of
1031     Statistical Mechanics, Sec.~\ref{introSection:statisticalMechanics},
1032     one can compute thermodynamic properties, analyze fluctuations of
1033     structural parameters, and investigate time-dependent processes of
1034     the molecule from the trajectories.
1035 tim 2721
1036 tim 2872 \subsubsection{\label{introSection:thermodynamicsProperties}\textbf{Thermodynamic Properties}}
1037 tim 2721
1038 tim 2872 Thermodynamic properties, which can be expressed in terms of some
1039 tim 2725 function of the coordinates and momenta of all particles in the
1040     system, can be directly computed from molecular dynamics. The usual
1041     way to measure the pressure is based on virial theorem of Clausius
1042     which states that the virial is equal to $-3Nk_BT$. For a system
1043     with forces between particles, the total virial, $W$, contains the
1044     contribution from external pressure and interaction between the
1045     particles:
1046     \[
1047     W = - 3PV + \left\langle {\sum\limits_{i < j} {r{}_{ij} \cdot
1048     f_{ij} } } \right\rangle
1049     \]
1050     where $f_{ij}$ is the force between particle $i$ and $j$ at a
1051     distance $r_{ij}$. Thus, the expression for the pressure is given
1052     by:
1053     \begin{equation}
1054     P = \frac{{Nk_B T}}{V} - \frac{1}{{3V}}\left\langle {\sum\limits_{i
1055     < j} {r{}_{ij} \cdot f_{ij} } } \right\rangle
1056     \end{equation}
1057    
1058 tim 2819 \subsubsection{\label{introSection:structuralProperties}\textbf{Structural Properties}}
1059 tim 2721
1060     Structural Properties of a simple fluid can be described by a set of
1061 tim 2872 distribution functions. Among these functions,the \emph{pair
1062 tim 2721 distribution function}, also known as \emph{radial distribution
1063 tim 2872 function}, is of most fundamental importance to liquid theory.
1064     Experimentally, pair distribution function can be gathered by
1065     Fourier transforming raw data from a series of neutron diffraction
1066     experiments and integrating over the surface factor
1067     \cite{Powles1973}. The experimental results can serve as a criterion
1068     to justify the correctness of a liquid model. Moreover, various
1069     equilibrium thermodynamic and structural properties can also be
1070     expressed in terms of radial distribution function \cite{Allen1987}.
1071 tim 2721
1072 tim 2872 The pair distribution functions $g(r)$ gives the probability that a
1073 tim 2721 particle $i$ will be located at a distance $r$ from a another
1074     particle $j$ in the system
1075     \[
1076     g(r) = \frac{V}{{N^2 }}\left\langle {\sum\limits_i {\sum\limits_{j
1077 tim 2874 \ne i} {\delta (r - r_{ij} )} } } \right\rangle = \frac{\rho
1078 tim 2872 (r)}{\rho}.
1079 tim 2721 \]
1080     Note that the delta function can be replaced by a histogram in
1081 tim 2881 computer simulation. Peaks in $g(r)$ represent solvent shells, and
1082     the height of these peaks gradually decreases to 1 as the liquid of
1083     large distance approaches the bulk density.
1084 tim 2721
1085    
1086 tim 2819 \subsubsection{\label{introSection:timeDependentProperties}\textbf{Time-dependent
1087     Properties}}
1088 tim 2721
1089     Time-dependent properties are usually calculated using \emph{time
1090 tim 2872 correlation functions}, which correlate random variables $A$ and $B$
1091     at two different times,
1092 tim 2721 \begin{equation}
1093     C_{AB} (t) = \left\langle {A(t)B(0)} \right\rangle.
1094     \label{introEquation:timeCorrelationFunction}
1095     \end{equation}
1096     If $A$ and $B$ refer to same variable, this kind of correlation
1097 tim 2872 function is called an \emph{autocorrelation function}. One example
1098     of an auto correlation function is the velocity auto-correlation
1099     function which is directly related to transport properties of
1100     molecular liquids:
1101 tim 2725 \[
1102     D = \frac{1}{3}\int\limits_0^\infty {\left\langle {v(t) \cdot v(0)}
1103     \right\rangle } dt
1104     \]
1105 tim 2872 where $D$ is diffusion constant. Unlike the velocity autocorrelation
1106     function, which is averaging over time origins and over all the
1107     atoms, the dipole autocorrelation functions are calculated for the
1108     entire system. The dipole autocorrelation function is given by:
1109 tim 2725 \[
1110     c_{dipole} = \left\langle {u_{tot} (t) \cdot u_{tot} (t)}
1111     \right\rangle
1112     \]
1113     Here $u_{tot}$ is the net dipole of the entire system and is given
1114     by
1115     \[
1116     u_{tot} (t) = \sum\limits_i {u_i (t)}
1117     \]
1118     In principle, many time correlation functions can be related with
1119     Fourier transforms of the infrared, Raman, and inelastic neutron
1120     scattering spectra of molecular liquids. In practice, one can
1121     extract the IR spectrum from the intensity of dipole fluctuation at
1122     each frequency using the following relationship:
1123     \[
1124     \hat c_{dipole} (v) = \int_{ - \infty }^\infty {c_{dipole} (t)e^{ -
1125     i2\pi vt} dt}
1126     \]
1127 tim 2721
1128 tim 2693 \section{\label{introSection:rigidBody}Dynamics of Rigid Bodies}
1129 tim 2692
1130 tim 2705 Rigid bodies are frequently involved in the modeling of different
1131     areas, from engineering, physics, to chemistry. For example,
1132     missiles and vehicle are usually modeled by rigid bodies. The
1133     movement of the objects in 3D gaming engine or other physics
1134 tim 2872 simulator is governed by rigid body dynamics. In molecular
1135     simulations, rigid bodies are used to simplify protein-protein
1136     docking studies\cite{Gray2003}.
