ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/tengDissertation/Introduction.tex
Revision: 2909
Committed: Thu Jun 29 23:00:35 2006 UTC (18 years ago) by tim
Content type: application/x-tex
File size: 75904 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 2907 Using equations of motion derived from Classical Mechanics,
7     Molecular Dynamics simulations are carried out by integrating the
8     equations of motion for a given system of particles. There are three
9     fundamental ideas behind classical mechanics. Firstly, one can
10     determine the state of a mechanical system at any time of interest;
11     Secondly, all the mechanical properties of the system at that time
12     can be determined by combining the knowledge of the properties of
13     the system with the specification of this state; Finally, the
14     specification of the state when further combined with the laws of
15     mechanics will also be sufficient to predict the future behavior of
16     the system.
17 tim 2685
18 tim 2693 \subsection{\label{introSection:newtonian}Newtonian Mechanics}
19 tim 2694 The discovery of Newton's three laws of mechanics which govern the
20     motion of particles is the foundation of the classical mechanics.
21 tim 2819 Newton's first law defines a class of inertial frames. Inertial
22 tim 2694 frames are reference frames where a particle not interacting with
23     other bodies will move with constant speed in the same direction.
24 tim 2819 With respect to inertial frames, Newton's second law has the form
25 tim 2694 \begin{equation}
26 tim 2819 F = \frac {dp}{dt} = \frac {mdv}{dt}
27 tim 2694 \label{introEquation:newtonSecondLaw}
28     \end{equation}
29     A point mass interacting with other bodies moves with the
30     acceleration along the direction of the force acting on it. Let
31 tim 2702 $F_{ij}$ be the force that particle $i$ exerts on particle $j$, and
32     $F_{ji}$ be the force that particle $j$ exerts on particle $i$.
33 tim 2819 Newton's third law states that
34 tim 2694 \begin{equation}
35 tim 2898 F_{ij} = -F_{ji}.
36 tim 2694 \label{introEquation:newtonThirdLaw}
37     \end{equation}
38     Conservation laws of Newtonian Mechanics play very important roles
39     in solving mechanics problems. The linear momentum of a particle is
40     conserved if it is free or it experiences no force. The second
41     conservation theorem concerns the angular momentum of a particle.
42     The angular momentum $L$ of a particle with respect to an origin
43     from which $r$ is measured is defined to be
44     \begin{equation}
45     L \equiv r \times p \label{introEquation:angularMomentumDefinition}
46     \end{equation}
47     The torque $\tau$ with respect to the same origin is defined to be
48     \begin{equation}
49 tim 2819 \tau \equiv r \times F \label{introEquation:torqueDefinition}
50 tim 2694 \end{equation}
51     Differentiating Eq.~\ref{introEquation:angularMomentumDefinition},
52     \[
53     \dot L = \frac{d}{{dt}}(r \times p) = (\dot r \times p) + (r \times
54     \dot p)
55     \]
56     since
57     \[
58     \dot r \times p = \dot r \times mv = m\dot r \times \dot r \equiv 0
59     \]
60     thus,
61     \begin{equation}
62 tim 2819 \dot L = r \times \dot p = \tau
63 tim 2694 \end{equation}
64     If there are no external torques acting on a body, the angular
65     momentum of it is conserved. The last conservation theorem state
66 tim 2899 that if all forces are conservative, energy is conserved,
67     \begin{equation}E = T + V. \label{introEquation:energyConservation}
68 tim 2696 \end{equation}
69 tim 2899 All of these conserved quantities are important factors to determine
70     the quality of numerical integration schemes for rigid bodies
71     \cite{Dullweber1997}.
72 tim 2694
73 tim 2693 \subsection{\label{introSection:lagrangian}Lagrangian Mechanics}
74 tim 2692
75 tim 2907 Newtonian Mechanics suffers from an important limitation: motion can
76 tim 2904 only be described in cartesian coordinate systems which make it
77     impossible to predict analytically the properties of the system even
78     if we know all of the details of the interaction. In order to
79     overcome some of the practical difficulties which arise in attempts
80 tim 2907 to apply Newton's equation to complex systems, approximate numerical
81 tim 2904 procedures may be developed.
82 tim 2692
83 tim 2819 \subsubsection{\label{introSection:halmiltonPrinciple}\textbf{Hamilton's
84     Principle}}
85 tim 2692
86     Hamilton introduced the dynamical principle upon which it is
87 tim 2819 possible to base all of mechanics and most of classical physics.
88 tim 2907 Hamilton's Principle may be stated as follows: the trajectory, along
89     which a dynamical system may move from one point to another within a
90     specified time, is derived by finding the path which minimizes the
91     time integral of the difference between the kinetic $K$, and
92     potential energies $U$,
93 tim 2692 \begin{equation}
94 tim 2899 \delta \int_{t_1 }^{t_2 } {(K - U)dt = 0}.
95 tim 2693 \label{introEquation:halmitonianPrinciple1}
96 tim 2692 \end{equation}
97     For simple mechanical systems, where the forces acting on the
98 tim 2819 different parts are derivable from a potential, the Lagrangian
99     function $L$ can be defined as the difference between the kinetic
100     energy of the system and its potential energy,
101 tim 2692 \begin{equation}
102 tim 2904 L \equiv K - U = L(q_i ,\dot q_i ).
103 tim 2692 \label{introEquation:lagrangianDef}
104     \end{equation}
105 tim 2904 Thus, Eq.~\ref{introEquation:halmitonianPrinciple1} becomes
106 tim 2692 \begin{equation}
107 tim 2904 \delta \int_{t_1 }^{t_2 } {L dt = 0} .
108 tim 2693 \label{introEquation:halmitonianPrinciple2}
109 tim 2692 \end{equation}
110    
111 tim 2819 \subsubsection{\label{introSection:equationOfMotionLagrangian}\textbf{The
112     Equations of Motion in Lagrangian Mechanics}}
113 tim 2692
114 tim 2850 For a system of $f$ degrees of freedom, the equations of motion in
115     the Lagrangian form is
116 tim 2692 \begin{equation}
117     \frac{d}{{dt}}\frac{{\partial L}}{{\partial \dot q_i }} -
118     \frac{{\partial L}}{{\partial q_i }} = 0,{\rm{ }}i = 1, \ldots,f
119 tim 2693 \label{introEquation:eqMotionLagrangian}
120 tim 2692 \end{equation}
121     where $q_{i}$ is generalized coordinate and $\dot{q_{i}}$ is
122     generalized velocity.
123    
124 tim 2693 \subsection{\label{introSection:hamiltonian}Hamiltonian Mechanics}
125 tim 2692
126     Arising from Lagrangian Mechanics, Hamiltonian Mechanics was
127     introduced by William Rowan Hamilton in 1833 as a re-formulation of
128     classical mechanics. If the potential energy of a system is
129 tim 2819 independent of velocities, the momenta can be defined as
130 tim 2692 \begin{equation}
131     p_i = \frac{\partial L}{\partial \dot q_i}
132     \label{introEquation:generalizedMomenta}
133     \end{equation}
134 tim 2693 The Lagrange equations of motion are then expressed by
135 tim 2692 \begin{equation}
136 tim 2693 p_i = \frac{{\partial L}}{{\partial q_i }}
137     \label{introEquation:generalizedMomentaDot}
138     \end{equation}
139     With the help of the generalized momenta, we may now define a new
140     quantity $H$ by the equation
141     \begin{equation}
142     H = \sum\limits_k {p_k \dot q_k } - L ,
143 tim 2692 \label{introEquation:hamiltonianDefByLagrangian}
144     \end{equation}
145     where $ \dot q_1 \ldots \dot q_f $ are generalized velocities and
146 tim 2898 $L$ is the Lagrangian function for the system. Differentiating
147     Eq.~\ref{introEquation:hamiltonianDefByLagrangian}, one can obtain
148 tim 2693 \begin{equation}
149     dH = \sum\limits_k {\left( {p_k d\dot q_k + \dot q_k dp_k -
150     \frac{{\partial L}}{{\partial q_k }}dq_k - \frac{{\partial
151     L}}{{\partial \dot q_k }}d\dot q_k } \right)} - \frac{{\partial
152 tim 2904 L}}{{\partial t}}dt . \label{introEquation:diffHamiltonian1}
153 tim 2693 \end{equation}
154 tim 2899 Making use of Eq.~\ref{introEquation:generalizedMomenta}, the second
155     and fourth terms in the parentheses cancel. Therefore,
156 tim 2693 Eq.~\ref{introEquation:diffHamiltonian1} can be rewritten as
157     \begin{equation}
158     dH = \sum\limits_k {\left( {\dot q_k dp_k - \dot p_k dq_k }
159 tim 2904 \right)} - \frac{{\partial L}}{{\partial t}}dt .
160 tim 2693 \label{introEquation:diffHamiltonian2}
161     \end{equation}
162     By identifying the coefficients of $dq_k$, $dp_k$ and dt, we can
163     find
164     \begin{equation}
165 tim 2819 \frac{{\partial H}}{{\partial p_k }} = \dot {q_k}
166 tim 2693 \label{introEquation:motionHamiltonianCoordinate}
167     \end{equation}
168     \begin{equation}
169 tim 2819 \frac{{\partial H}}{{\partial q_k }} = - \dot {p_k}
170 tim 2693 \label{introEquation:motionHamiltonianMomentum}
171     \end{equation}
172     and
173     \begin{equation}
174     \frac{{\partial H}}{{\partial t}} = - \frac{{\partial L}}{{\partial
175     t}}
176     \label{introEquation:motionHamiltonianTime}
177     \end{equation}
178 tim 2899 where Eq.~\ref{introEquation:motionHamiltonianCoordinate} and
179 tim 2693 Eq.~\ref{introEquation:motionHamiltonianMomentum} are Hamilton's
180     equation of motion. Due to their symmetrical formula, they are also
181 tim 2786 known as the canonical equations of motions \cite{Goldstein2001}.
182 tim 2693
183 tim 2692 An important difference between Lagrangian approach and the
184     Hamiltonian approach is that the Lagrangian is considered to be a
185 tim 2819 function of the generalized velocities $\dot q_i$ and coordinates
186     $q_i$, while the Hamiltonian is considered to be a function of the
187     generalized momenta $p_i$ and the conjugate coordinates $q_i$.
188     Hamiltonian Mechanics is more appropriate for application to
189     statistical mechanics and quantum mechanics, since it treats the
190     coordinate and its time derivative as independent variables and it
191     only works with 1st-order differential equations\cite{Marion1990}.
192 tim 2696 In Newtonian Mechanics, a system described by conservative forces
193 tim 2899 conserves the total energy
194     (Eq.~\ref{introEquation:energyConservation}). It follows that
195 tim 2904 Hamilton's equations of motion conserve the total Hamiltonian
196 tim 2696 \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 2904 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 tim 2907 possible states of a system. Each possible state of the system
218     corresponds to one unique point in the phase space. For mechanical
219     systems, the phase space usually consists of all possible values of
220     position and momentum variables. Consider a dynamic system of $f$
221     particles in a cartesian space, where each of the $6f$ coordinates
222     and momenta is assigned to one of $6f$ mutually orthogonal axes, the
223     phase space of this system is a $6f$ dimensional space. A point, $x
224     =
225 tim 2904 (\mathord{\buildrel{\lower3pt\hbox{$\scriptscriptstyle\rightharpoonup$}}
226     \over q} _1 , \ldots
227     ,\mathord{\buildrel{\lower3pt\hbox{$\scriptscriptstyle\rightharpoonup$}}
228     \over q} _f
229     ,\mathord{\buildrel{\lower3pt\hbox{$\scriptscriptstyle\rightharpoonup$}}
230     \over p} _1 \ldots
231     ,\mathord{\buildrel{\lower3pt\hbox{$\scriptscriptstyle\rightharpoonup$}}
232     \over p} _f )$ , with a unique set of values of $6f$ coordinates and
233     momenta is a phase space vector.
234 tim 2888 %%%fix me
235 tim 2700
236 tim 2888 In statistical mechanics, the condition of an ensemble at any time
237 tim 2700 can be regarded as appropriately specified by the density $\rho$
238     with which representative points are distributed over the phase
239 tim 2819 space. The density distribution for an ensemble with $f$ degrees of
240     freedom is defined as,
241 tim 2700 \begin{equation}
242     \rho = \rho (q_1 , \ldots ,q_f ,p_1 , \ldots ,p_f ,t).
