ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/tengDissertation/Introduction.tex
(Generate patch)

Comparing trunk/tengDissertation/Introduction.tex (file contents):
Revision 2692 by tim, Tue Apr 4 21:32:58 2006 UTC vs.
Revision 2895 by tim, Tue Jun 27 02:42:30 2006 UTC

# Line 1 | Line 1
1   \chapter{\label{chapt:introduction}INTRODUCTION AND THEORETICAL BACKGROUND}
2  
3 < \section{\label{introSection:molecularDynamics}Molecular Dynamics}
3 > \section{\label{introSection:classicalMechanics}Classical
4 > Mechanics}
5  
5 As a special discipline of molecular modeling, Molecular dynamics
6 has proven to be a powerful tool for studying the functions of
7 biological systems, providing structural, thermodynamic and
8 dynamical information.
9
10 \subsection{\label{introSection:classicalMechanics}Classical Mechanics}
11
6   Closely related to Classical Mechanics, Molecular Dynamics
7   simulations are carried out by integrating the equations of motion
8   for a given system of particles. There are three fundamental ideas
9 < behind classical mechanics. Firstly, One can determine the state of
9 > behind classical mechanics. Firstly, one can determine the state of
10   a mechanical system at any time of interest; Secondly, all the
11   mechanical properties of the system at that time can be determined
12   by combining the knowledge of the properties of the system with the
# Line 20 | Line 14 | sufficient to predict the future behavior of the syste
14   when further combine with the laws of mechanics will also be
15   sufficient to predict the future behavior of the system.
16  
17 < \subsubsection{\label{introSection:newtonian}Newtonian Mechanics}
17 > \subsection{\label{introSection:newtonian}Newtonian Mechanics}
18 > The discovery of Newton's three laws of mechanics which govern the
19 > motion of particles is the foundation of the classical mechanics.
20 > Newton's first law defines a class of inertial frames. Inertial
21 > frames are reference frames where a particle not interacting with
22 > other bodies will move with constant speed in the same direction.
23 > With respect to inertial frames, Newton's second law has the form
24 > \begin{equation}
25 > F = \frac {dp}{dt} = \frac {mdv}{dt}
26 > \label{introEquation:newtonSecondLaw}
27 > \end{equation}
28 > A point mass interacting with other bodies moves with the
29 > acceleration along the direction of the force acting on it. Let
30 > $F_{ij}$ be the force that particle $i$ exerts on particle $j$, and
31 > $F_{ji}$ be the force that particle $j$ exerts on particle $i$.
32 > Newton's third law states that
33 > \begin{equation}
34 > F_{ij} = -F_{ji}
35 > \label{introEquation:newtonThirdLaw}
36 > \end{equation}
37  
38 < \subsubsection{\label{introSection:lagrangian}Lagrangian Mechanics}
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 > \tau \equiv r \times F \label{introEquation:torqueDefinition}
50 > \end{equation}
51 > Differentiating Eq.~\ref{introEquation:angularMomentumDefinition},
52 > \[
53 > \dot L = \frac{d}{{dt}}(r \times p) = (\dot r \times p) + (r \times
54 > \dot p)
55 > \]
56 > since
57 > \[
58 > \dot r \times p = \dot r \times mv = m\dot r \times \dot r \equiv 0
59 > \]
60 > thus,
61 > \begin{equation}
62 > \dot L = r \times \dot p = \tau
63 > \end{equation}
64 > If there are no external torques acting on a body, the angular
65 > momentum of it is conserved. The last conservation theorem state
66 > that if all forces are conservative, Energy
67 > \begin{equation}E = T + V \label{introEquation:energyConservation}
68 > \end{equation}
69 > is conserved. All of these conserved quantities are
70 > important factors to determine the quality of numerical integration
71 > schemes for rigid bodies \cite{Dullweber1997}.
72  
73 < Newtonian Mechanics suffers from two important limitations: it
28 < describes their motion in special cartesian coordinate systems.
29 < Another limitation of Newtonian mechanics becomes obvious when we
30 < try to describe systems with large numbers of particles. It becomes
31 < very difficult to predict the properties of the system by carrying
32 < out calculations involving the each individual interaction between
33 < all the particles, even if we know all of the details of the
34 < interaction. In order to overcome some of the practical difficulties
35 < which arise in attempts to apply Newton's equation to complex
36 < system, alternative procedures may be developed.
73 > \subsection{\label{introSection:lagrangian}Lagrangian Mechanics}
74  
75 < \subsubsubsection{\label{introSection:halmiltonPrinciple}Hamilton's
76 < Principle}
75 > Newtonian Mechanics suffers from two important limitations: motions
76 > can only be described in cartesian coordinate systems. Moreover, it
77 > becomes impossible to predict analytically the properties of the
78 > system even if we know all of the details of the interaction. In
79 > order to overcome some of the practical difficulties which arise in
80 > attempts to apply Newton's equation to complex system, approximate
81 > numerical procedures may be developed.