1137 tim 2694
1138 tim 2705 It is very important to develop stable and efficient methods to
1139 tim 2872 integrate the equations of motion for orientational degrees of
1140     freedom. Euler angles are the natural choice to describe the
1141     rotational degrees of freedom. However, due to $\frac {1}{sin
1142     \theta}$ singularities, the numerical integration of corresponding
1143     equations of motion is very inefficient and inaccurate. Although an
1144     alternative integrator using multiple sets of Euler angles can
1145     overcome this difficulty\cite{Barojas1973}, the computational
1146     penalty and the loss of angular momentum conservation still remain.
1147     A singularity-free representation utilizing quaternions was
1148     developed by Evans in 1977\cite{Evans1977}. Unfortunately, this
1149     approach uses a nonseparable Hamiltonian resulting from the
1150     quaternion representation, which prevents the symplectic algorithm
1151     to be utilized. Another different approach is to apply holonomic
1152     constraints to the atoms belonging to the rigid body. Each atom
1153     moves independently under the normal forces deriving from potential
1154     energy and constraint forces which are used to guarantee the
1155     rigidness. However, due to their iterative nature, the SHAKE and
1156     Rattle algorithms also converge very slowly when the number of
1157     constraints increases\cite{Ryckaert1977, Andersen1983}.
1158 tim 2694
1159 tim 2872 A break-through in geometric literature suggests that, in order to
1160 tim 2705 develop a long-term integration scheme, one should preserve the
1161 tim 2872 symplectic structure of the flow. By introducing a conjugate
1162     momentum to the rotation matrix $Q$ and re-formulating Hamiltonian's
1163     equation, a symplectic integrator, RSHAKE\cite{Kol1997}, was
1164     proposed to evolve the Hamiltonian system in a constraint manifold
1165     by iteratively satisfying the orthogonality constraint $Q^T Q = 1$.
1166     An alternative method using the quaternion representation was
1167     developed by Omelyan\cite{Omelyan1998}. However, both of these
1168     methods are iterative and inefficient. In this section, we descibe a
1169 tim 2789 symplectic Lie-Poisson integrator for rigid body developed by
1170     Dullweber and his coworkers\cite{Dullweber1997} in depth.
1171 tim 2705
1172 tim 2872 \subsection{\label{introSection:constrainedHamiltonianRB}Constrained Hamiltonian for Rigid Bodies}
1173     The motion of a rigid body is Hamiltonian with the Hamiltonian
1174 tim 2713 function
1175 tim 2706 \begin{equation}
1176     H = \frac{1}{2}(p^T m^{ - 1} p) + \frac{1}{2}tr(PJ^{ - 1} P) +
1177     V(q,Q) + \frac{1}{2}tr[(QQ^T - 1)\Lambda ].
1178     \label{introEquation:RBHamiltonian}
1179     \end{equation}
1180     Here, $q$ and $Q$ are the position and rotation matrix for the
1181     rigid-body, $p$ and $P$ are conjugate momenta to $q$ and $Q$ , and
1182     $J$, a diagonal matrix, is defined by
1183     \[
1184     I_{ii}^{ - 1} = \frac{1}{2}\sum\limits_{i \ne j} {J_{jj}^{ - 1} }
1185     \]
1186     where $I_{ii}$ is the diagonal element of the inertia tensor. This
1187 tim 2872 constrained Hamiltonian equation is subjected to a holonomic
1188     constraint,
1189 tim 2706 \begin{equation}
1190 tim 2726 Q^T Q = 1, \label{introEquation:orthogonalConstraint}
1191 tim 2706 \end{equation}
1192 tim 2872 which is used to ensure rotation matrix's unitarity. Differentiating
1193     \ref{introEquation:orthogonalConstraint} and using Equation
1194     \ref{introEquation:RBMotionMomentum}, one may obtain,
1195 tim 2706 \begin{equation}
1196 tim 2707 Q^T PJ^{ - 1} + J^{ - 1} P^T Q = 0 . \\
1197 tim 2706 \label{introEquation:RBFirstOrderConstraint}
1198     \end{equation}
1199    
1200     Using Equation (\ref{introEquation:motionHamiltonianCoordinate},
1201     \ref{introEquation:motionHamiltonianMomentum}), one can write down
1202     the equations of motion,
1203    
1204 tim 2796 \begin{eqnarray}
1205     \frac{{dq}}{{dt}} & = & \frac{p}{m} \label{introEquation:RBMotionPosition}\\
1206     \frac{{dp}}{{dt}} & = & - \nabla _q V(q,Q) \label{introEquation:RBMotionMomentum}\\
1207     \frac{{dQ}}{{dt}} & = & PJ^{ - 1} \label{introEquation:RBMotionRotation}\\
1208     \frac{{dP}}{{dt}} & = & - \nabla _Q V(q,Q) - 2Q\Lambda . \label{introEquation:RBMotionP}
1209     \end{eqnarray}
1210    
1211 tim 2707 In general, there are two ways to satisfy the holonomic constraints.