243     \label{introEquation:densityDistribution}
244     \end{equation}
245     Governed by the principles of mechanics, the phase points change
246 tim 2907 their locations which changes the density at any time at phase
247 tim 2819 space. Hence, the density distribution is also to be taken as a
248 tim 2905 function of the time. The number of systems $\delta N$ at time $t$
249     can be determined by,
250 tim 2700 \begin{equation}
251     \delta N = \rho (q,p,t)dq_1 \ldots dq_f dp_1 \ldots dp_f.
252     \label{introEquation:deltaN}
253     \end{equation}
254 tim 2907 Assuming enough copies of the systems, we can sufficiently
255 tim 2819 approximate $\delta N$ without introducing discontinuity when we go
256     from one region in the phase space to another. By integrating over
257     the whole phase space,
258 tim 2700 \begin{equation}
259     N = \int { \ldots \int {\rho (q,p,t)dq_1 } ...dq_f dp_1 } ...dp_f
260     \label{introEquation:totalNumberSystem}
261     \end{equation}
262 tim 2907 gives us an expression for the total number of copies. Hence, the
263     probability per unit volume in the phase space can be obtained by,
264 tim 2700 \begin{equation}
265     \frac{{\rho (q,p,t)}}{N} = \frac{{\rho (q,p,t)}}{{\int { \ldots \int
266     {\rho (q,p,t)dq_1 } ...dq_f dp_1 } ...dp_f }}.
267     \label{introEquation:unitProbability}
268     \end{equation}
269 tim 2850 With the help of Eq.~\ref{introEquation:unitProbability} and the
270     knowledge of the system, it is possible to calculate the average
271 tim 2700 value of any desired quantity which depends on the coordinates and
272 tim 2907 momenta of the system. Even when the dynamics of the real system are
273 tim 2700 complex, or stochastic, or even discontinuous, the average
274 tim 2907 properties of the ensemble of possibilities as a whole remain well
275     defined. For a classical system in thermal equilibrium with its
276 tim 2819 environment, the ensemble average of a mechanical quantity, $\langle
277     A(q , p) \rangle_t$, takes the form of an integral over the phase
278     space of the system,
279 tim 2700 \begin{equation}
280     \langle A(q , p) \rangle_t = \frac{{\int { \ldots \int {A(q,p)\rho
281     (q,p,t)dq_1 } ...dq_f dp_1 } ...dp_f }}{{\int { \ldots \int {\rho
282 tim 2905 (q,p,t)dq_1 } ...dq_f dp_1 } ...dp_f }}.
283 tim 2700 \label{introEquation:ensembelAverage}
284     \end{equation}
285    
286     There are several different types of ensembles with different
287     statistical characteristics. As a function of macroscopic
288 tim 2819 parameters, such as temperature \textit{etc}, the partition function
289     can be used to describe the statistical properties of a system in
290 tim 2898 thermodynamic equilibrium. As an ensemble of systems, each of which
291     is known to be thermally isolated and conserve energy, the
292     Microcanonical ensemble (NVE) has a partition function like,
293 tim 2700 \begin{equation}
294 tim 2904 \Omega (N,V,E) = e^{\beta TS}. \label{introEquation:NVEPartition}
295 tim 2700 \end{equation}
296 tim 2905 A canonical ensemble (NVT) is an ensemble of systems, each of which
297 tim 2700 can share its energy with a large heat reservoir. The distribution
298     of the total energy amongst the possible dynamical states is given
299     by the partition function,
300     \begin{equation}
301 tim 2899 \Omega (N,V,T) = e^{ - \beta A}.
302 tim 2700 \label{introEquation:NVTPartition}
303     \end{equation}
304     Here, $A$ is the Helmholtz free energy which is defined as $ A = U -
305 tim 2819 TS$. Since most experiments are carried out under constant pressure
306 tim 2850 condition, the isothermal-isobaric ensemble (NPT) plays a very
307 tim 2819 important role in molecular simulations. The isothermal-isobaric
308     ensemble allow the system to exchange energy with a heat bath of
309     temperature $T$ and to change the volume as well. Its partition
310     function is given as
311 tim 2700 \begin{equation}
312     \Delta (N,P,T) = - e^{\beta G}.
313     \label{introEquation:NPTPartition}
314     \end{equation}
315     Here, $G = U - TS + PV$, and $G$ is called Gibbs free energy.
316    
317     \subsection{\label{introSection:liouville}Liouville's theorem}
318    
319 tim 2819 Liouville's theorem is the foundation on which statistical mechanics
320     rests. It describes the time evolution of the phase space
321 tim 2700 distribution function. In order to calculate the rate of change of
322 tim 2850 $\rho$, we begin from Eq.~\ref{introEquation:deltaN}. If we consider
323     the two faces perpendicular to the $q_1$ axis, which are located at
324     $q_1$ and $q_1 + \delta q_1$, the number of phase points leaving the
325     opposite face is given by the expression,
326 tim 2700 \begin{equation}
327     \left( {\rho + \frac{{\partial \rho }}{{\partial q_1 }}\delta q_1 }
328     \right)\left( {\dot q_1 + \frac{{\partial \dot q_1 }}{{\partial q_1
329     }}\delta q_1 } \right)\delta q_2 \ldots \delta q_f \delta p_1
330     \ldots \delta p_f .
331     \end{equation}
332     Summing all over the phase space, we obtain
333     \begin{equation}
334     \frac{{d(\delta N)}}{{dt}} = - \sum\limits_{i = 1}^f {\left[ {\rho
335     \left( {\frac{{\partial \dot q_i }}{{\partial q_i }} +
336     \frac{{\partial \dot p_i }}{{\partial p_i }}} \right) + \left(
337     {\frac{{\partial \rho }}{{\partial q_i }}\dot q_i + \frac{{\partial
338     \rho }}{{\partial p_i }}\dot p_i } \right)} \right]} \delta q_1
339     \ldots \delta q_f \delta p_1 \ldots \delta p_f .
340     \end{equation}
341     Differentiating the equations of motion in Hamiltonian formalism
342     (\ref{introEquation:motionHamiltonianCoordinate},
343     \ref{introEquation:motionHamiltonianMomentum}), we can show,
344     \begin{equation}
345     \sum\limits_i {\left( {\frac{{\partial \dot q_i }}{{\partial q_i }}
346     + \frac{{\partial \dot p_i }}{{\partial p_i }}} \right)} = 0 ,
347     \end{equation}
348     which cancels the first terms of the right hand side. Furthermore,
349 tim 2819 dividing $ \delta q_1 \ldots \delta q_f \delta p_1 \ldots \delta
350 tim 2700 p_f $ in both sides, we can write out Liouville's theorem in a
351     simple form,
352     \begin{equation}
353     \frac{{\partial \rho }}{{\partial t}} + \sum\limits_{i = 1}^f
354     {\left( {\frac{{\partial \rho }}{{\partial q_i }}\dot q_i +
355     \frac{{\partial \rho }}{{\partial p_i }}\dot p_i } \right)} = 0 .
356     \label{introEquation:liouvilleTheorem}
357     \end{equation}
358     Liouville's theorem states that the distribution function is
359     constant along any trajectory in phase space. In classical
360 tim 2907 statistical mechanics, since the number of system copies in an
361     ensemble is huge and constant, we can assume the local density has
362     no reason (other than classical mechanics) to change,
363 tim 2700 \begin{equation}
364     \frac{{\partial \rho }}{{\partial t}} = 0.
365     \label{introEquation:stationary}
366     \end{equation}
367     In such stationary system, the density of distribution $\rho$ can be
368     connected to the Hamiltonian $H$ through Maxwell-Boltzmann
369     distribution,
370     \begin{equation}
371     \rho \propto e^{ - \beta H}
372     \label{introEquation:densityAndHamiltonian}
373     \end{equation}
374    
375 tim 2819 \subsubsection{\label{introSection:phaseSpaceConservation}\textbf{Conservation of Phase Space}}
376 tim 2702 Lets consider a region in the phase space,
377     \begin{equation}
378     \delta v = \int { \ldots \int {dq_1 } ...dq_f dp_1 } ..dp_f .
379     \end{equation}
380     If this region is small enough, the density $\rho$ can be regarded
381 tim 2819 as uniform over the whole integral. Thus, the number of phase points
382     inside this region is given by,
383 tim 2702 \begin{equation}
384     \delta N = \rho \delta v = \rho \int { \ldots \int {dq_1 } ...dq_f
385     dp_1 } ..dp_f.
386     \end{equation}
387    
388     \begin{equation}
389     \frac{{d(\delta N)}}{{dt}} = \frac{{d\rho }}{{dt}}\delta v + \rho
390     \frac{d}{{dt}}(\delta v) = 0.
391     \end{equation}
392 tim 2907 With the help of the stationary assumption
393     (Eq.~\ref{introEquation:stationary}), we obtain the principle of
394 tim 2819 \emph{conservation of volume in phase space},
395 tim 2702 \begin{equation}
396     \frac{d}{{dt}}(\delta v) = \frac{d}{{dt}}\int { \ldots \int {dq_1 }
397     ...dq_f dp_1 } ..dp_f = 0.
398     \label{introEquation:volumePreserving}
399     \end{equation}
400    
401 tim 2819 \subsubsection{\label{introSection:liouvilleInOtherForms}\textbf{Liouville's Theorem in Other Forms}}
402 tim 2702
403 tim 2907 Liouville's theorem can be expressed in a variety of different forms
404 tim 2700 which are convenient within different contexts. For any two function
405     $F$ and $G$ of the coordinates and momenta of a system, the Poisson
406     bracket ${F, G}$ is defined as
407     \begin{equation}
408     \left\{ {F,G} \right\} = \sum\limits_i {\left( {\frac{{\partial
409     F}}{{\partial q_i }}\frac{{\partial G}}{{\partial p_i }} -
410     \frac{{\partial F}}{{\partial p_i }}\frac{{\partial G}}{{\partial
411     q_i }}} \right)}.
412     \label{introEquation:poissonBracket}
413     \end{equation}
414 tim 2905 Substituting equations of motion in Hamiltonian formalism
415     (Eq.~\ref{introEquation:motionHamiltonianCoordinate} ,
416     Eq.~\ref{introEquation:motionHamiltonianMomentum}) into
417 tim 2850 (Eq.~\ref{introEquation:liouvilleTheorem}), we can rewrite
418     Liouville's theorem using Poisson bracket notion,
419 tim 2700 \begin{equation}
420     \left( {\frac{{\partial \rho }}{{\partial t}}} \right) = - \left\{
421     {\rho ,H} \right\}.
422     \label{introEquation:liouvilleTheromInPoissin}
423     \end{equation}
424     Moreover, the Liouville operator is defined as
425     \begin{equation}
426     iL = \sum\limits_{i = 1}^f {\left( {\frac{{\partial H}}{{\partial
427     p_i }}\frac{\partial }{{\partial q_i }} - \frac{{\partial
428     H}}{{\partial q_i }}\frac{\partial }{{\partial p_i }}} \right)}
429     \label{introEquation:liouvilleOperator}
430     \end{equation}
431     In terms of Liouville operator, Liouville's equation can also be
432     expressed as
433     \begin{equation}
434     \left( {\frac{{\partial \rho }}{{\partial t}}} \right) = - iL\rho
435     \label{introEquation:liouvilleTheoremInOperator}
436     \end{equation}
437 tim 2907 which can help define a propagator $\rho (t) = e^{-iLt} \rho (0)$.
438 tim 2693 \subsection{\label{introSection:ergodic}The Ergodic Hypothesis}
439 tim 2692
440 tim 2695 Various thermodynamic properties can be calculated from Molecular
441     Dynamics simulation. By comparing experimental values with the
442     calculated properties, one can determine the accuracy of the
443 tim 2819 simulation and the quality of the underlying model. However, both
444     experiments and computer simulations are usually performed during a
445 tim 2695 certain time interval and the measurements are averaged over a
446 tim 2907 period of time which is different from the average behavior of
447 tim 2819 many-body system in Statistical Mechanics. Fortunately, the Ergodic
448     Hypothesis makes a connection between time average and the ensemble
449     average. It states that the time average and average over the
450 tim 2905 statistical ensemble are identical \cite{Frenkel1996, Leach2001}:
451 tim 2695 \begin{equation}
452 tim 2700 \langle A(q , p) \rangle_t = \mathop {\lim }\limits_{t \to \infty }
453     \frac{1}{t}\int\limits_0^t {A(q(t),p(t))dt = \int\limits_\Gamma
454     {A(q(t),p(t))} } \rho (q(t), p(t)) dqdp
455 tim 2695 \end{equation}
456 tim 2700 where $\langle A(q , p) \rangle_t$ is an equilibrium value of a
457     physical quantity and $\rho (p(t), q(t))$ is the equilibrium
458     distribution function. If an observation is averaged over a
459 tim 2907 sufficiently long time (longer than the relaxation time), all
460     accessible microstates in phase space are assumed to be equally
461     probed, giving a properly weighted statistical average. This allows
462     the researcher freedom of choice when deciding how best to measure a
463     given observable. In case an ensemble averaged approach sounds most
464 tim 2905 reasonable, the Monte Carlo methods\cite{Metropolis1949} can be
465 tim 2700 utilized. Or if the system lends itself to a time averaging
466     approach, the Molecular Dynamics techniques in
467     Sec.~\ref{introSection:molecularDynamics} will be the best
468     choice\cite{Frenkel1996}.