82  
83 + \subsubsection{\label{introSection:halmiltonPrinciple}\textbf{Hamilton's
84 + Principle}}
85 +
86   Hamilton introduced the dynamical principle upon which it is
87 < possible to base all of mechanics and, indeed, most of classical
88 < physics. Hamilton's Principle may be stated as follow,
87 > possible to base all of mechanics and most of classical physics.
88 > Hamilton's Principle may be stated as follows,
89  
90   The actual trajectory, along which a dynamical system may move from
91   one point to another within a specified time, is derived by finding
# Line 48 | Line 93 | the kinetic, $K$, and potential energies, $U$.
93   the kinetic, $K$, and potential energies, $U$.
94   \begin{equation}
95   \delta \int_{t_1 }^{t_2 } {(K - U)dt = 0} ,
96 < \lable{introEquation:halmitonianPrinciple1}
96 > \label{introEquation:halmitonianPrinciple1}
97   \end{equation}
98  
99   For simple mechanical systems, where the forces acting on the
100 < different part are derivable from a potential and the velocities are
101 < small compared with that of light, the Lagrangian function $L$ can
102 < be define as the difference between the kinetic energy of the system
58 < and its potential energy,
100 > different parts are derivable from a potential, the Lagrangian
101 > function $L$ can be defined as the difference between the kinetic
102 > energy of the system and its potential energy,
103   \begin{equation}
104   L \equiv K - U = L(q_i ,\dot q_i ) ,
105   \label{introEquation:lagrangianDef}
106   \end{equation}
107   then Eq.~\ref{introEquation:halmitonianPrinciple1} becomes
108   \begin{equation}
109 < \delta \int_{t_1 }^{t_2 } {K dt = 0} ,
110 < \lable{introEquation:halmitonianPrinciple2}
109 > \delta \int_{t_1 }^{t_2 } {L dt = 0} ,
110 > \label{introEquation:halmitonianPrinciple2}
111   \end{equation}
112  
113 < \subsubsubsection{\label{introSection:equationOfMotionLagrangian}The
114 < Equations of Motion in Lagrangian Mechanics}
113 > \subsubsection{\label{introSection:equationOfMotionLagrangian}\textbf{The
114 > Equations of Motion in Lagrangian Mechanics}}
115  
116 < for a holonomic system of $f$ degrees of freedom, the equations of
117 < motion in the Lagrangian form is
116 > For a system of $f$ degrees of freedom, the equations of motion in
117 > the Lagrangian form is
118   \begin{equation}
119   \frac{d}{{dt}}\frac{{\partial L}}{{\partial \dot q_i }} -
120   \frac{{\partial L}}{{\partial q_i }} = 0,{\rm{ }}i = 1, \ldots,f
121 < \lable{introEquation:eqMotionLagrangian}
121 > \label{introEquation:eqMotionLagrangian}
122   \end{equation}
123   where $q_{i}$ is generalized coordinate and $\dot{q_{i}}$ is
124   generalized velocity.
125  
126 < \subsubsection{\label{introSection:hamiltonian}Hamiltonian Mechanics}
126 > \subsection{\label{introSection:hamiltonian}Hamiltonian Mechanics}
127  
128   Arising from Lagrangian Mechanics, Hamiltonian Mechanics was
129   introduced by William Rowan Hamilton in 1833 as a re-formulation of
130   classical mechanics. If the potential energy of a system is
131 < independent of generalized velocities, the generalized momenta can
88 < be defined as
131 > independent of velocities, the momenta can be defined as
132   \begin{equation}
133   p_i = \frac{\partial L}{\partial \dot q_i}
134   \label{introEquation:generalizedMomenta}
135   \end{equation}
136 < With the help of these momenta, we may now define a new quantity $H$
94 < by the equation
136 > The Lagrange equations of motion are then expressed by
137   \begin{equation}
138 < H = p_1 \dot q_1  +  \ldots  + p_f \dot q_f  - L,
138 > p_i  = \frac{{\partial L}}{{\partial q_i }}
139 > \label{introEquation:generalizedMomentaDot}
140 > \end{equation}
141 >
142 > With the help of the generalized momenta, we may now define a new
143 > quantity $H$ by the equation
144 > \begin{equation}
145 > H = \sum\limits_k {p_k \dot q_k }  - L ,
146   \label{introEquation:hamiltonianDefByLagrangian}
147   \end{equation}
148   where $ \dot q_1  \ldots \dot q_f $ are generalized velocities and
149   $L$ is the Lagrangian function for the system.