1212 tim 2872 We can use a constraint force provided by a Lagrange multiplier on
1213     the normal manifold to keep the motion on constraint space. Or we
1214     can simply evolve the system on the constraint manifold. These two
1215     methods have been proved to be equivalent. The holonomic constraint
1216     and equations of motions define a constraint manifold for rigid
1217     bodies
1218 tim 2707 \[
1219     M = \left\{ {(Q,P):Q^T Q = 1,Q^T PJ^{ - 1} + J^{ - 1} P^T Q = 0}
1220     \right\}.
1221     \]
1222 tim 2706
1223 tim 2707 Unfortunately, this constraint manifold is not the cotangent bundle
1224 tim 2888 $T^* SO(3)$ which can be consider as a symplectic manifold on Lie
1225     rotation group $SO(3)$. However, it turns out that under symplectic
1226 tim 2707 transformation, the cotangent space and the phase space are
1227 tim 2872 diffeomorphic. By introducing
1228 tim 2706 \[
1229 tim 2707 \tilde Q = Q,\tilde P = \frac{1}{2}\left( {P - QP^T Q} \right),
1230 tim 2706 \]
1231 tim 2707 the mechanical system subject to a holonomic constraint manifold $M$
1232     can be re-formulated as a Hamiltonian system on the cotangent space
1233     \[
1234     T^* SO(3) = \left\{ {(\tilde Q,\tilde P):\tilde Q^T \tilde Q =
1235     1,\tilde Q^T \tilde PJ^{ - 1} + J^{ - 1} P^T \tilde Q = 0} \right\}
1236     \]
1237 tim 2706
1238 tim 2707 For a body fixed vector $X_i$ with respect to the center of mass of
1239     the rigid body, its corresponding lab fixed vector $X_0^{lab}$ is
1240     given as
1241     \begin{equation}
1242     X_i^{lab} = Q X_i + q.
1243     \end{equation}
1244     Therefore, potential energy $V(q,Q)$ is defined by
1245     \[
1246     V(q,Q) = V(Q X_0 + q).
1247     \]
1248 tim 2713 Hence, the force and torque are given by
1249 tim 2707 \[
1250 tim 2713 \nabla _q V(q,Q) = F(q,Q) = \sum\limits_i {F_i (q,Q)},
1251 tim 2707 \]
1252 tim 2713 and
1253 tim 2707 \[
1254     \nabla _Q V(q,Q) = F(q,Q)X_i^t
1255     \]
1256 tim 2713 respectively.
1257 tim 2695
1258 tim 2707 As a common choice to describe the rotation dynamics of the rigid
1259 tim 2872 body, the angular momentum on the body fixed frame $\Pi = Q^t P$ is
1260     introduced to rewrite the equations of motion,
1261 tim 2707 \begin{equation}
1262     \begin{array}{l}
1263 tim 2888 \dot \Pi = J^{ - 1} \Pi ^T \Pi + Q^T \sum\limits_i {F_i (q,Q)X_i^T } - \Lambda \\
1264     \dot Q = Q\Pi {\rm{ }}J^{ - 1} \\
1265 tim 2707 \end{array}
1266     \label{introEqaution:RBMotionPI}
1267     \end{equation}
1268     , as well as holonomic constraints,
1269     \[
1270     \begin{array}{l}
1271     \Pi J^{ - 1} + J^{ - 1} \Pi ^t = 0 \\
1272     Q^T Q = 1 \\
1273     \end{array}
1274     \]
1275 tim 2692
1276 tim 2707 For a vector $v(v_1 ,v_2 ,v_3 ) \in R^3$ and a matrix $\hat v \in
1277     so(3)^ \star$, the hat-map isomorphism,
1278     \begin{equation}
1279     v(v_1 ,v_2 ,v_3 ) \Leftrightarrow \hat v = \left(
1280     {\begin{array}{*{20}c}
1281     0 & { - v_3 } & {v_2 } \\
1282     {v_3 } & 0 & { - v_1 } \\
1283     { - v_2 } & {v_1 } & 0 \\
1284     \end{array}} \right),
1285     \label{introEquation:hatmapIsomorphism}
1286     \end{equation}
1287     will let us associate the matrix products with traditional vector
1288     operations
1289     \[
1290     \hat vu = v \times u
1291     \]
1292     Using \ref{introEqaution:RBMotionPI}, one can construct a skew
1293     matrix,
1294 tim 2888
1295 tim 2890 \begin{eqnarray*}
1296 tim 2888 (\dot \Pi - \dot \Pi ^T ){\rm{ }} = {\rm{ }}(\Pi - \Pi ^T ){\rm{
1297     }}(J^{ - 1} \Pi + \Pi J^{ - 1} ) + \sum\limits_i {[Q^T F_i
1298     (r,Q)X_i^T - X_i F_i (r,Q)^T Q]} - (\Lambda - \Lambda ^T ).
1299     \label{introEquation:skewMatrixPI}
1300     \end{eqnarray*}
1301    
1302 tim 2707 Since $\Lambda$ is symmetric, the last term of Equation
1303 tim 2713 \ref{introEquation:skewMatrixPI} is zero, which implies the Lagrange
1304     multiplier $\Lambda$ is absent from the equations of motion. This
1305 tim 2872 unique property eliminates the requirement of iterations which can
1306 tim 2789 not be avoided in other methods\cite{Kol1997, Omelyan1998}.