469 tim 2694
470 tim 2697 \section{\label{introSection:geometricIntegratos}Geometric Integrators}
471 tim 2819 A variety of numerical integrators have been proposed to simulate
472     the motions of atoms in MD simulation. They usually begin with
473     initial conditionals and move the objects in the direction governed
474     by the differential equations. However, most of them ignore the
475     hidden physical laws contained within the equations. Since 1990,
476     geometric integrators, which preserve various phase-flow invariants
477 tim 2907 such as symplectic structure, volume and time reversal symmetry,
478     were developed to address this issue\cite{Dullweber1997,
479     McLachlan1998, Leimkuhler1999}. The velocity Verlet method, which
480     happens to be a simple example of symplectic integrator, continues
481     to gain popularity in the molecular dynamics community. This fact
482     can be partly explained by its geometric nature.
483 tim 2697
484 tim 2819 \subsection{\label{introSection:symplecticManifold}Symplectic Manifolds}
485     A \emph{manifold} is an abstract mathematical space. It looks
486     locally like Euclidean space, but when viewed globally, it may have
487     more complicated structure. A good example of manifold is the
488     surface of Earth. It seems to be flat locally, but it is round if
489     viewed as a whole. A \emph{differentiable manifold} (also known as
490     \emph{smooth manifold}) is a manifold on which it is possible to
491 tim 2907 apply calculus\cite{Hirsch1997}. A \emph{symplectic manifold} is
492     defined as a pair $(M, \omega)$ which consists of a
493 tim 2697 \emph{differentiable manifold} $M$ and a close, non-degenerated,
494     bilinear symplectic form, $\omega$. A symplectic form on a vector
495     space $V$ is a function $\omega(x, y)$ which satisfies
496     $\omega(\lambda_1x_1+\lambda_2x_2, y) = \lambda_1\omega(x_1, y)+
497     \lambda_2\omega(x_2, y)$, $\omega(x, y) = - \omega(y, x)$ and
498 tim 2907 $\omega(x, x) = 0$\cite{McDuff1998}. The cross product operation in
499     vector field is an example of symplectic form. One of the
500     motivations to study \emph{symplectic manifolds} in Hamiltonian
501     Mechanics is that a symplectic manifold can represent all possible
502     configurations of the system and the phase space of the system can
503     be described by it's cotangent bundle\cite{Jost2002}. Every
504     symplectic manifold is even dimensional. For instance, in Hamilton
505     equations, coordinate and momentum always appear in pairs.
506 tim 2697
507 tim 2698 \subsection{\label{introSection:ODE}Ordinary Differential Equations}
508 tim 2697
509 tim 2819 For an ordinary differential system defined as
510 tim 2698 \begin{equation}
511     \dot x = f(x)
512     \end{equation}
513 tim 2819 where $x = x(q,p)^T$, this system is a canonical Hamiltonian, if
514 tim 2907 $f(x) = J\nabla _x H(x)$. Here, $H = H (q, p)$ is Hamiltonian
515 tim 2905 function and $J$ is the skew-symmetric matrix
516 tim 2698 \begin{equation}
517     J = \left( {\begin{array}{*{20}c}
518     0 & I \\
519     { - I} & 0 \\
520     \end{array}} \right)
521     \label{introEquation:canonicalMatrix}
522     \end{equation}
523     where $I$ is an identity matrix. Using this notation, Hamiltonian
524     system can be rewritten as,
525     \begin{equation}
526 tim 2905 \frac{d}{{dt}}x = J\nabla _x H(x).
527 tim 2698 \label{introEquation:compactHamiltonian}
528     \end{equation}In this case, $f$ is
529 tim 2899 called a \emph{Hamiltonian vector field}. Another generalization of
530     Hamiltonian dynamics is Poisson Dynamics\cite{Olver1986},
531 tim 2698 \begin{equation}
532     \dot x = J(x)\nabla _x H \label{introEquation:poissonHamiltonian}
533     \end{equation}
534     The most obvious change being that matrix $J$ now depends on $x$.
535    
536 tim 2907 \subsection{\label{introSection:exactFlow}Exact Propagator}
537 tim 2702
538 tim 2905 Let $x(t)$ be the exact solution of the ODE
539     system,$\frac{{dx}}{{dt}} = f(x) \label{introEquation:ODE}$, we can
540 tim 2907 define its exact propagator(solution) $\varphi_\tau$
541 tim 2905 \[ x(t+\tau)
542     =\varphi_\tau(x(t))
543 tim 2698 \]
544     where $\tau$ is a fixed time step and $\varphi$ is a map from phase
545 tim 2907 space to itself. The propagator has the continuous group property,
546 tim 2698 \begin{equation}
547 tim 2702 \varphi _{\tau _1 } \circ \varphi _{\tau _2 } = \varphi _{\tau _1
548     + \tau _2 } .
549     \end{equation}
550     In particular,
551     \begin{equation}
552     \varphi _\tau \circ \varphi _{ - \tau } = I
553     \end{equation}
554 tim 2907 Therefore, the exact propagator is self-adjoint,
555 tim 2702 \begin{equation}
556     \varphi _\tau = \varphi _{ - \tau }^{ - 1}.
557     \end{equation}
558 tim 2907 The exact propagator can also be written in terms of operator,
559 tim 2702 \begin{equation}
560     \varphi _\tau (x) = e^{\tau \sum\limits_i {f_i (x)\frac{\partial
561     }{{\partial x_i }}} } (x) \equiv \exp (\tau f)(x).
562     \label{introEquation:exponentialOperator}
563     \end{equation}
564 tim 2907 In most cases, it is not easy to find the exact propagator
565     $\varphi_\tau$. Instead, we use an approximate map, $\psi_\tau$,
566     which is usually called an integrator. The order of an integrator
567     $\psi_\tau$ is $p$, if the Taylor series of $\psi_\tau$ agree to
568     order $p$,
569 tim 2702 \begin{equation}
570 tim 2872 \psi_\tau(x) = x + \tau f(x) + O(\tau^{p+1})
571 tim 2698 \end{equation}
572    
573 tim 2702 \subsection{\label{introSection:geometricProperties}Geometric Properties}
574    
575 tim 2872 The hidden geometric properties\cite{Budd1999, Marsden1998} of an
576 tim 2907 ODE and its propagator play important roles in numerical studies.
577     Many of them can be found in systems which occur naturally in
578     applications. Let $\varphi$ be the propagator of Hamiltonian vector
579     field, $\varphi$ is a \emph{symplectic} propagator if it satisfies,
580 tim 2698 \begin{equation}
581 tim 2703 {\varphi '}^T J \varphi ' = J.
582 tim 2698 \end{equation}
583     According to Liouville's theorem, the symplectic volume is invariant
584 tim 2907 under a Hamiltonian propagator, which is the basis for classical
585     statistical mechanics. Furthermore, the propagator of a Hamiltonian
586     vector field on a symplectic manifold can be shown to be a
587 tim 2699 symplectomorphism. As to the Poisson system,
588 tim 2698 \begin{equation}
589 tim 2703 {\varphi '}^T J \varphi ' = J \circ \varphi
590 tim 2698 \end{equation}
591 tim 2898 is the property that must be preserved by the integrator. It is
592 tim 2907 possible to construct a \emph{volume-preserving} propagator for a
593     source free ODE ($ \nabla \cdot f = 0 $), if the propagator
594     satisfies $ \det d\varphi = 1$. One can show easily that a
595     symplectic propagator will be volume-preserving. Changing the
596     variables $y = h(x)$ in an ODE (Eq.~\ref{introEquation:ODE}) will
597     result in a new system,
598 tim 2698 \[
599     \dot y = \tilde f(y) = ((dh \cdot f)h^{ - 1} )(y).
600     \]
601     The vector filed $f$ has reversing symmetry $h$ if $f = - \tilde f$.
602 tim 2907 In other words, the propagator of this vector field is reversible if
603     and only if $ h \circ \varphi ^{ - 1} = \varphi \circ h $. A
604     conserved quantity of a general differential function is a function
605     $ G:R^{2d} \to R^d $ which is constant for all solutions of the ODE
606     $\frac{{dx}}{{dt}} = f(x)$ ,
607 tim 2705 \[
608     \frac{{dG(x(t))}}{{dt}} = 0.
609     \]
610 tim 2907 Using the chain rule, one may obtain,
611 tim 2705 \[
612 tim 2905 \sum\limits_i {\frac{{dG}}{{dx_i }}} f_i (x) = f \dot \nabla G,
613 tim 2705 \]
614 tim 2907 which is the condition for conserved quantities. For a canonical
615     Hamiltonian system, the time evolution of an arbitrary smooth
616     function $G$ is given by,
617 tim 2789 \begin{eqnarray}
618 tim 2905 \frac{{dG(x(t))}}{{dt}} & = & [\nabla _x G(x(t))]^T \dot x(t) \notag\\
619     & = & [\nabla _x G(x(t))]^T J\nabla _x H(x(t)).
620 tim 2705 \label{introEquation:firstIntegral1}
621 tim 2789 \end{eqnarray}
622 tim 2905 Using poisson bracket notion, Eq.~\ref{introEquation:firstIntegral1}
623     can be rewritten as
624 tim 2705 \[
625     \frac{d}{{dt}}G(x(t)) = \left\{ {G,H} \right\}(x(t)).
626     \]
627 tim 2907 Therefore, the sufficient condition for $G$ to be a conserved
628     quantity of a Hamiltonian system is $\left\{ {G,H} \right\} = 0.$ As
629     is well known, the Hamiltonian (or energy) H of a Hamiltonian system
630     is a conserved quantity, which is due to the fact $\{ H,H\} = 0$.
631     When designing any numerical methods, one should always try to
632     preserve the structural properties of the original ODE and its
633     propagator.
634 tim 2702
635 tim 2699 \subsection{\label{introSection:constructionSymplectic}Construction of Symplectic Methods}
636     A lot of well established and very effective numerical methods have
637 tim 2907 been successful precisely because of their symplectic nature even
638 tim 2699 though this fact was not recognized when they were first
639 tim 2872 constructed. The most famous example is the Verlet-leapfrog method
640 tim 2819 in molecular dynamics. In general, symplectic integrators can be
641 tim 2699 constructed using one of four different methods.
642     \begin{enumerate}
643     \item Generating functions
644     \item Variational methods
645     \item Runge-Kutta methods
646     \item Splitting methods
647     \end{enumerate}
648 tim 2907 Generating functions\cite{Channell1990} tend to lead to methods
649 tim 2789 which are cumbersome and difficult to use. In dissipative systems,
650     variational methods can capture the decay of energy
651 tim 2907 accurately\cite{Kane2000}. Since they are geometrically unstable
652 tim 2789 against non-Hamiltonian perturbations, ordinary implicit Runge-Kutta
653     methods are not suitable for Hamiltonian system. Recently, various
654 tim 2905 high-order explicit Runge-Kutta methods \cite{Owren1992,Chen2003}
655     have been developed to overcome this instability. However, due to
656     computational penalty involved in implementing the Runge-Kutta
657     methods, they have not attracted much attention from the Molecular
658     Dynamics community. Instead, splitting methods have been widely
659     accepted since they exploit natural decompositions of the
660     system\cite{Tuckerman1992, McLachlan1998}.