150  
151 + Differentiating Eq.~\ref{introEquation:hamiltonianDefByLagrangian},
152 + one can obtain
153 + \begin{equation}
154 + dH = \sum\limits_k {\left( {p_k d\dot q_k  + \dot q_k dp_k  -
155 + \frac{{\partial L}}{{\partial q_k }}dq_k  - \frac{{\partial
156 + L}}{{\partial \dot q_k }}d\dot q_k } \right)}  - \frac{{\partial
157 + L}}{{\partial t}}dt \label{introEquation:diffHamiltonian1}
158 + \end{equation}
159 + Making use of  Eq.~\ref{introEquation:generalizedMomenta}, the
160 + second and fourth terms in the parentheses cancel. Therefore,
161 + Eq.~\ref{introEquation:diffHamiltonian1} can be rewritten as
162 + \begin{equation}
163 + dH = \sum\limits_k {\left( {\dot q_k dp_k  - \dot p_k dq_k }
164 + \right)}  - \frac{{\partial L}}{{\partial t}}dt
165 + \label{introEquation:diffHamiltonian2}
166 + \end{equation}
167 + By identifying the coefficients of $dq_k$, $dp_k$ and dt, we can
168 + find
169 + \begin{equation}
170 + \frac{{\partial H}}{{\partial p_k }} = \dot {q_k}
171 + \label{introEquation:motionHamiltonianCoordinate}
172 + \end{equation}
173 + \begin{equation}
174 + \frac{{\partial H}}{{\partial q_k }} =  - \dot {p_k}
175 + \label{introEquation:motionHamiltonianMomentum}
176 + \end{equation}
177 + and
178 + \begin{equation}
179 + \frac{{\partial H}}{{\partial t}} =  - \frac{{\partial L}}{{\partial
180 + t}}
181 + \label{introEquation:motionHamiltonianTime}
182 + \end{equation}
183 +
184 + Eq.~\ref{introEquation:motionHamiltonianCoordinate} and
185 + Eq.~\ref{introEquation:motionHamiltonianMomentum} are Hamilton's
186 + equation of motion. Due to their symmetrical formula, they are also
187 + known as the canonical equations of motions \cite{Goldstein2001}.
188 +
189   An important difference between Lagrangian approach and the
190   Hamiltonian approach is that the Lagrangian is considered to be a
191 < function of the generalized velocities $\dot q_i$ and the
192 < generalized coordinates $q_i$, while the Hamiltonian is considered
193 < to be a function of the generalized momenta $p_i$ and the conjugate
194 < generalized coordinate $q_i$. Hamiltonian Mechanics is more
195 < appropriate for application to statistical mechanics and quantum
196 < mechanics, since it treats the coordinate and its time derivative as
197 < independent variables and it only works with 1st-order differential
111 < equations.
191 > function of the generalized velocities $\dot q_i$ and coordinates
192 > $q_i$, while the Hamiltonian is considered to be a function of the
193 > generalized momenta $p_i$ and the conjugate coordinates $q_i$.
194 > Hamiltonian Mechanics is more appropriate for application to
195 > statistical mechanics and quantum mechanics, since it treats the
196 > coordinate and its time derivative as independent variables and it
197 > only works with 1st-order differential equations\cite{Marion1990}.
198  
199 + In Newtonian Mechanics, a system described by conservative forces
200 + conserves the total energy \ref{introEquation:energyConservation}.
201 + It follows that Hamilton's equations of motion conserve the total
202 + Hamiltonian.
203 + \begin{equation}
204 + \frac{{dH}}{{dt}} = \sum\limits_i {\left( {\frac{{\partial
205 + H}}{{\partial q_i }}\dot q_i  + \frac{{\partial H}}{{\partial p_i
206 + }}\dot p_i } \right)}  = \sum\limits_i {\left( {\frac{{\partial
207 + H}}{{\partial q_i }}\frac{{\partial H}}{{\partial p_i }} -
208 + \frac{{\partial H}}{{\partial p_i }}\frac{{\partial H}}{{\partial
209 + q_i }}} \right) = 0} \label{introEquation:conserveHalmitonian}
210 + \end{equation}
211  
212 < \subsubsection{\label{introSection:canonicalTransformation}Canonical Transformation}
212 > \section{\label{introSection:statisticalMechanics}Statistical
213 > Mechanics}
214  
215 < \subsection{\label{introSection:statisticalMechanics}Statistical Mechanics}
117 <
118 < The thermodynamic behaviors and properties  of Molecular Dynamics
215 > The thermodynamic behaviors and properties of Molecular Dynamics
216   simulation are governed by the principle of Statistical Mechanics.
217   The following section will give a brief introduction to some of the
218 < Statistical Mechanics concepts presented in this dissertation.
218 > Statistical Mechanics concepts and theorem presented in this
219 > dissertation.
220  
221 < \subsubsection{\label{introSection::ensemble}Ensemble}
221 > \subsection{\label{introSection:ensemble}Phase Space and Ensemble}
222  
223 < \subsubsection{\label{introSection:ergodic}The Ergodic Hypothesis}
224 <
225 < \subsection{\label{introSection:rigidBody}Dynamics of Rigid Bodies}
226 <
227 < \subsection{\label{introSection:correlationFunctions}Correlation Functions}
228 <
229 < \section{\label{introSection:langevinDynamics}Langevin Dynamics}
230 <
231 < \subsection{\label{introSection:generalizedLangevinDynamics}Generalized Langevin Dynamics}
223 > Mathematically, phase space is the space which represents all
224 > possible states. Each possible state of the system corresponds to
225 > one unique point in the phase space. For mechanical systems, the
226 > phase space usually consists of all possible values of position and
227 > momentum variables. Consider a dynamic system of $f$ particles in a
228 > cartesian space, where each of the $6f$ coordinates and momenta is
229 > assigned to one of $6f$ mutually orthogonal axes, the phase space of
230 > this system is a $6f$ dimensional space. A point, $x = (\rightarrow
231 > q_1 , \ldots ,\rightarrow q_f ,\rightarrow p_1 , \ldots ,\rightarrow
232 > p_f )$, with a unique set of values of $6f$ coordinates and momenta
233 > is a phase space vector.