1307 tim 2707
1308 tim 2872 Applying the hat-map isomorphism, we obtain the equation of motion
1309     for angular momentum on body frame
1310 tim 2713 \begin{equation}
1311     \dot \pi = \pi \times I^{ - 1} \pi + \sum\limits_i {\left( {Q^T
1312     F_i (r,Q)} \right) \times X_i }.
1313     \label{introEquation:bodyAngularMotion}
1314     \end{equation}
1315 tim 2707 In the same manner, the equation of motion for rotation matrix is
1316     given by
1317     \[
1318 tim 2713 \dot Q = Qskew(I^{ - 1} \pi )
1319 tim 2707 \]
1320    
1321 tim 2713 \subsection{\label{introSection:SymplecticFreeRB}Symplectic
1322     Lie-Poisson Integrator for Free Rigid Body}
1323 tim 2707
1324 tim 2872 If there are no external forces exerted on the rigid body, the only
1325     contribution to the rotational motion is from the kinetic energy
1326     (the first term of \ref{introEquation:bodyAngularMotion}). The free
1327     rigid body is an example of a Lie-Poisson system with Hamiltonian
1328     function
1329 tim 2713 \begin{equation}
1330     T^r (\pi ) = T_1 ^r (\pi _1 ) + T_2^r (\pi _2 ) + T_3^r (\pi _3 )
1331     \label{introEquation:rotationalKineticRB}
1332     \end{equation}
1333     where $T_i^r (\pi _i ) = \frac{{\pi _i ^2 }}{{2I_i }}$ and
1334     Lie-Poisson structure matrix,
1335     \begin{equation}
1336     J(\pi ) = \left( {\begin{array}{*{20}c}
1337     0 & {\pi _3 } & { - \pi _2 } \\
1338     { - \pi _3 } & 0 & {\pi _1 } \\
1339     {\pi _2 } & { - \pi _1 } & 0 \\
1340     \end{array}} \right)
1341     \end{equation}
1342     Thus, the dynamics of free rigid body is governed by
1343     \begin{equation}
1344     \frac{d}{{dt}}\pi = J(\pi )\nabla _\pi T^r (\pi )
1345     \end{equation}
1346 tim 2707
1347 tim 2713 One may notice that each $T_i^r$ in Equation
1348     \ref{introEquation:rotationalKineticRB} can be solved exactly. For
1349     instance, the equations of motion due to $T_1^r$ are given by
1350     \begin{equation}
1351     \frac{d}{{dt}}\pi = R_1 \pi ,\frac{d}{{dt}}Q = QR_1
1352     \label{introEqaution:RBMotionSingleTerm}
1353     \end{equation}
1354     where
1355     \[ R_1 = \left( {\begin{array}{*{20}c}
1356     0 & 0 & 0 \\
1357     0 & 0 & {\pi _1 } \\
1358     0 & { - \pi _1 } & 0 \\
1359     \end{array}} \right).
1360     \]
1361     The solutions of Equation \ref{introEqaution:RBMotionSingleTerm} is
1362 tim 2707 \[
1363 tim 2713 \pi (\Delta t) = e^{\Delta tR_1 } \pi (0),Q(\Delta t) =
1364     Q(0)e^{\Delta tR_1 }
1365 tim 2707 \]
1366 tim 2713 with
1367 tim 2707 \[
1368 tim 2713 e^{\Delta tR_1 } = \left( {\begin{array}{*{20}c}
1369     0 & 0 & 0 \\
1370     0 & {\cos \theta _1 } & {\sin \theta _1 } \\
1371     0 & { - \sin \theta _1 } & {\cos \theta _1 } \\
1372     \end{array}} \right),\theta _1 = \frac{{\pi _1 }}{{I_1 }}\Delta t.
1373 tim 2707 \]
1374 tim 2719 To reduce the cost of computing expensive functions in $e^{\Delta
1375 tim 2872 tR_1 }$, we can use Cayley transformation to obtain a single-aixs
1376     propagator,
1377 tim 2713 \[
1378     e^{\Delta tR_1 } \approx (1 - \Delta tR_1 )^{ - 1} (1 + \Delta tR_1
1379     )
1380     \]
1381 tim 2720 The flow maps for $T_2^r$ and $T_3^r$ can be found in the same
1382 tim 2872 manner. In order to construct a second-order symplectic method, we
1383     split the angular kinetic Hamiltonian function can into five terms
1384 tim 2707 \[
1385 tim 2713 T^r (\pi ) = \frac{1}{2}T_1 ^r (\pi _1 ) + \frac{1}{2}T_2^r (\pi _2
1386     ) + T_3^r (\pi _3 ) + \frac{1}{2}T_2^r (\pi _2 ) + \frac{1}{2}T_1 ^r
1387 tim 2872 (\pi _1 ).
1388     \]
1389     By concatenating the propagators corresponding to these five terms,
1390     we can obtain an symplectic integrator,
1391 tim 2713 \[
1392     \varphi _{\Delta t,T^r } = \varphi _{\Delta t/2,\pi _1 } \circ
1393 tim 2707 \varphi _{\Delta t/2,\pi _2 } \circ \varphi _{\Delta t,\pi _3 }
1394     \circ \varphi _{\Delta t/2,\pi _2 } \circ \varphi _{\Delta t/2,\pi
1395 tim 2713 _1 }.