661 tim 2702
662 tim 2819 \subsubsection{\label{introSection:splittingMethod}\textbf{Splitting Methods}}
663 tim 2702
664     The main idea behind splitting methods is to decompose the discrete
665 tim 2907 $\varphi_h$ as a composition of simpler propagators,
666 tim 2699 \begin{equation}
667     \varphi _h = \varphi _{h_1 } \circ \varphi _{h_2 } \ldots \circ
668     \varphi _{h_n }
669     \label{introEquation:FlowDecomposition}
670     \end{equation}
671 tim 2907 where each of the sub-propagator is chosen such that each represent
672     a simpler integration of the system. Suppose that a Hamiltonian
673     system takes the form,
674 tim 2702 \[
675     H = H_1 + H_2.
676     \]
677     Here, $H_1$ and $H_2$ may represent different physical processes of
678     the system. For instance, they may relate to kinetic and potential
679     energy respectively, which is a natural decomposition of the
680 tim 2907 problem. If $H_1$ and $H_2$ can be integrated using exact
681     propagators $\varphi_1(t)$ and $\varphi_2(t)$, respectively, a
682     simple first order expression is then given by the Lie-Trotter
683     formula
684 tim 2699 \begin{equation}
685 tim 2702 \varphi _h = \varphi _{1,h} \circ \varphi _{2,h},
686     \label{introEquation:firstOrderSplitting}
687     \end{equation}
688     where $\varphi _h$ is the result of applying the corresponding
689     continuous $\varphi _i$ over a time $h$. By definition, as
690     $\varphi_i(t)$ is the exact solution of a Hamiltonian system, it
691     must follow that each operator $\varphi_i(t)$ is a symplectic map.
692 tim 2907 It is easy to show that any composition of symplectic propagators
693     yields a symplectic map,
694 tim 2702 \begin{equation}
695 tim 2699 (\varphi '\phi ')^T J\varphi '\phi ' = \phi '^T \varphi '^T J\varphi
696 tim 2702 '\phi ' = \phi '^T J\phi ' = J,
697 tim 2699 \label{introEquation:SymplecticFlowComposition}
698     \end{equation}
699 tim 2702 where $\phi$ and $\psi$ both are symplectic maps. Thus operator
700     splitting in this context automatically generates a symplectic map.
701 tim 2905 The Lie-Trotter
702     splitting(Eq.~\ref{introEquation:firstOrderSplitting}) introduces
703 tim 2907 local errors proportional to $h^2$, while the Strang splitting gives
704     a second-order decomposition,
705 tim 2702 \begin{equation}
706     \varphi _h = \varphi _{1,h/2} \circ \varphi _{2,h} \circ \varphi
707 tim 2706 _{1,h/2} , \label{introEquation:secondOrderSplitting}
708 tim 2702 \end{equation}
709 tim 2907 which has a local error proportional to $h^3$. The Strang
710 tim 2819 splitting's popularity in molecular simulation community attribute
711     to its symmetric property,
712 tim 2702 \begin{equation}
713     \varphi _h^{ - 1} = \varphi _{ - h}.
714 tim 2703 \label{introEquation:timeReversible}
715 tim 2882 \end{equation}
716 tim 2702
717 tim 2872 \subsubsection{\label{introSection:exampleSplittingMethod}\textbf{Examples of the Splitting Method}}
718 tim 2702 The classical equation for a system consisting of interacting
719     particles can be written in Hamiltonian form,
720     \[
721     H = T + V
722     \]
723     where $T$ is the kinetic energy and $V$ is the potential energy.
724 tim 2872 Setting $H_1 = T, H_2 = V$ and applying the Strang splitting, one
725 tim 2702 obtains the following:
726     \begin{align}
727     q(\Delta t) &= q(0) + \dot{q}(0)\Delta t +
728     \frac{F[q(0)]}{m}\frac{\Delta t^2}{2}, %
729     \label{introEquation:Lp10a} \\%
730     %
731     \dot{q}(\Delta t) &= \dot{q}(0) + \frac{\Delta t}{2m}
732     \biggl [F[q(0)] + F[q(\Delta t)] \biggr]. %
733     \label{introEquation:Lp10b}
734     \end{align}
735     where $F(t)$ is the force at time $t$. This integration scheme is
736     known as \emph{velocity verlet} which is
737     symplectic(\ref{introEquation:SymplecticFlowComposition}),
738     time-reversible(\ref{introEquation:timeReversible}) and
739     volume-preserving (\ref{introEquation:volumePreserving}). These
740     geometric properties attribute to its long-time stability and its
741     popularity in the community. However, the most commonly used
742     velocity verlet integration scheme is written as below,
743     \begin{align}
744     \dot{q}\biggl (\frac{\Delta t}{2}\biggr ) &=
745     \dot{q}(0) + \frac{\Delta t}{2m}\, F[q(0)], \label{introEquation:Lp9a}\\%
746     %
747     q(\Delta t) &= q(0) + \Delta t\, \dot{q}\biggl (\frac{\Delta t}{2}\biggr ),%
748     \label{introEquation:Lp9b}\\%
749     %
750     \dot{q}(\Delta t) &= \dot{q}\biggl (\frac{\Delta t}{2}\biggr ) +
751 tim 2872 \frac{\Delta t}{2m}\, F[q(t)]. \label{introEquation:Lp9c}
752 tim 2702 \end{align}
753     From the preceding splitting, one can see that the integration of
754     the equations of motion would follow:
755     \begin{enumerate}
756     \item calculate the velocities at the half step, $\frac{\Delta t}{2}$, from the forces calculated at the initial position.
757    
758     \item Use the half step velocities to move positions one whole step, $\Delta t$.
759    
760 tim 2872 \item Evaluate the forces at the new positions, $\mathbf{q}(\Delta t)$, and use the new forces to complete the velocity move.
761 tim 2702
762     \item Repeat from step 1 with the new position, velocities, and forces assuming the roles of the initial values.
763     \end{enumerate}
764 tim 2872 By simply switching the order of the propagators in the splitting
765     and composing a new integrator, the \emph{position verlet}
766     integrator, can be generated,
767 tim 2702 \begin{align}
768     \dot q(\Delta t) &= \dot q(0) + \Delta tF(q(0))\left[ {q(0) +
769     \frac{{\Delta t}}{{2m}}\dot q(0)} \right], %
770     \label{introEquation:positionVerlet1} \\%
771     %
772 tim 2703 q(\Delta t) &= q(0) + \frac{{\Delta t}}{2}\left[ {\dot q(0) + \dot
773 tim 2702 q(\Delta t)} \right]. %
774 tim 2719 \label{introEquation:positionVerlet2}
775 tim 2702 \end{align}
776    
777 tim 2819 \subsubsection{\label{introSection:errorAnalysis}\textbf{Error Analysis and Higher Order Methods}}
778 tim 2702
779 tim 2872 The Baker-Campbell-Hausdorff formula can be used to determine the
780 tim 2907 local error of a splitting method in terms of the commutator of the
781 tim 2702 operators(\ref{introEquation:exponentialOperator}) associated with
782 tim 2907 the sub-propagator. For operators $hX$ and $hY$ which are associated
783     with $\varphi_1(t)$ and $\varphi_2(t)$ respectively , we have
784 tim 2702 \begin{equation}
785     \exp (hX + hY) = \exp (hZ)
786     \end{equation}
787     where
788     \begin{equation}
789     hZ = hX + hY + \frac{{h^2 }}{2}[X,Y] + \frac{{h^3 }}{2}\left(
790     {[X,[X,Y]] + [Y,[Y,X]]} \right) + \ldots .
791     \end{equation}
792 tim 2907 Here, $[X,Y]$ is the commutator of operator $X$ and $Y$ given by
793 tim 2702 \[
794     [X,Y] = XY - YX .
795     \]
796 tim 2872 Applying the Baker-Campbell-Hausdorff formula\cite{Varadarajan1974}
797 tim 2907 to the Strang splitting, we can obtain
798 tim 2779 \begin{eqnarray*}
799 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 \\
800     & & \mbox{} + h^2 [X,X]/8 + h^2 [Y,Y]/8 \\
801 tim 2905 & & \mbox{} + h^3 [Y,[Y,X]]/12 - h^3[X,[X,Y]]/24 + \ldots
802     ).
803 tim 2779 \end{eqnarray*}
804 tim 2905 Since $ [X,Y] + [Y,X] = 0$ and $ [X,X] = 0$, the dominant local
805 tim 2907 error of Strang splitting is proportional to $h^3$. The same
806 tim 2905 procedure can be applied to a general splitting of the form
807 tim 2702 \begin{equation}
808     \varphi _{b_m h}^2 \circ \varphi _{a_m h}^1 \circ \varphi _{b_{m -
809     1} h}^2 \circ \ldots \circ \varphi _{a_1 h}^1 .
810     \end{equation}
811 tim 2872 A careful choice of coefficient $a_1 \ldots b_m$ will lead to higher
812     order methods. Yoshida proposed an elegant way to compose higher
813 tim 2789 order methods based on symmetric splitting\cite{Yoshida1990}. Given
814     a symmetric second order base method $ \varphi _h^{(2)} $, a
815     fourth-order symmetric method can be constructed by composing,
816 tim 2702 \[
817     \varphi _h^{(4)} = \varphi _{\alpha h}^{(2)} \circ \varphi _{\beta
818     h}^{(2)} \circ \varphi _{\alpha h}^{(2)}
819     \]
820     where $ \alpha = - \frac{{2^{1/3} }}{{2 - 2^{1/3} }}$ and $ \beta
821     = \frac{{2^{1/3} }}{{2 - 2^{1/3} }}$. Moreover, a symmetric
822     integrator $ \varphi _h^{(2n + 2)}$ can be composed by
823     \begin{equation}
824     \varphi _h^{(2n + 2)} = \varphi _{\alpha h}^{(2n)} \circ \varphi
825 tim 2872 _{\beta h}^{(2n)} \circ \varphi _{\alpha h}^{(2n)},
826 tim 2702 \end{equation}
827 tim 2872 if the weights are chosen as
828 tim 2702 \[
829     \alpha = - \frac{{2^{1/(2n + 1)} }}{{2 - 2^{1/(2n + 1)} }},\beta =
830     \frac{{2^{1/(2n + 1)} }}{{2 - 2^{1/(2n + 1)} }} .
831     \]
832    
833 tim 2694 \section{\label{introSection:molecularDynamics}Molecular Dynamics}
834    
835 tim 2720 As one of the principal tools of molecular modeling, Molecular
836     dynamics has proven to be a powerful tool for studying the functions
837     of biological systems, providing structural, thermodynamic and
838     dynamical information. The basic idea of molecular dynamics is that
839     macroscopic properties are related to microscopic behavior and
840     microscopic behavior can be calculated from the trajectories in
841 tim 2907 simulations. For instance, instantaneous temperature of a
842     Hamiltonian system of $N$ particles can be measured by
843 tim 2720 \[
844 tim 2725 T = \sum\limits_{i = 1}^N {\frac{{m_i v_i^2 }}{{fk_B }}}
845 tim 2720 \]
846     where $m_i$ and $v_i$ are the mass and velocity of $i$th particle
847     respectively, $f$ is the number of degrees of freedom, and $k_B$ is
848 tim 2907 the Boltzman constant.
849 tim 2694
850 tim 2720 A typical molecular dynamics run consists of three essential steps:
851     \begin{enumerate}
852     \item Initialization
853     \begin{enumerate}
854     \item Preliminary preparation
855     \item Minimization
856     \item Heating
857     \item Equilibration
858     \end{enumerate}
859     \item Production
860     \item Analysis
861     \end{enumerate}
862     These three individual steps will be covered in the following
863     sections. Sec.~\ref{introSec:initialSystemSettings} deals with the
864 tim 2801 initialization of a simulation. Sec.~\ref{introSection:production}
865 tim 2907 will discuss issues of production runs.
866 tim 2801 Sec.~\ref{introSection:Analysis} provides the theoretical tools for
867 tim 2907 analysis of trajectories.
868 tim 2719
869 tim 2720 \subsection{\label{introSec:initialSystemSettings}Initialization}
870 tim 2719
871 tim 2819 \subsubsection{\textbf{Preliminary preparation}}
872 tim 2719
873 tim 2720 When selecting the starting structure of a molecule for molecular
874     simulation, one may retrieve its Cartesian coordinates from public
875     databases, such as RCSB Protein Data Bank \textit{etc}. Although
876     thousands of crystal structures of molecules are discovered every
877     year, many more remain unknown due to the difficulties of
878 tim 2872 purification and crystallization. Even for molecules with known
879 tim 2907 structures, some important information is missing. For example, a
880 tim 2720 missing hydrogen atom which acts as donor in hydrogen bonding must
881 tim 2907 be added. Moreover, in order to include electrostatic interactions,
882 tim 2720 one may need to specify the partial charges for individual atoms.
883     Under some circumstances, we may even need to prepare the system in
884 tim 2872 a special configuration. For instance, when studying transport
885     phenomenon in membrane systems, we may prepare the lipids in a
886     bilayer structure instead of placing lipids randomly in solvent,
887     since we are not interested in the slow self-aggregation process.