234 > %%%fix me
235  
236 < \subsection{\label{introSection:hydroynamics}Hydrodynamics}
236 > In statistical mechanics, the condition of an ensemble at any time
237 > can be regarded as appropriately specified by the density $\rho$
238 > with which representative points are distributed over the phase
239 > space. The density distribution for an ensemble with $f$ degrees of
240 > freedom is defined as,
241 > \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 > their locations which would change the density at any time at phase
247 > space. Hence, the density distribution is also to be taken as a
248 > function of the time.
249 >
250 > The number of systems $\delta N$ at time $t$ can be determined by,
251 > \begin{equation}
252 > \delta N = \rho (q,p,t)dq_1  \ldots dq_f dp_1  \ldots dp_f.
253 > \label{introEquation:deltaN}
254 > \end{equation}
255 > Assuming a large enough population of systems, we can sufficiently
256 > approximate $\delta N$ without introducing discontinuity when we go
257 > from one region in the phase space to another. By integrating over
258 > the whole phase space,
259 > \begin{equation}
260 > N = \int { \ldots \int {\rho (q,p,t)dq_1 } ...dq_f dp_1 } ...dp_f
261 > \label{introEquation:totalNumberSystem}
262 > \end{equation}
263 > gives us an expression for the total number of the systems. Hence,
264 > the probability per unit in the phase space can be obtained by,
265 > \begin{equation}
266 > \frac{{\rho (q,p,t)}}{N} = \frac{{\rho (q,p,t)}}{{\int { \ldots \int
267 > {\rho (q,p,t)dq_1 } ...dq_f dp_1 } ...dp_f }}.
268 > \label{introEquation:unitProbability}
269 > \end{equation}
270 > With the help of Eq.~\ref{introEquation:unitProbability} and the
271 > knowledge of the system, it is possible to calculate the average
272 > value of any desired quantity which depends on the coordinates and
273 > momenta of the system. Even when the dynamics of the real system is
274 > complex, or stochastic, or even discontinuous, the average
275 > properties of the ensemble of possibilities as a whole remaining
276 > well defined. For a classical system in thermal equilibrium with its
277 > environment, the ensemble average of a mechanical quantity, $\langle
278 > A(q , p) \rangle_t$, takes the form of an integral over the phase
279 > space of the system,
280 > \begin{equation}
281 > \langle  A(q , p) \rangle_t = \frac{{\int { \ldots \int {A(q,p)\rho
282 > (q,p,t)dq_1 } ...dq_f dp_1 } ...dp_f }}{{\int { \ldots \int {\rho
283 > (q,p,t)dq_1 } ...dq_f dp_1 } ...dp_f }}
284 > \label{introEquation:ensembelAverage}
285 > \end{equation}
286 >
287 > There are several different types of ensembles with different
288 > statistical characteristics. As a function of macroscopic
289 > parameters, such as temperature \textit{etc}, the partition function
290 > can be used to describe the statistical properties of a system in
291 > thermodynamic equilibrium.
292 >
293 > As an ensemble of systems, each of which is known to be thermally
294 > isolated and conserve energy, the Microcanonical ensemble (NVE) has
295 > a partition function like,
296 > \begin{equation}
297 > \Omega (N,V,E) = e^{\beta TS} \label{introEquation:NVEPartition}.
298 > \end{equation}
299 > A canonical ensemble (NVT)is an ensemble of systems, each of which
300 > can share its energy with a large heat reservoir. The distribution
301 > of the total energy amongst the possible dynamical states is given
302 > by the partition function,
303 > \begin{equation}
304 > \Omega (N,V,T) = e^{ - \beta A}
305 > \label{introEquation:NVTPartition}
306 > \end{equation}
307 > Here, $A$ is the Helmholtz free energy which is defined as $ A = U -
308 > TS$. Since most experiments are carried out under constant pressure
309 > condition, the isothermal-isobaric ensemble (NPT) plays a very
310 > important role in molecular simulations. The isothermal-isobaric
311 > ensemble allow the system to exchange energy with a heat bath of
312 > temperature $T$ and to change the volume as well. Its partition
313 > function is given as
314 > \begin{equation}
315 > \Delta (N,P,T) =  - e^{\beta G}.
316 > \label{introEquation:NPTPartition}
317 > \end{equation}
318 > Here, $G = U - TS + PV$, and $G$ is called Gibbs free energy.