1396 tim 2707 \]
1397    
1398 tim 2713 The non-canonical Lie-Poisson bracket ${F, G}$ of two function
1399     $F(\pi )$ and $G(\pi )$ is defined by
1400 tim 2707 \[
1401 tim 2713 \{ F,G\} (\pi ) = [\nabla _\pi F(\pi )]^T J(\pi )\nabla _\pi G(\pi
1402     )
1403     \]
1404     If the Poisson bracket of a function $F$ with an arbitrary smooth
1405     function $G$ is zero, $F$ is a \emph{Casimir}, which is the
1406     conserved quantity in Poisson system. We can easily verify that the
1407     norm of the angular momentum, $\parallel \pi
1408     \parallel$, is a \emph{Casimir}. Let$ F(\pi ) = S(\frac{{\parallel
1409     \pi \parallel ^2 }}{2})$ for an arbitrary function $ S:R \to R$ ,
1410     then by the chain rule
1411     \[
1412     \nabla _\pi F(\pi ) = S'(\frac{{\parallel \pi \parallel ^2
1413     }}{2})\pi
1414     \]
1415     Thus $ [\nabla _\pi F(\pi )]^T J(\pi ) = - S'(\frac{{\parallel \pi
1416     \parallel ^2 }}{2})\pi \times \pi = 0 $. This explicit
1417 tim 2872 Lie-Poisson integrator is found to be both extremely efficient and
1418     stable. These properties can be explained by the fact the small
1419     angle approximation is used and the norm of the angular momentum is
1420     conserved.
1421 tim 2713
1422     \subsection{\label{introSection:RBHamiltonianSplitting} Hamiltonian
1423     Splitting for Rigid Body}
1424    
1425     The Hamiltonian of rigid body can be separated in terms of kinetic
1426     energy and potential energy,
1427     \[
1428     H = T(p,\pi ) + V(q,Q)
1429     \]
1430     The equations of motion corresponding to potential energy and
1431     kinetic energy are listed in the below table,
1432 tim 2776 \begin{table}
1433 tim 2889 \caption{EQUATIONS OF MOTION DUE TO POTENTIAL AND KINETIC ENERGIES}
1434 tim 2713 \begin{center}
1435     \begin{tabular}{|l|l|}
1436     \hline
1437     % after \\: \hline or \cline{col1-col2} \cline{col3-col4} ...
1438     Potential & Kinetic \\
1439     $\frac{{dq}}{{dt}} = \frac{p}{m}$ & $\frac{d}{{dt}}q = p$ \\
1440     $\frac{d}{{dt}}p = - \frac{{\partial V}}{{\partial q}}$ & $ \frac{d}{{dt}}p = 0$ \\
1441     $\frac{d}{{dt}}Q = 0$ & $ \frac{d}{{dt}}Q = Qskew(I^{ - 1} j)$ \\
1442     $ \frac{d}{{dt}}\pi = \sum\limits_i {\left( {Q^T F_i (r,Q)} \right) \times X_i }$ & $\frac{d}{{dt}}\pi = \pi \times I^{ - 1} \pi$\\
1443     \hline
1444     \end{tabular}
1445     \end{center}
1446 tim 2776 \end{table}
1447 tim 2872 A second-order symplectic method is now obtained by the composition
1448     of the position and velocity propagators,
1449 tim 2713 \[
1450     \varphi _{\Delta t} = \varphi _{\Delta t/2,V} \circ \varphi
1451     _{\Delta t,T} \circ \varphi _{\Delta t/2,V}.
1452     \]
1453 tim 2719 Moreover, $\varphi _{\Delta t/2,V}$ can be divided into two
1454 tim 2872 sub-propagators which corresponding to force and torque
1455     respectively,
1456 tim 2713 \[
1457 tim 2707 \varphi _{\Delta t/2,V} = \varphi _{\Delta t/2,F} \circ \varphi
1458 tim 2713 _{\Delta t/2,\tau }.
1459 tim 2707 \]
1460 tim 2713 Since the associated operators of $\varphi _{\Delta t/2,F} $ and
1461 tim 2872 $\circ \varphi _{\Delta t/2,\tau }$ commute, the composition order
1462     inside $\varphi _{\Delta t/2,V}$ does not matter. Furthermore, the
1463     kinetic energy can be separated to translational kinetic term, $T^t
1464     (p)$, and rotational kinetic term, $T^r (\pi )$,
1465 tim 2713 \begin{equation}
1466     T(p,\pi ) =T^t (p) + T^r (\pi ).
1467     \end{equation}
1468     where $ T^t (p) = \frac{1}{2}p^T m^{ - 1} p $ and $T^r (\pi )$ is
1469     defined by \ref{introEquation:rotationalKineticRB}. Therefore, the
1470 tim 2872 corresponding propagators are given by
1471 tim 2713 \[
1472     \varphi _{\Delta t,T} = \varphi _{\Delta t,T^t } \circ \varphi
1473     _{\Delta t,T^r }.
1474     \]
1475 tim 2872 Finally, we obtain the overall symplectic propagators for freely
1476     moving rigid bodies
1477 tim 2713 \begin{equation}
1478     \begin{array}{c}
1479     \varphi _{\Delta t} = \varphi _{\Delta t/2,F} \circ \varphi _{\Delta t/2,\tau } \\
1480     \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 } \\
1481     \circ \varphi _{\Delta t/2,\tau } \circ \varphi _{\Delta t/2,F} .\\
1482     \end{array}
1483     \label{introEquation:overallRBFlowMaps}
1484     \end{equation}
1485 tim 2707
1486 tim 2685 \section{\label{introSection:langevinDynamics}Langevin Dynamics}
1487 tim 2716 As an alternative to newtonian dynamics, Langevin dynamics, which
1488     mimics a simple heat bath with stochastic and dissipative forces,
1489     has been applied in a variety of studies. This section will review
1490 tim 2872 the theory of Langevin dynamics. A brief derivation of generalized
1491     Langevin equation will be given first. Following that, we will
1492     discuss the physical meaning of the terms appearing in the equation
1493     as well as the calculation of friction tensor from hydrodynamics
1494     theory.