888 tim 2694
889 tim 2819 \subsubsection{\textbf{Minimization}}
890 tim 2705
891 tim 2720 It is quite possible that some of molecules in the system from
892 tim 2872 preliminary preparation may be overlapping with each other. This
893     close proximity leads to high initial potential energy which
894     consequently jeopardizes any molecular dynamics simulations. To
895     remove these steric overlaps, one typically performs energy
896     minimization to find a more reasonable conformation. Several energy
897     minimization methods have been developed to exploit the energy
898     surface and to locate the local minimum. While converging slowly
899     near the minimum, steepest descent method is extremely robust when
900     systems are strongly anharmonic. Thus, it is often used to refine
901 tim 2907 structures from crystallographic data. Relying on the Hessian,
902     advanced methods like Newton-Raphson converge rapidly to a local
903     minimum, but become unstable if the energy surface is far from
904 tim 2872 quadratic. Another factor that must be taken into account, when
905 tim 2720 choosing energy minimization method, is the size of the system.
906     Steepest descent and conjugate gradient can deal with models of any
907 tim 2872 size. Because of the limits on computer memory to store the hessian
908 tim 2907 matrix and the computing power needed to diagonalize these matrices,
909     most Newton-Raphson methods can not be used with very large systems.
910 tim 2694
911 tim 2819 \subsubsection{\textbf{Heating}}
912 tim 2720
913 tim 2907 Typically, heating is performed by assigning random velocities
914 tim 2872 according to a Maxwell-Boltzman distribution for a desired
915     temperature. Beginning at a lower temperature and gradually
916     increasing the temperature by assigning larger random velocities, we
917 tim 2907 end up setting the temperature of the system to a final temperature
918     at which the simulation will be conducted. In heating phase, we
919     should also keep the system from drifting or rotating as a whole. To
920     do this, the net linear momentum and angular momentum of the system
921     is shifted to zero after each resampling from the Maxwell -Boltzman
922     distribution.
923 tim 2720
924 tim 2819 \subsubsection{\textbf{Equilibration}}
925 tim 2720
926     The purpose of equilibration is to allow the system to evolve
927     spontaneously for a period of time and reach equilibrium. The
928     procedure is continued until various statistical properties, such as
929     temperature, pressure, energy, volume and other structural
930     properties \textit{etc}, become independent of time. Strictly
931     speaking, minimization and heating are not necessary, provided the
932     equilibration process is long enough. However, these steps can serve
933     as a means to arrive at an equilibrated structure in an effective
934     way.
935    
936     \subsection{\label{introSection:production}Production}
937    
938 tim 2872 The production run is the most important step of the simulation, in
939 tim 2725 which the equilibrated structure is used as a starting point and the
940     motions of the molecules are collected for later analysis. In order
941     to capture the macroscopic properties of the system, the molecular
942 tim 2872 dynamics simulation must be performed by sampling correctly and
943     efficiently from the relevant thermodynamic ensemble.
944 tim 2720
945 tim 2725 The most expensive part of a molecular dynamics simulation is the
946     calculation of non-bonded forces, such as van der Waals force and
947     Coulombic forces \textit{etc}. For a system of $N$ particles, the
948     complexity of the algorithm for pair-wise interactions is $O(N^2 )$,
949 tim 2907 which makes large simulations prohibitive in the absence of any
950 tim 2905 algorithmic tricks. A natural approach to avoid system size issues
951     is to represent the bulk behavior by a finite number of the
952 tim 2907 particles. However, this approach will suffer from surface effects
953     at the edges of the simulation. To offset this, \textit{Periodic
954     boundary conditions} (see Fig.~\ref{introFig:pbc}) were developed to
955     simulate bulk properties with a relatively small number of
956     particles. In this method, the simulation box is replicated
957     throughout space to form an infinite lattice. During the simulation,
958     when a particle moves in the primary cell, its image in other cells
959     move in exactly the same direction with exactly the same
960 tim 2905 orientation. Thus, as a particle leaves the primary cell, one of its
961     images will enter through the opposite face.
962 tim 2789 \begin{figure}
963     \centering
964     \includegraphics[width=\linewidth]{pbc.eps}
965     \caption[An illustration of periodic boundary conditions]{A 2-D
966     illustration of periodic boundary conditions. As one particle leaves
967     the left of the simulation box, an image of it enters the right.}
968     \label{introFig:pbc}
969     \end{figure}
970 tim 2725
971     %cutoff and minimum image convention
972     Another important technique to improve the efficiency of force
973 tim 2907 evaluation is to apply spherical cutoffs where particles farther
974     than a predetermined distance are not included in the calculation
975 tim 2725 \cite{Frenkel1996}. The use of a cutoff radius will cause a
976 tim 2730 discontinuity in the potential energy curve. Fortunately, one can
977 tim 2907 shift a simple radial potential to ensure the potential curve go
978 tim 2872 smoothly to zero at the cutoff radius. The cutoff strategy works
979     well for Lennard-Jones interaction because of its short range
980     nature. However, simply truncating the electrostatic interaction
981     with the use of cutoffs has been shown to lead to severe artifacts
982     in simulations. The Ewald summation, in which the slowly decaying
983     Coulomb potential is transformed into direct and reciprocal sums
984     with rapid and absolute convergence, has proved to minimize the
985     periodicity artifacts in liquid simulations. Taking the advantages
986     of the fast Fourier transform (FFT) for calculating discrete Fourier
987     transforms, the particle mesh-based
988 tim 2789 methods\cite{Hockney1981,Shimada1993, Luty1994} are accelerated from
989 tim 2872 $O(N^{3/2})$ to $O(N logN)$. An alternative approach is the
990     \emph{fast multipole method}\cite{Greengard1987, Greengard1994},
991     which treats Coulombic interactions exactly at short range, and
992     approximate the potential at long range through multipolar
993     expansion. In spite of their wide acceptance at the molecular
994     simulation community, these two methods are difficult to implement
995     correctly and efficiently. Instead, we use a damped and
996     charge-neutralized Coulomb potential method developed by Wolf and
997     his coworkers\cite{Wolf1999}. The shifted Coulomb potential for
998     particle $i$ and particle $j$ at distance $r_{rj}$ is given by:
999 tim 2725 \begin{equation}
1000     V(r_{ij})= \frac{q_i q_j \textrm{erfc}(\alpha
1001     r_{ij})}{r_{ij}}-\lim_{r_{ij}\rightarrow
1002     R_\textrm{c}}\left\{\frac{q_iq_j \textrm{erfc}(\alpha
1003     r_{ij})}{r_{ij}}\right\}. \label{introEquation:shiftedCoulomb}
1004     \end{equation}
1005     where $\alpha$ is the convergence parameter. Due to the lack of
1006     inherent periodicity and rapid convergence,this method is extremely
1007     efficient and easy to implement.
1008 tim 2789 \begin{figure}
1009     \centering
1010     \includegraphics[width=\linewidth]{shifted_coulomb.eps}
1011     \caption[An illustration of shifted Coulomb potential]{An
1012     illustration of shifted Coulomb potential.}
1013     \label{introFigure:shiftedCoulomb}
1014     \end{figure}
1015 tim 2725
1016     %multiple time step
1017    
1018 tim 2720 \subsection{\label{introSection:Analysis} Analysis}
1019    
1020 tim 2872 Recently, advanced visualization technique have become applied to
1021 tim 2721 monitor the motions of molecules. Although the dynamics of the
1022     system can be described qualitatively from animation, quantitative
1023 tim 2907 trajectory analysis is more useful. According to the principles of
1024 tim 2905 Statistical Mechanics in
1025     Sec.~\ref{introSection:statisticalMechanics}, one can compute
1026     thermodynamic properties, analyze fluctuations of structural
1027     parameters, and investigate time-dependent processes of the molecule
1028     from the trajectories.
1029 tim 2721
1030 tim 2872 \subsubsection{\label{introSection:thermodynamicsProperties}\textbf{Thermodynamic Properties}}
1031 tim 2721
1032 tim 2872 Thermodynamic properties, which can be expressed in terms of some
1033 tim 2725 function of the coordinates and momenta of all particles in the
1034     system, can be directly computed from molecular dynamics. The usual
1035     way to measure the pressure is based on virial theorem of Clausius
1036     which states that the virial is equal to $-3Nk_BT$. For a system
1037     with forces between particles, the total virial, $W$, contains the
1038     contribution from external pressure and interaction between the
1039     particles:
1040     \[
1041     W = - 3PV + \left\langle {\sum\limits_{i < j} {r{}_{ij} \cdot
1042     f_{ij} } } \right\rangle
1043     \]
1044     where $f_{ij}$ is the force between particle $i$ and $j$ at a
1045     distance $r_{ij}$. Thus, the expression for the pressure is given
1046     by:
1047     \begin{equation}
1048     P = \frac{{Nk_B T}}{V} - \frac{1}{{3V}}\left\langle {\sum\limits_{i
1049     < j} {r{}_{ij} \cdot f_{ij} } } \right\rangle
1050     \end{equation}
1051    
1052 tim 2819 \subsubsection{\label{introSection:structuralProperties}\textbf{Structural Properties}}
1053 tim 2721
1054     Structural Properties of a simple fluid can be described by a set of
1055 tim 2872 distribution functions. Among these functions,the \emph{pair
1056 tim 2721 distribution function}, also known as \emph{radial distribution
1057 tim 2872 function}, is of most fundamental importance to liquid theory.
1058 tim 2907 Experimentally, pair distribution functions can be gathered by
1059 tim 2872 Fourier transforming raw data from a series of neutron diffraction
1060     experiments and integrating over the surface factor
1061     \cite{Powles1973}. The experimental results can serve as a criterion
1062     to justify the correctness of a liquid model. Moreover, various
1063     equilibrium thermodynamic and structural properties can also be
1064 tim 2907 expressed in terms of the radial distribution function
1065     \cite{Allen1987}. The pair distribution functions $g(r)$ gives the
1066     probability that a particle $i$ will be located at a distance $r$
1067     from a another particle $j$ in the system
1068 tim 2905 \begin{equation}
1069 tim 2721 g(r) = \frac{V}{{N^2 }}\left\langle {\sum\limits_i {\sum\limits_{j
1070 tim 2874 \ne i} {\delta (r - r_{ij} )} } } \right\rangle = \frac{\rho
1071 tim 2872 (r)}{\rho}.
1072 tim 2905 \end{equation}
1073 tim 2721 Note that the delta function can be replaced by a histogram in
1074 tim 2881 computer simulation. Peaks in $g(r)$ represent solvent shells, and
1075     the height of these peaks gradually decreases to 1 as the liquid of
1076     large distance approaches the bulk density.
1077 tim 2721
1078    
1079 tim 2819 \subsubsection{\label{introSection:timeDependentProperties}\textbf{Time-dependent
1080     Properties}}
1081 tim 2721
1082     Time-dependent properties are usually calculated using \emph{time
1083 tim 2872 correlation functions}, which correlate random variables $A$ and $B$
1084     at two different times,
1085 tim 2721 \begin{equation}
1086     C_{AB} (t) = \left\langle {A(t)B(0)} \right\rangle.
1087     \label{introEquation:timeCorrelationFunction}
1088     \end{equation}
1089     If $A$ and $B$ refer to same variable, this kind of correlation
1090 tim 2872 function is called an \emph{autocorrelation function}. One example
1091     of an auto correlation function is the velocity auto-correlation
1092     function which is directly related to transport properties of
1093     molecular liquids:
1094 tim 2725 \[
1095     D = \frac{1}{3}\int\limits_0^\infty {\left\langle {v(t) \cdot v(0)}
1096     \right\rangle } dt
1097     \]
1098 tim 2872 where $D$ is diffusion constant. Unlike the velocity autocorrelation
1099 tim 2907 function, which is averaged over time origins and over all the
1100     atoms, the dipole autocorrelation functions is calculated for the
1101 tim 2872 entire system. The dipole autocorrelation function is given by:
1102 tim 2725 \[
1103     c_{dipole} = \left\langle {u_{tot} (t) \cdot u_{tot} (t)}
1104     \right\rangle
1105     \]
1106     Here $u_{tot}$ is the net dipole of the entire system and is given
1107     by
1108     \[
1109 tim 2905 u_{tot} (t) = \sum\limits_i {u_i (t)}.
1110 tim 2725 \]
1111 tim 2907 In principle, many time correlation functions can be related to
1112 tim 2725 Fourier transforms of the infrared, Raman, and inelastic neutron
1113     scattering spectra of molecular liquids. In practice, one can
1114 tim 2907 extract the IR spectrum from the intensity of the molecular dipole
1115     fluctuation at each frequency using the following relationship:
1116 tim 2725 \[
1117     \hat c_{dipole} (v) = \int_{ - \infty }^\infty {c_{dipole} (t)e^{ -
1118 tim 2905 i2\pi vt} dt}.
