ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/tengDissertation/Introduction.tex
Revision: 2911
Committed: Thu Jun 29 23:56:11 2006 UTC (18 years, 2 months ago) by tim
Content type: application/x-tex
File size: 74455 byte(s)
Log Message:
version 1.0.0

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