319 >
320 > \subsection{\label{introSection:liouville}Liouville's theorem}
321 >
322 > Liouville's theorem is the foundation on which statistical mechanics
323 > rests. It describes the time evolution of the phase space
324 > distribution function. In order to calculate the rate of change of
325 > $\rho$, we begin from Eq.~\ref{introEquation:deltaN}. If we consider
326 > the two faces perpendicular to the $q_1$ axis, which are located at
327 > $q_1$ and $q_1 + \delta q_1$, the number of phase points leaving the
328 > opposite face is given by the expression,
329 > \begin{equation}
330 > \left( {\rho  + \frac{{\partial \rho }}{{\partial q_1 }}\delta q_1 }
331 > \right)\left( {\dot q_1  + \frac{{\partial \dot q_1 }}{{\partial q_1
332 > }}\delta q_1 } \right)\delta q_2  \ldots \delta q_f \delta p_1
333 > \ldots \delta p_f .
334 > \end{equation}
335 > Summing all over the phase space, we obtain
336 > \begin{equation}
337 > \frac{{d(\delta N)}}{{dt}} =  - \sum\limits_{i = 1}^f {\left[ {\rho
338 > \left( {\frac{{\partial \dot q_i }}{{\partial q_i }} +
339 > \frac{{\partial \dot p_i }}{{\partial p_i }}} \right) + \left(
340 > {\frac{{\partial \rho }}{{\partial q_i }}\dot q_i  + \frac{{\partial
341 > \rho }}{{\partial p_i }}\dot p_i } \right)} \right]} \delta q_1
342 > \ldots \delta q_f \delta p_1  \ldots \delta p_f .
343 > \end{equation}
344 > Differentiating the equations of motion in Hamiltonian formalism
345 > (\ref{introEquation:motionHamiltonianCoordinate},
346 > \ref{introEquation:motionHamiltonianMomentum}), we can show,
347 > \begin{equation}
348 > \sum\limits_i {\left( {\frac{{\partial \dot q_i }}{{\partial q_i }}
349 > + \frac{{\partial \dot p_i }}{{\partial p_i }}} \right)}  = 0 ,
350 > \end{equation}
351 > which cancels the first terms of the right hand side. Furthermore,
352 > dividing $ \delta q_1  \ldots \delta q_f \delta p_1  \ldots \delta
353 > p_f $ in both sides, we can write out Liouville's theorem in a
354 > simple form,
355 > \begin{equation}
356 > \frac{{\partial \rho }}{{\partial t}} + \sum\limits_{i = 1}^f
357 > {\left( {\frac{{\partial \rho }}{{\partial q_i }}\dot q_i  +
358 > \frac{{\partial \rho }}{{\partial p_i }}\dot p_i } \right)}  = 0 .
359 > \label{introEquation:liouvilleTheorem}
360 > \end{equation}
361 >
362 > Liouville's theorem states that the distribution function is
363 > constant along any trajectory in phase space. In classical
364 > statistical mechanics, since the number of members in an ensemble is
365 > huge and constant, we can assume the local density has no reason
366 > (other than classical mechanics) to change,
367 > \begin{equation}
368 > \frac{{\partial \rho }}{{\partial t}} = 0.
369 > \label{introEquation:stationary}
370 > \end{equation}
371 > In such stationary system, the density of distribution $\rho$ can be
372 > connected to the Hamiltonian $H$ through Maxwell-Boltzmann
373 > distribution,
374 > \begin{equation}
375 > \rho  \propto e^{ - \beta H}
376 > \label{introEquation:densityAndHamiltonian}
377 > \end{equation}
378 >
379 > \subsubsection{\label{introSection:phaseSpaceConservation}\textbf{Conservation of Phase Space}}
380 > Lets consider a region in the phase space,
381 > \begin{equation}
382 > \delta v = \int { \ldots \int {dq_1 } ...dq_f dp_1 } ..dp_f .
383 > \end{equation}
384 > If this region is small enough, the density $\rho$ can be regarded
385 > as uniform over the whole integral. Thus, the number of phase points
386 > inside this region is given by,
387 > \begin{equation}
388 > \delta N = \rho \delta v = \rho \int { \ldots \int {dq_1 } ...dq_f
389 > dp_1 } ..dp_f.
390 > \end{equation}
391 >
392 > \begin{equation}
393 > \frac{{d(\delta N)}}{{dt}} = \frac{{d\rho }}{{dt}}\delta v + \rho
394 > \frac{d}{{dt}}(\delta v) = 0.
395 > \end{equation}
396 > With the help of stationary assumption
397 > (\ref{introEquation:stationary}), we obtain the principle of the
398 > \emph{conservation of volume in phase space},
399 > \begin{equation}
400 > \frac{d}{{dt}}(\delta v) = \frac{d}{{dt}}\int { \ldots \int {dq_1 }
401 > ...dq_f dp_1 } ..dp_f  = 0.
402 > \label{introEquation:volumePreserving}
403 > \end{equation}
404 >
405 > \subsubsection{\label{introSection:liouvilleInOtherForms}\textbf{Liouville's Theorem in Other Forms}}
406 >
407 > Liouville's theorem can be expresses in a variety of different forms
408 > which are convenient within different contexts. For any two function
409 > $F$ and $G$ of the coordinates and momenta of a system, the Poisson
410 > bracket ${F, G}$ is defined as
411 > \begin{equation}
412 > \left\{ {F,G} \right\} = \sum\limits_i {\left( {\frac{{\partial
413 > F}}{{\partial q_i }}\frac{{\partial G}}{{\partial p_i }} -
414 > \frac{{\partial F}}{{\partial p_i }}\frac{{\partial G}}{{\partial
415 > q_i }}} \right)}.