1119 tim 2725 \]
1120 tim 2721
1121 tim 2693 \section{\label{introSection:rigidBody}Dynamics of Rigid Bodies}
1122 tim 2692
1123 tim 2705 Rigid bodies are frequently involved in the modeling of different
1124     areas, from engineering, physics, to chemistry. For example,
1125 tim 2907 missiles and vehicles are usually modeled by rigid bodies. The
1126     movement of the objects in 3D gaming engines or other physics
1127     simulators is governed by rigid body dynamics. In molecular
1128 tim 2872 simulations, rigid bodies are used to simplify protein-protein
1129     docking studies\cite{Gray2003}.
1130 tim 2694
1131 tim 2705 It is very important to develop stable and efficient methods to
1132 tim 2872 integrate the equations of motion for orientational degrees of
1133     freedom. Euler angles are the natural choice to describe the
1134     rotational degrees of freedom. However, due to $\frac {1}{sin
1135     \theta}$ singularities, the numerical integration of corresponding
1136 tim 2907 equations of these motion is very inefficient and inaccurate.
1137     Although an alternative integrator using multiple sets of Euler
1138     angles can overcome this difficulty\cite{Barojas1973}, the
1139     computational penalty and the loss of angular momentum conservation
1140     still remain. A singularity-free representation utilizing
1141     quaternions was developed by Evans in 1977\cite{Evans1977}.
1142     Unfortunately, this approach uses a nonseparable Hamiltonian
1143     resulting from the quaternion representation, which prevents the
1144     symplectic algorithm from being utilized. Another different approach
1145     is to apply holonomic constraints to the atoms belonging to the
1146     rigid body. Each atom moves independently under the normal forces
1147     deriving from potential energy and constraint forces which are used
1148     to guarantee the rigidness. However, due to their iterative nature,
1149     the SHAKE and Rattle algorithms also converge very slowly when the
1150     number of constraints increases\cite{Ryckaert1977, Andersen1983}.
1151 tim 2694
1152 tim 2872 A break-through in geometric literature suggests that, in order to
1153 tim 2705 develop a long-term integration scheme, one should preserve the
1154 tim 2907 symplectic structure of the propagator. By introducing a conjugate
1155 tim 2872 momentum to the rotation matrix $Q$ and re-formulating Hamiltonian's
1156     equation, a symplectic integrator, RSHAKE\cite{Kol1997}, was
1157     proposed to evolve the Hamiltonian system in a constraint manifold
1158     by iteratively satisfying the orthogonality constraint $Q^T Q = 1$.
1159     An alternative method using the quaternion representation was
1160     developed by Omelyan\cite{Omelyan1998}. However, both of these
1161     methods are iterative and inefficient. In this section, we descibe a
1162 tim 2907 symplectic Lie-Poisson integrator for rigid bodies developed by
1163 tim 2789 Dullweber and his coworkers\cite{Dullweber1997} in depth.
1164 tim 2705
1165 tim 2872 \subsection{\label{introSection:constrainedHamiltonianRB}Constrained Hamiltonian for Rigid Bodies}
1166     The motion of a rigid body is Hamiltonian with the Hamiltonian
1167 tim 2713 function
1168 tim 2706 \begin{equation}
1169     H = \frac{1}{2}(p^T m^{ - 1} p) + \frac{1}{2}tr(PJ^{ - 1} P) +
1170     V(q,Q) + \frac{1}{2}tr[(QQ^T - 1)\Lambda ].
1171     \label{introEquation:RBHamiltonian}
1172     \end{equation}
1173 tim 2909 Here, $q$ and $Q$ are the position vector and rotation matrix for
1174     the rigid-body, $p$ and $P$ are conjugate momenta to $q$ and $Q$ ,
1175     and $J$, a diagonal matrix, is defined by
1176 tim 2706 \[
1177     I_{ii}^{ - 1} = \frac{1}{2}\sum\limits_{i \ne j} {J_{jj}^{ - 1} }
1178     \]
1179     where $I_{ii}$ is the diagonal element of the inertia tensor. This
1180 tim 2872 constrained Hamiltonian equation is subjected to a holonomic
1181     constraint,
1182 tim 2706 \begin{equation}
1183 tim 2726 Q^T Q = 1, \label{introEquation:orthogonalConstraint}
1184 tim 2706 \end{equation}
1185 tim 2909 which is used to ensure the rotation matrix's unitarity. Using
1186     Equation (\ref{introEquation:motionHamiltonianCoordinate},
1187 tim 2706 \ref{introEquation:motionHamiltonianMomentum}), one can write down
1188     the equations of motion,
1189 tim 2796 \begin{eqnarray}
1190 tim 2905 \frac{{dq}}{{dt}} & = & \frac{p}{m}, \label{introEquation:RBMotionPosition}\\
1191     \frac{{dp}}{{dt}} & = & - \nabla _q V(q,Q), \label{introEquation:RBMotionMomentum}\\
1192     \frac{{dQ}}{{dt}} & = & PJ^{ - 1}, \label{introEquation:RBMotionRotation}\\
1193 tim 2796 \frac{{dP}}{{dt}} & = & - \nabla _Q V(q,Q) - 2Q\Lambda . \label{introEquation:RBMotionP}
1194     \end{eqnarray}
1195 tim 2909 Differentiating Eq.~\ref{introEquation:orthogonalConstraint} and
1196     using Eq.~\ref{introEquation:RBMotionMomentum}, one may obtain,
1197     \begin{equation}
1198     Q^T PJ^{ - 1} + J^{ - 1} P^T Q = 0 . \\
1199     \label{introEquation:RBFirstOrderConstraint}
1200     \end{equation}
1201 tim 2707 In general, there are two ways to satisfy the holonomic constraints.
1202 tim 2872 We can use a constraint force provided by a Lagrange multiplier on
1203 tim 2909 the normal manifold to keep the motion on the constraint space. Or
1204     we can simply evolve the system on the constraint manifold. These
1205     two methods have been proved to be equivalent. The holonomic
1206     constraint and equations of motions define a constraint manifold for
1207     rigid bodies
1208 tim 2707 \[
1209     M = \left\{ {(Q,P):Q^T Q = 1,Q^T PJ^{ - 1} + J^{ - 1} P^T Q = 0}
1210     \right\}.
1211     \]
1212 tim 2909 Unfortunately, this constraint manifold is not $T^* SO(3)$ which is
1213     a symplectic manifold on Lie rotation group $SO(3)$. However, it
1214     turns out that under symplectic transformation, the cotangent space
1215     and the phase space are diffeomorphic. By introducing
1216 tim 2706 \[
1217 tim 2707 \tilde Q = Q,\tilde P = \frac{1}{2}\left( {P - QP^T Q} \right),
1218 tim 2706 \]
1219 tim 2707 the mechanical system subject to a holonomic constraint manifold $M$
1220     can be re-formulated as a Hamiltonian system on the cotangent space
1221     \[
1222     T^* SO(3) = \left\{ {(\tilde Q,\tilde P):\tilde Q^T \tilde Q =
1223     1,\tilde Q^T \tilde PJ^{ - 1} + J^{ - 1} P^T \tilde Q = 0} \right\}
1224     \]
1225     For a body fixed vector $X_i$ with respect to the center of mass of
1226     the rigid body, its corresponding lab fixed vector $X_0^{lab}$ is
1227     given as
1228     \begin{equation}
1229     X_i^{lab} = Q X_i + q.
1230     \end{equation}
1231     Therefore, potential energy $V(q,Q)$ is defined by
1232     \[
1233     V(q,Q) = V(Q X_0 + q).
1234     \]
1235 tim 2713 Hence, the force and torque are given by
1236 tim 2707 \[
1237 tim 2713 \nabla _q V(q,Q) = F(q,Q) = \sum\limits_i {F_i (q,Q)},
1238 tim 2707 \]
1239 tim 2713 and
1240 tim 2707 \[
1241     \nabla _Q V(q,Q) = F(q,Q)X_i^t
1242     \]
1243 tim 2899 respectively. As a common choice to describe the rotation dynamics
1244     of the rigid body, the angular momentum on the body fixed frame $\Pi
1245     = Q^t P$ is introduced to rewrite the equations of motion,
1246 tim 2707 \begin{equation}
1247     \begin{array}{l}
1248 tim 2899 \dot \Pi = J^{ - 1} \Pi ^T \Pi + Q^T \sum\limits_i {F_i (q,Q)X_i^T } - \Lambda, \\
1249     \dot Q = Q\Pi {\rm{ }}J^{ - 1}, \\
1250 tim 2707 \end{array}
1251     \label{introEqaution:RBMotionPI}
1252     \end{equation}
1253 tim 2905 as well as holonomic constraints $\Pi J^{ - 1} + J^{ - 1} \Pi ^t =
1254     0$ and $Q^T Q = 1$. For a vector $v(v_1 ,v_2 ,v_3 ) \in R^3$ and a
1255     matrix $\hat v \in so(3)^ \star$, the hat-map isomorphism,
1256 tim 2707 \begin{equation}
1257     v(v_1 ,v_2 ,v_3 ) \Leftrightarrow \hat v = \left(
1258     {\begin{array}{*{20}c}
1259     0 & { - v_3 } & {v_2 } \\
1260     {v_3 } & 0 & { - v_1 } \\
1261     { - v_2 } & {v_1 } & 0 \\
1262     \end{array}} \right),
1263     \label{introEquation:hatmapIsomorphism}
1264     \end{equation}
1265     will let us associate the matrix products with traditional vector
1266     operations
1267     \[
1268 tim 2899 \hat vu = v \times u.
1269 tim 2707 \]
1270 tim 2899 Using Eq.~\ref{introEqaution:RBMotionPI}, one can construct a skew
1271 tim 2707 matrix,
1272 tim 2899 \begin{eqnarray}
1273 tim 2905 (\dot \Pi - \dot \Pi ^T )&= &(\Pi - \Pi ^T )(J^{ - 1} \Pi + \Pi J^{ - 1} ) \notag \\
1274     & & + \sum\limits_i {[Q^T F_i (r,Q)X_i^T - X_i F_i (r,Q)^T Q]} -
1275     (\Lambda - \Lambda ^T ). \label{introEquation:skewMatrixPI}
1276 tim 2899 \end{eqnarray}
1277     Since $\Lambda$ is symmetric, the last term of
1278     Eq.~\ref{introEquation:skewMatrixPI} is zero, which implies the
1279     Lagrange multiplier $\Lambda$ is absent from the equations of
1280     motion. This unique property eliminates the requirement of
1281     iterations which can not be avoided in other methods\cite{Kol1997,
1282     Omelyan1998}. Applying the hat-map isomorphism, we obtain the
1283 tim 2909 equation of motion for angular momentum in the body frame
1284 tim 2713 \begin{equation}
1285     \dot \pi = \pi \times I^{ - 1} \pi + \sum\limits_i {\left( {Q^T
1286     F_i (r,Q)} \right) \times X_i }.
1287     \label{introEquation:bodyAngularMotion}
1288     \end{equation}
1289 tim 2707 In the same manner, the equation of motion for rotation matrix is
1290     given by
1291     \[
1292 tim 2899 \dot Q = Qskew(I^{ - 1} \pi ).
1293 tim 2707 \]
1294    
1295 tim 2713 \subsection{\label{introSection:SymplecticFreeRB}Symplectic
1296 tim 2909 Lie-Poisson Integrator for Free Rigid Bodies}
1297 tim 2707
1298 tim 2872 If there are no external forces exerted on the rigid body, the only
1299     contribution to the rotational motion is from the kinetic energy
1300     (the first term of \ref{introEquation:bodyAngularMotion}). The free
1301     rigid body is an example of a Lie-Poisson system with Hamiltonian
1302     function
1303 tim 2713 \begin{equation}
1304     T^r (\pi ) = T_1 ^r (\pi _1 ) + T_2^r (\pi _2 ) + T_3^r (\pi _3 )
1305     \label{introEquation:rotationalKineticRB}
1306     \end{equation}
1307     where $T_i^r (\pi _i ) = \frac{{\pi _i ^2 }}{{2I_i }}$ and
1308     Lie-Poisson structure matrix,
1309     \begin{equation}
1310     J(\pi ) = \left( {\begin{array}{*{20}c}
1311     0 & {\pi _3 } & { - \pi _2 } \\
1312     { - \pi _3 } & 0 & {\pi _1 } \\
1313     {\pi _2 } & { - \pi _1 } & 0 \\
1314 tim 2899 \end{array}} \right).
1315 tim 2713 \end{equation}
1316     Thus, the dynamics of free rigid body is governed by
1317     \begin{equation}
1318 tim 2899 \frac{d}{{dt}}\pi = J(\pi )\nabla _\pi T^r (\pi ).