1495 tim 2685
1496 tim 2719 \subsection{\label{introSection:generalizedLangevinDynamics}Derivation of Generalized Langevin Equation}
1497 tim 2685
1498 tim 2872 A harmonic bath model, in which an effective set of harmonic
1499 tim 2719 oscillators are used to mimic the effect of a linearly responding
1500     environment, has been widely used in quantum chemistry and
1501     statistical mechanics. One of the successful applications of
1502 tim 2872 Harmonic bath model is the derivation of the Generalized Langevin
1503     Dynamics (GLE). Lets consider a system, in which the degree of
1504 tim 2719 freedom $x$ is assumed to couple to the bath linearly, giving a
1505     Hamiltonian of the form
1506 tim 2696 \begin{equation}
1507     H = \frac{{p^2 }}{{2m}} + U(x) + H_B + \Delta U(x,x_1 , \ldots x_N)
1508 tim 2719 \label{introEquation:bathGLE}.
1509 tim 2696 \end{equation}
1510 tim 2872 Here $p$ is a momentum conjugate to $x$, $m$ is the mass associated
1511     with this degree of freedom, $H_B$ is a harmonic bath Hamiltonian,
1512 tim 2696 \[
1513 tim 2719 H_B = \sum\limits_{\alpha = 1}^N {\left\{ {\frac{{p_\alpha ^2
1514     }}{{2m_\alpha }} + \frac{1}{2}m_\alpha \omega _\alpha ^2 }
1515     \right\}}
1516 tim 2696 \]
1517 tim 2719 where the index $\alpha$ runs over all the bath degrees of freedom,
1518     $\omega _\alpha$ are the harmonic bath frequencies, $m_\alpha$ are
1519 tim 2872 the harmonic bath masses, and $\Delta U$ is a bilinear system-bath
1520 tim 2719 coupling,
1521 tim 2696 \[
1522     \Delta U = - \sum\limits_{\alpha = 1}^N {g_\alpha x_\alpha x}
1523     \]
1524 tim 2872 where $g_\alpha$ are the coupling constants between the bath
1525 tim 2874 coordinates ($x_ \alpha$) and the system coordinate ($x$).
1526 tim 2872 Introducing
1527 tim 2696 \[
1528 tim 2719 W(x) = U(x) - \sum\limits_{\alpha = 1}^N {\frac{{g_\alpha ^2
1529     }}{{2m_\alpha w_\alpha ^2 }}} x^2
1530     \] and combining the last two terms in Equation
1531     \ref{introEquation:bathGLE}, we may rewrite the Harmonic bath
1532     Hamiltonian as
1533 tim 2696 \[
1534     H = \frac{{p^2 }}{{2m}} + W(x) + \sum\limits_{\alpha = 1}^N
1535     {\left\{ {\frac{{p_\alpha ^2 }}{{2m_\alpha }} + \frac{1}{2}m_\alpha
1536     w_\alpha ^2 \left( {x_\alpha - \frac{{g_\alpha }}{{m_\alpha
1537     w_\alpha ^2 }}x} \right)^2 } \right\}}
1538     \]
1539     Since the first two terms of the new Hamiltonian depend only on the
1540     system coordinates, we can get the equations of motion for
1541 tim 2872 Generalized Langevin Dynamics by Hamilton's equations,
1542 tim 2719 \begin{equation}
1543     m\ddot x = - \frac{{\partial W(x)}}{{\partial x}} -
1544     \sum\limits_{\alpha = 1}^N {g_\alpha \left( {x_\alpha -
1545     \frac{{g_\alpha }}{{m_\alpha w_\alpha ^2 }}x} \right)},
1546     \label{introEquation:coorMotionGLE}
1547     \end{equation}
1548     and
1549     \begin{equation}
1550     m\ddot x_\alpha = - m_\alpha w_\alpha ^2 \left( {x_\alpha -
1551     \frac{{g_\alpha }}{{m_\alpha w_\alpha ^2 }}x} \right).
1552     \label{introEquation:bathMotionGLE}
1553     \end{equation}
1554 tim 2696
1555 tim 2719 In order to derive an equation for $x$, the dynamics of the bath
1556     variables $x_\alpha$ must be solved exactly first. As an integral
1557     transform which is particularly useful in solving linear ordinary
1558 tim 2872 differential equations,the Laplace transform is the appropriate tool
1559     to solve this problem. The basic idea is to transform the difficult
1560 tim 2719 differential equations into simple algebra problems which can be
1561 tim 2872 solved easily. Then, by applying the inverse Laplace transform, also
1562     known as the Bromwich integral, we can retrieve the solutions of the
1563 tim 2719 original problems.