416 > \label{introEquation:poissonBracket}
417 > \end{equation}
418 > Substituting equations of motion in Hamiltonian formalism(
419 > Eq.~\ref{introEquation:motionHamiltonianCoordinate} ,
420 > Eq.~\ref{introEquation:motionHamiltonianMomentum} ) into
421 > (Eq.~\ref{introEquation:liouvilleTheorem}), we can rewrite
422 > Liouville's theorem using Poisson bracket notion,
423 > \begin{equation}
424 > \left( {\frac{{\partial \rho }}{{\partial t}}} \right) =  - \left\{
425 > {\rho ,H} \right\}.
426 > \label{introEquation:liouvilleTheromInPoissin}
427 > \end{equation}
428 > Moreover, the Liouville operator is defined as
429 > \begin{equation}
430 > iL = \sum\limits_{i = 1}^f {\left( {\frac{{\partial H}}{{\partial
431 > p_i }}\frac{\partial }{{\partial q_i }} - \frac{{\partial
432 > H}}{{\partial q_i }}\frac{\partial }{{\partial p_i }}} \right)}
433 > \label{introEquation:liouvilleOperator}
434 > \end{equation}
435 > In terms of Liouville operator, Liouville's equation can also be
436 > expressed as
437 > \begin{equation}
438 > \left( {\frac{{\partial \rho }}{{\partial t}}} \right) =  - iL\rho
439 > \label{introEquation:liouvilleTheoremInOperator}
440 > \end{equation}
441 >
442 > \subsection{\label{introSection:ergodic}The Ergodic Hypothesis}
443 >
444 > Various thermodynamic properties can be calculated from Molecular
445 > Dynamics simulation. By comparing experimental values with the
446 > calculated properties, one can determine the accuracy of the
447 > simulation and the quality of the underlying model. However, both
448 > experiments and computer simulations are usually performed during a
449 > certain time interval and the measurements are averaged over a
450 > period of them which is different from the average behavior of
451 > many-body system in Statistical Mechanics. Fortunately, the Ergodic
452 > Hypothesis makes a connection between time average and the ensemble
453 > average. It states that the time average and average over the
454 > statistical ensemble are identical \cite{Frenkel1996, Leach2001}.
455 > \begin{equation}
456 > \langle A(q , p) \rangle_t = \mathop {\lim }\limits_{t \to \infty }
457 > \frac{1}{t}\int\limits_0^t {A(q(t),p(t))dt = \int\limits_\Gamma
458 > {A(q(t),p(t))} } \rho (q(t), p(t)) dqdp
459 > \end{equation}
460 > where $\langle  A(q , p) \rangle_t$ is an equilibrium value of a
461 > physical quantity and $\rho (p(t), q(t))$ is the equilibrium
462 > distribution function. If an observation is averaged over a
463 > sufficiently long time (longer than relaxation time), all accessible
464 > microstates in phase space are assumed to be equally probed, giving
465 > a properly weighted statistical average. This allows the researcher
466 > freedom of choice when deciding how best to measure a given
467 > observable. In case an ensemble averaged approach sounds most
468 > reasonable, the Monte Carlo techniques\cite{Metropolis1949} can be
469 > utilized. Or if the system lends itself to a time averaging
470 > approach, the Molecular Dynamics techniques in
471 > Sec.~\ref{introSection:molecularDynamics} will be the best
472 > choice\cite{Frenkel1996}.
473 >
474 > \section{\label{introSection:geometricIntegratos}Geometric Integrators}
475 > A variety of numerical integrators have been proposed to simulate
476 > the motions of atoms in MD simulation. They usually begin with
477 > initial conditionals and move the objects in the direction governed
478 > by the differential equations. However, most of them ignore the
479 > hidden physical laws contained within the equations. Since 1990,
480 > geometric integrators, which preserve various phase-flow invariants
481 > such as symplectic structure, volume and time reversal symmetry, are
482 > developed to address this issue\cite{Dullweber1997, McLachlan1998,
483 > Leimkuhler1999}. The velocity Verlet method, which happens to be a
484 > simple example of symplectic integrator, continues to gain
485 > popularity in the molecular dynamics community. This fact can be
486 > partly explained by its geometric nature.