1319 tim 2713 \end{equation}
1320 tim 2905 One may notice that each $T_i^r$ in
1321     Eq.~\ref{introEquation:rotationalKineticRB} can be solved exactly.
1322     For instance, the equations of motion due to $T_1^r$ are given by
1323 tim 2713 \begin{equation}
1324     \frac{d}{{dt}}\pi = R_1 \pi ,\frac{d}{{dt}}Q = QR_1
1325     \label{introEqaution:RBMotionSingleTerm}
1326     \end{equation}
1327 tim 2905 with
1328 tim 2713 \[ R_1 = \left( {\begin{array}{*{20}c}
1329     0 & 0 & 0 \\
1330     0 & 0 & {\pi _1 } \\
1331     0 & { - \pi _1 } & 0 \\
1332     \end{array}} \right).
1333     \]
1334 tim 2905 The solutions of Eq.~\ref{introEqaution:RBMotionSingleTerm} is
1335 tim 2707 \[
1336 tim 2713 \pi (\Delta t) = e^{\Delta tR_1 } \pi (0),Q(\Delta t) =
1337     Q(0)e^{\Delta tR_1 }
1338 tim 2707 \]
1339 tim 2713 with
1340 tim 2707 \[
1341 tim 2713 e^{\Delta tR_1 } = \left( {\begin{array}{*{20}c}
1342     0 & 0 & 0 \\
1343     0 & {\cos \theta _1 } & {\sin \theta _1 } \\
1344     0 & { - \sin \theta _1 } & {\cos \theta _1 } \\
1345     \end{array}} \right),\theta _1 = \frac{{\pi _1 }}{{I_1 }}\Delta t.
1346 tim 2707 \]
1347 tim 2719 To reduce the cost of computing expensive functions in $e^{\Delta
1348 tim 2909 tR_1 }$, we can use the Cayley transformation to obtain a
1349     single-aixs propagator,
1350     \begin{eqnarray*}
1351     e^{\Delta tR_1 } & \approx & (1 - \Delta tR_1 )^{ - 1} (1 + \Delta
1352     tR_1 ) \\
1353     %
1354     & \approx & \left( \begin{array}{ccc}
1355     1 & 0 & 0 \\
1356     0 & \frac{1-\theta^2 / 4}{1 + \theta^2 / 4} & -\frac{\theta}{1+
1357     \theta^2 / 4} \\
1358     0 & \frac{\theta}{1+ \theta^2 / 4} & \frac{1-\theta^2 / 4}{1 +
1359     \theta^2 / 4}
1360     \end{array}
1361     \right).
1362     \end{eqnarray*}
1363     The propagators for $T_2^r$ and $T_3^r$ can be found in the same
1364 tim 2872 manner. In order to construct a second-order symplectic method, we
1365 tim 2905 split the angular kinetic Hamiltonian function into five terms
1366 tim 2707 \[
1367 tim 2713 T^r (\pi ) = \frac{1}{2}T_1 ^r (\pi _1 ) + \frac{1}{2}T_2^r (\pi _2
1368     ) + T_3^r (\pi _3 ) + \frac{1}{2}T_2^r (\pi _2 ) + \frac{1}{2}T_1 ^r
1369 tim 2872 (\pi _1 ).
1370     \]
1371     By concatenating the propagators corresponding to these five terms,
1372     we can obtain an symplectic integrator,
1373 tim 2713 \[
1374     \varphi _{\Delta t,T^r } = \varphi _{\Delta t/2,\pi _1 } \circ
1375 tim 2707 \varphi _{\Delta t/2,\pi _2 } \circ \varphi _{\Delta t,\pi _3 }
1376     \circ \varphi _{\Delta t/2,\pi _2 } \circ \varphi _{\Delta t/2,\pi
1377 tim 2713 _1 }.
1378 tim 2707 \]
1379 tim 2713 The non-canonical Lie-Poisson bracket ${F, G}$ of two function
1380     $F(\pi )$ and $G(\pi )$ is defined by
1381 tim 2707 \[
1382 tim 2713 \{ F,G\} (\pi ) = [\nabla _\pi F(\pi )]^T J(\pi )\nabla _\pi G(\pi
1383 tim 2899 ).
1384 tim 2713 \]
1385     If the Poisson bracket of a function $F$ with an arbitrary smooth
1386     function $G$ is zero, $F$ is a \emph{Casimir}, which is the
1387     conserved quantity in Poisson system. We can easily verify that the
1388     norm of the angular momentum, $\parallel \pi
1389 tim 2909 \parallel$, is a \emph{Casimir}\cite{McLachlan1993}. Let$ F(\pi ) = S(\frac{{\parallel
1390 tim 2713 \pi \parallel ^2 }}{2})$ for an arbitrary function $ S:R \to R$ ,
1391     then by the chain rule
1392     \[
1393     \nabla _\pi F(\pi ) = S'(\frac{{\parallel \pi \parallel ^2
1394 tim 2899 }}{2})\pi.
1395 tim 2713 \]
1396 tim 2899 Thus, $ [\nabla _\pi F(\pi )]^T J(\pi ) = - S'(\frac{{\parallel
1397     \pi
1398 tim 2713 \parallel ^2 }}{2})\pi \times \pi = 0 $. This explicit
1399 tim 2872 Lie-Poisson integrator is found to be both extremely efficient and
1400     stable. These properties can be explained by the fact the small
1401     angle approximation is used and the norm of the angular momentum is
1402     conserved.
1403 tim 2713
1404     \subsection{\label{introSection:RBHamiltonianSplitting} Hamiltonian
1405     Splitting for Rigid Body}
1406    
1407     The Hamiltonian of rigid body can be separated in terms of kinetic
1408 tim 2905 energy and potential energy,$H = T(p,\pi ) + V(q,Q)$. The equations
1409     of motion corresponding to potential energy and kinetic energy are
1410     listed in the below table,
1411 tim 2776 \begin{table}
1412 tim 2889 \caption{EQUATIONS OF MOTION DUE TO POTENTIAL AND KINETIC ENERGIES}
1413 tim 2713 \begin{center}
1414     \begin{tabular}{|l|l|}
1415     \hline
1416     % after \\: \hline or \cline{col1-col2} \cline{col3-col4} ...
1417     Potential & Kinetic \\
1418     $\frac{{dq}}{{dt}} = \frac{p}{m}$ & $\frac{d}{{dt}}q = p$ \\
1419     $\frac{d}{{dt}}p = - \frac{{\partial V}}{{\partial q}}$ & $ \frac{d}{{dt}}p = 0$ \\
1420     $\frac{d}{{dt}}Q = 0$ & $ \frac{d}{{dt}}Q = Qskew(I^{ - 1} j)$ \\
1421     $ \frac{d}{{dt}}\pi = \sum\limits_i {\left( {Q^T F_i (r,Q)} \right) \times X_i }$ & $\frac{d}{{dt}}\pi = \pi \times I^{ - 1} \pi$\\
1422     \hline
1423     \end{tabular}
1424     \end{center}
1425 tim 2776 \end{table}
1426 tim 2872 A second-order symplectic method is now obtained by the composition
1427     of the position and velocity propagators,
1428 tim 2713 \[
1429     \varphi _{\Delta t} = \varphi _{\Delta t/2,V} \circ \varphi
1430     _{\Delta t,T} \circ \varphi _{\Delta t/2,V}.
1431     \]
1432 tim 2719 Moreover, $\varphi _{\Delta t/2,V}$ can be divided into two
1433 tim 2872 sub-propagators which corresponding to force and torque
1434     respectively,
1435 tim 2713 \[
1436 tim 2707 \varphi _{\Delta t/2,V} = \varphi _{\Delta t/2,F} \circ \varphi
1437 tim 2713 _{\Delta t/2,\tau }.
1438 tim 2707 \]
1439 tim 2713 Since the associated operators of $\varphi _{\Delta t/2,F} $ and
1440 tim 2872 $\circ \varphi _{\Delta t/2,\tau }$ commute, the composition order
1441     inside $\varphi _{\Delta t/2,V}$ does not matter. Furthermore, the
1442     kinetic energy can be separated to translational kinetic term, $T^t
1443     (p)$, and rotational kinetic term, $T^r (\pi )$,
1444 tim 2713 \begin{equation}
1445     T(p,\pi ) =T^t (p) + T^r (\pi ).
1446     \end{equation}
1447     where $ T^t (p) = \frac{1}{2}p^T m^{ - 1} p $ and $T^r (\pi )$ is
1448 tim 2905 defined by Eq.~\ref{introEquation:rotationalKineticRB}. Therefore,
1449     the corresponding propagators are given by
1450 tim 2713 \[
1451     \varphi _{\Delta t,T} = \varphi _{\Delta t,T^t } \circ \varphi
1452     _{\Delta t,T^r }.
1453     \]
1454 tim 2872 Finally, we obtain the overall symplectic propagators for freely
1455     moving rigid bodies
1456 tim 2905 \begin{eqnarray}
1457     \varphi _{\Delta t} &=& \varphi _{\Delta t/2,F} \circ \varphi _{\Delta t/2,\tau } \notag\\
1458     & & \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 } \notag\\
1459 tim 2908 & & \circ \varphi _{\Delta t/2,\tau } \circ \varphi _{\Delta t/2,F} .
1460 tim 2713 \label{introEquation:overallRBFlowMaps}
1461 tim 2905 \end{eqnarray}
1462 tim 2707
1463 tim 2685 \section{\label{introSection:langevinDynamics}Langevin Dynamics}
1464 tim 2716 As an alternative to newtonian dynamics, Langevin dynamics, which
1465     mimics a simple heat bath with stochastic and dissipative forces,
1466     has been applied in a variety of studies. This section will review
1467 tim 2872 the theory of Langevin dynamics. A brief derivation of generalized
1468     Langevin equation will be given first. Following that, we will
1469     discuss the physical meaning of the terms appearing in the equation
1470     as well as the calculation of friction tensor from hydrodynamics
1471     theory.
1472 tim 2685
1473 tim 2719 \subsection{\label{introSection:generalizedLangevinDynamics}Derivation of Generalized Langevin Equation}
1474 tim 2685
1475 tim 2872 A harmonic bath model, in which an effective set of harmonic
1476 tim 2719 oscillators are used to mimic the effect of a linearly responding
1477     environment, has been widely used in quantum chemistry and
1478     statistical mechanics. One of the successful applications of
1479 tim 2872 Harmonic bath model is the derivation of the Generalized Langevin
1480     Dynamics (GLE). Lets consider a system, in which the degree of
1481 tim 2719 freedom $x$ is assumed to couple to the bath linearly, giving a
1482     Hamiltonian of the form
1483 tim 2696 \begin{equation}
1484     H = \frac{{p^2 }}{{2m}} + U(x) + H_B + \Delta U(x,x_1 , \ldots x_N)
1485 tim 2719 \label{introEquation:bathGLE}.
1486 tim 2696 \end{equation}
1487 tim 2872 Here $p$ is a momentum conjugate to $x$, $m$ is the mass associated
1488     with this degree of freedom, $H_B$ is a harmonic bath Hamiltonian,
1489 tim 2696 \[
1490 tim 2719 H_B = \sum\limits_{\alpha = 1}^N {\left\{ {\frac{{p_\alpha ^2
1491     }}{{2m_\alpha }} + \frac{1}{2}m_\alpha \omega _\alpha ^2 }
1492     \right\}}
1493 tim 2696 \]
1494 tim 2719 where the index $\alpha$ runs over all the bath degrees of freedom,
1495     $\omega _\alpha$ are the harmonic bath frequencies, $m_\alpha$ are
1496 tim 2872 the harmonic bath masses, and $\Delta U$ is a bilinear system-bath
1497 tim 2719 coupling,
1498 tim 2696 \[
1499     \Delta U = - \sum\limits_{\alpha = 1}^N {g_\alpha x_\alpha x}
1500     \]
1501 tim 2872 where $g_\alpha$ are the coupling constants between the bath
1502 tim 2874 coordinates ($x_ \alpha$) and the system coordinate ($x$).
1503 tim 2872 Introducing
1504 tim 2696 \[
1505 tim 2719 W(x) = U(x) - \sum\limits_{\alpha = 1}^N {\frac{{g_\alpha ^2
1506     }}{{2m_\alpha w_\alpha ^2 }}} x^2
1507 tim 2899 \]
1508     and combining the last two terms in Eq.~\ref{introEquation:bathGLE}, we may rewrite the Harmonic bath Hamiltonian as
1509 tim 2696 \[
1510     H = \frac{{p^2 }}{{2m}} + W(x) + \sum\limits_{\alpha = 1}^N
1511     {\left\{ {\frac{{p_\alpha ^2 }}{{2m_\alpha }} + \frac{1}{2}m_\alpha
1512     w_\alpha ^2 \left( {x_\alpha - \frac{{g_\alpha }}{{m_\alpha
1513 tim 2899 w_\alpha ^2 }}x} \right)^2 } \right\}}.