1564 tim 2696
1565 tim 2719 Let $f(t)$ be a function defined on $ [0,\infty ) $. The Laplace
1566     transform of f(t) is a new function defined as
1567 tim 2696 \[
1568 tim 2719 L(f(t)) \equiv F(p) = \int_0^\infty {f(t)e^{ - pt} dt}
1569 tim 2696 \]
1570 tim 2719 where $p$ is real and $L$ is called the Laplace Transform
1571     Operator. Below are some important properties of Laplace transform
1572 tim 2696
1573 tim 2789 \begin{eqnarray*}
1574     L(x + y) & = & L(x) + L(y) \\
1575     L(ax) & = & aL(x) \\
1576     L(\dot x) & = & pL(x) - px(0) \\
1577     L(\ddot x)& = & p^2 L(x) - px(0) - \dot x(0) \\
1578     L\left( {\int_0^t {g(t - \tau )h(\tau )d\tau } } \right)& = & G(p)H(p) \\
1579     \end{eqnarray*}
1580    
1581    
1582 tim 2872 Applying the Laplace transform to the bath coordinates, we obtain
1583 tim 2789 \begin{eqnarray*}
1584     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) \\
1585     L(x_\alpha ) & = & \frac{{\frac{{g_\alpha }}{{\omega _\alpha }}L(x) + px_\alpha (0) + \dot x_\alpha (0)}}{{p^2 + \omega _\alpha ^2 }} \\
1586     \end{eqnarray*}
1587    
1588 tim 2719 By the same way, the system coordinates become
1589 tim 2789 \begin{eqnarray*}
1590     mL(\ddot x) & = & - \frac{1}{p}\frac{{\partial W(x)}}{{\partial x}} \\
1591     & & \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\}} \\
1592     \end{eqnarray*}
1593 tim 2696
1594 tim 2719 With the help of some relatively important inverse Laplace
1595     transformations:
1596 tim 2696 \[
1597 tim 2719 \begin{array}{c}
1598     L(\cos at) = \frac{p}{{p^2 + a^2 }} \\
1599     L(\sin at) = \frac{a}{{p^2 + a^2 }} \\
1600     L(1) = \frac{1}{p} \\
1601     \end{array}
1602 tim 2696 \]
1603 tim 2719 , we obtain
1604 tim 2794 \begin{eqnarray*}
1605     m\ddot x & = & - \frac{{\partial W(x)}}{{\partial x}} -
1606 tim 2696 \sum\limits_{\alpha = 1}^N {\left\{ {\left( { - \frac{{g_\alpha ^2
1607     }}{{m_\alpha \omega _\alpha ^2 }}} \right)\int_0^t {\cos (\omega
1608 tim 2794 _\alpha t)\dot x(t - \tau )d\tau } } \right\}} \\
1609     & & + \sum\limits_{\alpha = 1}^N {\left\{ {\left[ {g_\alpha
1610     x_\alpha (0) - \frac{{g_\alpha }}{{m_\alpha \omega _\alpha }}}
1611     \right]\cos (\omega _\alpha t) + \frac{{g_\alpha \dot x_\alpha
1612     (0)}}{{\omega _\alpha }}\sin (\omega _\alpha t)} \right\}}
1613     \end{eqnarray*}
1614     \begin{eqnarray*}
1615     m\ddot x & = & - \frac{{\partial W(x)}}{{\partial x}} - \int_0^t
1616 tim 2696 {\sum\limits_{\alpha = 1}^N {\left( { - \frac{{g_\alpha ^2
1617     }}{{m_\alpha \omega _\alpha ^2 }}} \right)\cos (\omega _\alpha
1618 tim 2794 t)\dot x(t - \tau )d} \tau } \\
1619     & & + \sum\limits_{\alpha = 1}^N {\left\{ {\left[ {g_\alpha
1620     x_\alpha (0) - \frac{{g_\alpha }}{{m_\alpha \omega _\alpha }}}
1621     \right]\cos (\omega _\alpha t) + \frac{{g_\alpha \dot x_\alpha
1622     (0)}}{{\omega _\alpha }}\sin (\omega _\alpha t)} \right\}}
1623     \end{eqnarray*}
1624 tim 2719 Introducing a \emph{dynamic friction kernel}
1625 tim 2696 \begin{equation}
1626 tim 2719 \xi (t) = \sum\limits_{\alpha = 1}^N {\left( { - \frac{{g_\alpha ^2
1627     }}{{m_\alpha \omega _\alpha ^2 }}} \right)\cos (\omega _\alpha t)}
1628     \label{introEquation:dynamicFrictionKernelDefinition}
1629     \end{equation}
1630     and \emph{a random force}
1631     \begin{equation}
1632     R(t) = \sum\limits_{\alpha = 1}^N {\left( {g_\alpha x_\alpha (0)
1633     - \frac{{g_\alpha ^2 }}{{m_\alpha \omega _\alpha ^2 }}x(0)}
1634     \right)\cos (\omega _\alpha t)} + \frac{{\dot x_\alpha
1635     (0)}}{{\omega _\alpha }}\sin (\omega _\alpha t),
1636     \label{introEquation:randomForceDefinition}
1637     \end{equation}
1638     the equation of motion can be rewritten as
1639     \begin{equation}
1640 tim 2696 m\ddot x = - \frac{{\partial W}}{{\partial x}} - \int_0^t {\xi
1641     (t)\dot x(t - \tau )d\tau } + R(t)
1642     \label{introEuqation:GeneralizedLangevinDynamics}
1643     \end{equation}
1644 tim 2719 which is known as the \emph{generalized Langevin equation}.