487 >
488 > \subsection{\label{introSection:symplecticManifold}Symplectic Manifolds}
489 > A \emph{manifold} is an abstract mathematical space. It looks
490 > locally like Euclidean space, but when viewed globally, it may have
491 > more complicated structure. A good example of manifold is the
492 > surface of Earth. It seems to be flat locally, but it is round if
493 > viewed as a whole. A \emph{differentiable manifold} (also known as
494 > \emph{smooth manifold}) is a manifold on which it is possible to
495 > apply calculus on \emph{differentiable manifold}. A \emph{symplectic
496 > manifold} is defined as a pair $(M, \omega)$ which consists of a
497 > \emph{differentiable manifold} $M$ and a close, non-degenerated,
498 > bilinear symplectic form, $\omega$. A symplectic form on a vector
499 > space $V$ is a function $\omega(x, y)$ which satisfies
500 > $\omega(\lambda_1x_1+\lambda_2x_2, y) = \lambda_1\omega(x_1, y)+
501 > \lambda_2\omega(x_2, y)$, $\omega(x, y) = - \omega(y, x)$ and
502 > $\omega(x, x) = 0$. The cross product operation in vector field is
503 > an example of symplectic form.
504 >
505 > One of the motivations to study \emph{symplectic manifolds} in
506 > Hamiltonian Mechanics is that a symplectic manifold can represent
507 > all possible configurations of the system and the phase space of the
508 > system can be described by it's cotangent bundle. Every symplectic
509 > manifold is even dimensional. For instance, in Hamilton equations,
510 > coordinate and momentum always appear in pairs.
511 >
512 > \subsection{\label{introSection:ODE}Ordinary Differential Equations}
513 >
514 > For an ordinary differential system defined as
515 > \begin{equation}
516 > \dot x = f(x)
517 > \end{equation}
518 > where $x = x(q,p)^T$, this system is a canonical Hamiltonian, if
519 > \begin{equation}
520 > f(r) = J\nabla _x H(r).
521 > \end{equation}
522 > $H = H (q, p)$ is Hamiltonian function and $J$ is the skew-symmetric
523 > matrix
524 > \begin{equation}
525 > J = \left( {\begin{array}{*{20}c}
526 >   0 & I  \\
527 >   { - I} & 0  \\
528 > \end{array}} \right)
529 > \label{introEquation:canonicalMatrix}
530 > \end{equation}
531 > where $I$ is an identity matrix. Using this notation, Hamiltonian
532 > system can be rewritten as,
533 > \begin{equation}
534 > \frac{d}{{dt}}x = J\nabla _x H(x)
535 > \label{introEquation:compactHamiltonian}
536 > \end{equation}In this case, $f$ is
537 > called a \emph{Hamiltonian vector field}.
538 >
539 > Another generalization of Hamiltonian dynamics is Poisson
540 > Dynamics\cite{Olver1986},
541 > \begin{equation}
542 > \dot x = J(x)\nabla _x H \label{introEquation:poissonHamiltonian}
543 > \end{equation}
544 > The most obvious change being that matrix $J$ now depends on $x$.
545 >
546 > \subsection{\label{introSection:exactFlow}Exact Flow}
547 >
548 > Let $x(t)$ be the exact solution of the ODE system,
549 > \begin{equation}
550 > \frac{{dx}}{{dt}} = f(x) \label{introEquation:ODE}
551 > \end{equation}
552 > The exact flow(solution) $\varphi_\tau$ is defined by
553 > \[
554 > x(t+\tau) =\varphi_\tau(x(t))
555 > \]
556 > where $\tau$ is a fixed time step and $\varphi$ is a map from phase
557 > space to itself. The flow has the continuous group property,
558 > \begin{equation}
559 > \varphi _{\tau _1 }  \circ \varphi _{\tau _2 }  = \varphi _{\tau _1
560 > + \tau _2 } .
561 > \end{equation}
562 > In particular,
563 > \begin{equation}
564 > \varphi _\tau   \circ \varphi _{ - \tau }  = I
565 > \end{equation}
566 > Therefore, the exact flow is self-adjoint,
567 > \begin{equation}
568 > \varphi _\tau   = \varphi _{ - \tau }^{ - 1}.
569 > \end{equation}
570 > The exact flow can also be written in terms of the of an operator,
571 > \begin{equation}
572 > \varphi _\tau  (x) = e^{\tau \sum\limits_i {f_i (x)\frac{\partial
573 > }{{\partial x_i }}} } (x) \equiv \exp (\tau f)(x).
574 > \label{introEquation:exponentialOperator}
575 > \end{equation}
576 >
577 > In most cases, it is not easy to find the exact flow $\varphi_\tau$.
578 > Instead, we use an approximate map, $\psi_\tau$, which is usually
579 > called integrator. The order of an integrator $\psi_\tau$ is $p$, if
580 > the Taylor series of $\psi_\tau$ agree to order $p$,
581 > \begin{equation}
582 > \psi_\tau(x) = x + \tau f(x) + O(\tau^{p+1})
583 > \end{equation}
584 >
585 > \subsection{\label{introSection:geometricProperties}Geometric Properties}
586 >
587 > The hidden geometric properties\cite{Budd1999, Marsden1998} of an
588 > ODE and its flow play important roles in numerical studies. Many of
589 > them can be found in systems which occur naturally in applications.
590 >
591 > Let $\varphi$ be the flow of Hamiltonian vector field, $\varphi$ is
592 > a \emph{symplectic} flow if it satisfies,
593 > \begin{equation}
594 > {\varphi '}^T J \varphi ' = J.