1514 tim 2696 \]
1515     Since the first two terms of the new Hamiltonian depend only on the
1516     system coordinates, we can get the equations of motion for
1517 tim 2872 Generalized Langevin Dynamics by Hamilton's equations,
1518 tim 2719 \begin{equation}
1519     m\ddot x = - \frac{{\partial W(x)}}{{\partial x}} -
1520     \sum\limits_{\alpha = 1}^N {g_\alpha \left( {x_\alpha -
1521     \frac{{g_\alpha }}{{m_\alpha w_\alpha ^2 }}x} \right)},
1522     \label{introEquation:coorMotionGLE}
1523     \end{equation}
1524     and
1525     \begin{equation}
1526     m\ddot x_\alpha = - m_\alpha w_\alpha ^2 \left( {x_\alpha -
1527     \frac{{g_\alpha }}{{m_\alpha w_\alpha ^2 }}x} \right).
1528     \label{introEquation:bathMotionGLE}
1529     \end{equation}
1530     In order to derive an equation for $x$, the dynamics of the bath
1531     variables $x_\alpha$ must be solved exactly first. As an integral
1532     transform which is particularly useful in solving linear ordinary
1533 tim 2872 differential equations,the Laplace transform is the appropriate tool
1534     to solve this problem. The basic idea is to transform the difficult
1535 tim 2719 differential equations into simple algebra problems which can be
1536 tim 2909 solved easily. Then, by applying the inverse Laplace transform, we
1537     can retrieve the solutions of the original problems. Let $f(t)$ be a
1538     function defined on $ [0,\infty ) $, the Laplace transform of $f(t)$
1539     is a new function defined as
1540 tim 2696 \[
1541 tim 2719 L(f(t)) \equiv F(p) = \int_0^\infty {f(t)e^{ - pt} dt}
1542 tim 2696 \]
1543 tim 2719 where $p$ is real and $L$ is called the Laplace Transform
1544     Operator. Below are some important properties of Laplace transform
1545 tim 2789 \begin{eqnarray*}
1546     L(x + y) & = & L(x) + L(y) \\
1547     L(ax) & = & aL(x) \\
1548     L(\dot x) & = & pL(x) - px(0) \\
1549     L(\ddot x)& = & p^2 L(x) - px(0) - \dot x(0) \\
1550     L\left( {\int_0^t {g(t - \tau )h(\tau )d\tau } } \right)& = & G(p)H(p) \\
1551     \end{eqnarray*}
1552 tim 2872 Applying the Laplace transform to the bath coordinates, we obtain
1553 tim 2789 \begin{eqnarray*}
1554 tim 2905 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), \\
1555     L(x_\alpha ) & = & \frac{{\frac{{g_\alpha }}{{\omega _\alpha }}L(x) + px_\alpha (0) + \dot x_\alpha (0)}}{{p^2 + \omega _\alpha ^2 }}. \\
1556 tim 2789 \end{eqnarray*}
1557 tim 2909 In the same way, the system coordinates become
1558 tim 2789 \begin{eqnarray*}
1559 tim 2899 mL(\ddot x) & = &
1560     - \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\}} \\
1561 tim 2905 & & - \frac{1}{p}\frac{{\partial W(x)}}{{\partial x}}.
1562 tim 2789 \end{eqnarray*}
1563 tim 2719 With the help of some relatively important inverse Laplace
1564     transformations:
1565 tim 2696 \[
1566 tim 2719 \begin{array}{c}
1567     L(\cos at) = \frac{p}{{p^2 + a^2 }} \\
1568     L(\sin at) = \frac{a}{{p^2 + a^2 }} \\
1569     L(1) = \frac{1}{p} \\
1570     \end{array}
1571 tim 2696 \]
1572 tim 2899 we obtain
1573 tim 2794 \begin{eqnarray*}
1574     m\ddot x & = & - \frac{{\partial W(x)}}{{\partial x}} -
1575 tim 2696 \sum\limits_{\alpha = 1}^N {\left\{ {\left( { - \frac{{g_\alpha ^2
1576     }}{{m_\alpha \omega _\alpha ^2 }}} \right)\int_0^t {\cos (\omega
1577 tim 2794 _\alpha t)\dot x(t - \tau )d\tau } } \right\}} \\
1578     & & + \sum\limits_{\alpha = 1}^N {\left\{ {\left[ {g_\alpha
1579     x_\alpha (0) - \frac{{g_\alpha }}{{m_\alpha \omega _\alpha }}}
1580     \right]\cos (\omega _\alpha t) + \frac{{g_\alpha \dot x_\alpha
1581 tim 2909 (0)}}{{\omega _\alpha }}\sin (\omega _\alpha t)} \right\}}\\
1582     %
1583     & = & -
1584     \frac{{\partial W(x)}}{{\partial x}} - \int_0^t {\sum\limits_{\alpha
1585     = 1}^N {\left( { - \frac{{g_\alpha ^2 }}{{m_\alpha \omega _\alpha
1586     ^2 }}} \right)\cos (\omega _\alpha
1587 tim 2794 t)\dot x(t - \tau )d} \tau } \\
1588     & & + \sum\limits_{\alpha = 1}^N {\left\{ {\left[ {g_\alpha
1589     x_\alpha (0) - \frac{{g_\alpha }}{{m_\alpha \omega _\alpha }}}
1590     \right]\cos (\omega _\alpha t) + \frac{{g_\alpha \dot x_\alpha
1591     (0)}}{{\omega _\alpha }}\sin (\omega _\alpha t)} \right\}}
1592     \end{eqnarray*}
1593 tim 2719 Introducing a \emph{dynamic friction kernel}
1594 tim 2696 \begin{equation}
1595 tim 2719 \xi (t) = \sum\limits_{\alpha = 1}^N {\left( { - \frac{{g_\alpha ^2
1596     }}{{m_\alpha \omega _\alpha ^2 }}} \right)\cos (\omega _\alpha t)}
1597     \label{introEquation:dynamicFrictionKernelDefinition}
1598     \end{equation}
1599     and \emph{a random force}
1600     \begin{equation}
1601     R(t) = \sum\limits_{\alpha = 1}^N {\left( {g_\alpha x_\alpha (0)
1602     - \frac{{g_\alpha ^2 }}{{m_\alpha \omega _\alpha ^2 }}x(0)}
1603     \right)\cos (\omega _\alpha t)} + \frac{{\dot x_\alpha
1604     (0)}}{{\omega _\alpha }}\sin (\omega _\alpha t),
1605     \label{introEquation:randomForceDefinition}
1606     \end{equation}
1607     the equation of motion can be rewritten as
1608     \begin{equation}
1609 tim 2696 m\ddot x = - \frac{{\partial W}}{{\partial x}} - \int_0^t {\xi
1610     (t)\dot x(t - \tau )d\tau } + R(t)
1611     \label{introEuqation:GeneralizedLangevinDynamics}
1612     \end{equation}
1613 tim 2719 which is known as the \emph{generalized Langevin equation}.
1614    
1615 tim 2819 \subsubsection{\label{introSection:randomForceDynamicFrictionKernel}\textbf{Random Force and Dynamic Friction Kernel}}
1616 tim 2719
1617     One may notice that $R(t)$ depends only on initial conditions, which
1618     implies it is completely deterministic within the context of a
1619     harmonic bath. However, it is easy to verify that $R(t)$ is totally
1620 tim 2905 uncorrelated to $x$ and $\dot x$,$\left\langle {x(t)R(t)}
1621     \right\rangle = 0, \left\langle {\dot x(t)R(t)} \right\rangle =
1622     0.$ This property is what we expect from a truly random process. As
1623     long as the model chosen for $R(t)$ was a gaussian distribution in
1624 tim 2872 general, the stochastic nature of the GLE still remains.
1625 tim 2719 %dynamic friction kernel
1626     The convolution integral
1627 tim 2696 \[
1628 tim 2719 \int_0^t {\xi (t)\dot x(t - \tau )d\tau }
1629 tim 2696 \]
1630 tim 2719 depends on the entire history of the evolution of $x$, which implies
1631     that the bath retains memory of previous motions. In other words,
1632     the bath requires a finite time to respond to change in the motion
1633     of the system. For a sluggish bath which responds slowly to changes
1634     in the system coordinate, we may regard $\xi(t)$ as a constant
1635     $\xi(t) = \Xi_0$. Hence, the convolution integral becomes
1636     \[
1637     \int_0^t {\xi (t)\dot x(t - \tau )d\tau } = \xi _0 (x(t) - x(0))
1638     \]
1639 tim 2899 and Eq.~\ref{introEuqation:GeneralizedLangevinDynamics} becomes
1640 tim 2719 \[
1641     m\ddot x = - \frac{\partial }{{\partial x}}\left( {W(x) +
1642     \frac{1}{2}\xi _0 (x - x_0 )^2 } \right) + R(t),
1643     \]
1644 tim 2872 which can be used to describe the effect of dynamic caging in
1645     viscous solvents. The other extreme is the bath that responds
1646     infinitely quickly to motions in the system. Thus, $\xi (t)$ can be
1647     taken as a $delta$ function in time:
1648 tim 2719 \[
1649     \xi (t) = 2\xi _0 \delta (t)
1650     \]
1651     Hence, the convolution integral becomes
1652     \[
1653     \int_0^t {\xi (t)\dot x(t - \tau )d\tau } = 2\xi _0 \int_0^t
1654     {\delta (t)\dot x(t - \tau )d\tau } = \xi _0 \dot x(t),
1655     \]
1656 tim 2899 and Eq.~\ref{introEuqation:GeneralizedLangevinDynamics} becomes
1657 tim 2719 \begin{equation}
1658     m\ddot x = - \frac{{\partial W(x)}}{{\partial x}} - \xi _0 \dot
1659     x(t) + R(t) \label{introEquation:LangevinEquation}
1660     \end{equation}
1661     which is known as the Langevin equation. The static friction
1662     coefficient $\xi _0$ can either be calculated from spectral density
1663 tim 2850 or be determined by Stokes' law for regular shaped particles. A
1664 tim 2909 brief review on calculating friction tensors for arbitrary shaped
1665 tim 2720 particles is given in Sec.~\ref{introSection:frictionTensor}.
1666 tim 2696
1667 tim 2819 \subsubsection{\label{introSection:secondFluctuationDissipation}\textbf{The Second Fluctuation Dissipation Theorem}}
1668 tim 2719
1669 tim 2905 Defining a new set of coordinates
1670 tim 2696 \[
1671     q_\alpha (t) = x_\alpha (t) - \frac{1}{{m_\alpha \omega _\alpha
1672 tim 2905 ^2 }}x(0),
1673     \]
1674 tim 2719 we can rewrite $R(T)$ as
1675 tim 2696 \[
1676 tim 2719 R(t) = \sum\limits_{\alpha = 1}^N {g_\alpha q_\alpha (t)}.
1677 tim 2696 \]
1678     And since the $q$ coordinates are harmonic oscillators,
1679 tim 2789 \begin{eqnarray*}
1680     \left\langle {q_\alpha ^2 } \right\rangle & = & \frac{{kT}}{{m_\alpha \omega _\alpha ^2 }} \\
1681     \left\langle {q_\alpha (t)q_\alpha (0)} \right\rangle & = & \left\langle {q_\alpha ^2 (0)} \right\rangle \cos (\omega _\alpha t) \\
1682     \left\langle {q_\alpha (t)q_\beta (0)} \right\rangle & = &\delta _{\alpha \beta } \left\langle {q_\alpha (t)q_\alpha (0)} \right\rangle \\
1683     \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 } } \\
1684     & = &\sum\limits_\alpha {g_\alpha ^2 \left\langle {q_\alpha ^2 (0)} \right\rangle \cos (\omega _\alpha t)} \\
1685 tim 2908 & = &kT\xi (t)
1686 tim 2789 \end{eqnarray*}
1687 tim 2719 Thus, we recover the \emph{second fluctuation dissipation theorem}
1688 tim 2696 \begin{equation}
1689     \xi (t) = \left\langle {R(t)R(0)} \right\rangle
1690 tim 2905 \label{introEquation:secondFluctuationDissipation},
1691 tim 2696 \end{equation}
1692 tim 2905 which acts as a constraint on the possible ways in which one can
1693     model the random force and friction kernel.