1645    
1646 tim 2819 \subsubsection{\label{introSection:randomForceDynamicFrictionKernel}\textbf{Random Force and Dynamic Friction Kernel}}
1647 tim 2719
1648     One may notice that $R(t)$ depends only on initial conditions, which
1649     implies it is completely deterministic within the context of a
1650     harmonic bath. However, it is easy to verify that $R(t)$ is totally
1651     uncorrelated to $x$ and $\dot x$,
1652 tim 2696 \[
1653 tim 2719 \begin{array}{l}
1654     \left\langle {x(t)R(t)} \right\rangle = 0, \\
1655     \left\langle {\dot x(t)R(t)} \right\rangle = 0. \\
1656     \end{array}
1657 tim 2696 \]
1658 tim 2719 This property is what we expect from a truly random process. As long
1659 tim 2872 as the model chosen for $R(t)$ was a gaussian distribution in
1660     general, the stochastic nature of the GLE still remains.
1661 tim 2696
1662 tim 2719 %dynamic friction kernel
1663     The convolution integral
1664 tim 2696 \[
1665 tim 2719 \int_0^t {\xi (t)\dot x(t - \tau )d\tau }
1666 tim 2696 \]
1667 tim 2719 depends on the entire history of the evolution of $x$, which implies
1668     that the bath retains memory of previous motions. In other words,
1669     the bath requires a finite time to respond to change in the motion
1670     of the system. For a sluggish bath which responds slowly to changes
1671     in the system coordinate, we may regard $\xi(t)$ as a constant
1672     $\xi(t) = \Xi_0$. Hence, the convolution integral becomes
1673     \[
1674     \int_0^t {\xi (t)\dot x(t - \tau )d\tau } = \xi _0 (x(t) - x(0))
1675     \]
1676     and Equation \ref{introEuqation:GeneralizedLangevinDynamics} becomes
1677     \[
1678     m\ddot x = - \frac{\partial }{{\partial x}}\left( {W(x) +
1679     \frac{1}{2}\xi _0 (x - x_0 )^2 } \right) + R(t),
1680     \]
1681 tim 2872 which can be used to describe the effect of dynamic caging in
1682     viscous solvents. The other extreme is the bath that responds
1683     infinitely quickly to motions in the system. Thus, $\xi (t)$ can be
1684     taken as a $delta$ function in time:
1685 tim 2719 \[
1686     \xi (t) = 2\xi _0 \delta (t)
1687     \]
1688     Hence, the convolution integral becomes
1689     \[
1690     \int_0^t {\xi (t)\dot x(t - \tau )d\tau } = 2\xi _0 \int_0^t
1691     {\delta (t)\dot x(t - \tau )d\tau } = \xi _0 \dot x(t),
1692     \]
1693     and Equation \ref{introEuqation:GeneralizedLangevinDynamics} becomes
1694     \begin{equation}
1695     m\ddot x = - \frac{{\partial W(x)}}{{\partial x}} - \xi _0 \dot
1696     x(t) + R(t) \label{introEquation:LangevinEquation}
1697     \end{equation}
1698     which is known as the Langevin equation. The static friction
1699     coefficient $\xi _0$ can either be calculated from spectral density
1700 tim 2850 or be determined by Stokes' law for regular shaped particles. A
1701 tim 2719 briefly review on calculating friction tensor for arbitrary shaped
1702 tim 2720 particles is given in Sec.~\ref{introSection:frictionTensor}.
1703 tim 2696
1704 tim 2819 \subsubsection{\label{introSection:secondFluctuationDissipation}\textbf{The Second Fluctuation Dissipation Theorem}}
1705 tim 2719
1706     Defining a new set of coordinates,
1707 tim 2696 \[
1708     q_\alpha (t) = x_\alpha (t) - \frac{1}{{m_\alpha \omega _\alpha
1709     ^2 }}x(0)
1710 tim 2719 \],
1711     we can rewrite $R(T)$ as
1712 tim 2696 \[
1713 tim 2719 R(t) = \sum\limits_{\alpha = 1}^N {g_\alpha q_\alpha (t)}.
1714 tim 2696 \]
1715     And since the $q$ coordinates are harmonic oscillators,
1716 tim 2789
1717     \begin{eqnarray*}
1718     \left\langle {q_\alpha ^2 } \right\rangle & = & \frac{{kT}}{{m_\alpha \omega _\alpha ^2 }} \\
1719     \left\langle {q_\alpha (t)q_\alpha (0)} \right\rangle & = & \left\langle {q_\alpha ^2 (0)} \right\rangle \cos (\omega _\alpha t) \\
1720     \left\langle {q_\alpha (t)q_\beta (0)} \right\rangle & = &\delta _{\alpha \beta } \left\langle {q_\alpha (t)q_\alpha (0)} \right\rangle \\
1721     \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 } } \\
1722     & = &\sum\limits_\alpha {g_\alpha ^2 \left\langle {q_\alpha ^2 (0)} \right\rangle \cos (\omega _\alpha t)} \\
1723     & = &kT\xi (t) \\
1724     \end{eqnarray*}
1725    
1726 tim 2719 Thus, we recover the \emph{second fluctuation dissipation theorem}
1727 tim 2696 \begin{equation}
1728     \xi (t) = \left\langle {R(t)R(0)} \right\rangle
1729 tim 2719 \label{introEquation:secondFluctuationDissipation}.
1730 tim 2696 \end{equation}
1731 tim 2719 In effect, it acts as a constraint on the possible ways in which one
1732     can model the random force and friction kernel.