595 > \end{equation}
596 > According to Liouville's theorem, the symplectic volume is invariant
597 > under a Hamiltonian flow, which is the basis for classical
598 > statistical mechanics. Furthermore, the flow of a Hamiltonian vector
599 > field on a symplectic manifold can be shown to be a
600 > symplectomorphism. As to the Poisson system,
601 > \begin{equation}
602 > {\varphi '}^T J \varphi ' = J \circ \varphi
603 > \end{equation}
604 > is the property that must be preserved by the integrator.
605 >
606 > It is possible to construct a \emph{volume-preserving} flow for a
607 > source free ODE ($ \nabla \cdot f = 0 $), if the flow satisfies $
608 > \det d\varphi  = 1$. One can show easily that a symplectic flow will
609 > be volume-preserving.
610 >
611 > Changing the variables $y = h(x)$ in an ODE
612 > (Eq.~\ref{introEquation:ODE}) will result in a new system,
613 > \[
614 > \dot y = \tilde f(y) = ((dh \cdot f)h^{ - 1} )(y).
615 > \]
616 > The vector filed $f$ has reversing symmetry $h$ if $f = - \tilde f$.
617 > In other words, the flow of this vector field is reversible if and
618 > only if $ h \circ \varphi ^{ - 1}  = \varphi  \circ h $.
619 >
620 > A \emph{first integral}, or conserved quantity of a general
621 > differential function is a function $ G:R^{2d}  \to R^d $ which is
622 > constant for all solutions of the ODE $\frac{{dx}}{{dt}} = f(x)$ ,
623 > \[
624 > \frac{{dG(x(t))}}{{dt}} = 0.
625 > \]
626 > Using chain rule, one may obtain,
627 > \[
628 > \sum\limits_i {\frac{{dG}}{{dx_i }}} f_i (x) = f \bullet \nabla G,
629 > \]
630 > which is the condition for conserving \emph{first integral}. For a
631 > canonical Hamiltonian system, the time evolution of an arbitrary
632 > smooth function $G$ is given by,
633 >
634 > \begin{eqnarray}
635 > \frac{{dG(x(t))}}{{dt}} & = & [\nabla _x G(x(t))]^T \dot x(t) \\
636 >                        & = & [\nabla _x G(x(t))]^T J\nabla _x H(x(t)). \\
637 > \label{introEquation:firstIntegral1}
638 > \end{eqnarray}
639 >
640 >
641 > Using poisson bracket notion, Equation
642 > \ref{introEquation:firstIntegral1} can be rewritten as
643 > \[
644 > \frac{d}{{dt}}G(x(t)) = \left\{ {G,H} \right\}(x(t)).
645 > \]
646 > Therefore, the sufficient condition for $G$ to be the \emph{first
647 > integral} of a Hamiltonian system is
648 > \[
649 > \left\{ {G,H} \right\} = 0.
650 > \]
651 > As well known, the Hamiltonian (or energy) H of a Hamiltonian system
652 > is a \emph{first integral}, which is due to the fact $\{ H,H\}  =
653 > 0$.
654 >
655 > When designing any numerical methods, one should always try to
656 > preserve the structural properties of the original ODE and its flow.
657 >
658 > \subsection{\label{introSection:constructionSymplectic}Construction of Symplectic Methods}
659 > A lot of well established and very effective numerical methods have
660 > been successful precisely because of their symplecticities even
661 > though this fact was not recognized when they were first
662 > constructed. The most famous example is the Verlet-leapfrog method
663 > in molecular dynamics. In general, symplectic integrators can be
664 > constructed using one of four different methods.
665 > \begin{enumerate}
666 > \item Generating functions
667 > \item Variational methods
668 > \item Runge-Kutta methods
669 > \item Splitting methods
670 > \end{enumerate}
671 >
672 > Generating function\cite{Channell1990} tends to lead to methods
673 > which are cumbersome and difficult to use. In dissipative systems,
674 > variational methods can capture the decay of energy
675 > accurately\cite{Kane2000}. Since their geometrically unstable nature
676 > against non-Hamiltonian perturbations, ordinary implicit Runge-Kutta
677 > methods are not suitable for Hamiltonian system. Recently, various
678 > high-order explicit Runge-Kutta methods
679 > \cite{Owren1992,Chen2003}have been developed to overcome this
680 > instability. However, due to computational penalty involved in
681 > implementing the Runge-Kutta methods, they have not attracted much
682 > attention from the Molecular Dynamics community. Instead, splitting
683 > methods have been widely accepted since they exploit natural
684 > decompositions of the system\cite{Tuckerman1992, McLachlan1998}.
685 >
686 > \subsubsection{\label{introSection:splittingMethod}\textbf{Splitting Methods}}
687 >
688 > The main idea behind splitting methods is to decompose the discrete
689 > $\varphi_h$ as a composition of simpler flows,
690 > \begin{equation}
691 > \varphi _h  = \varphi _{h_1 }  \circ \varphi _{h_2 }  \ldots  \circ
692 > \varphi _{h_n }
693 > \label{introEquation:FlowDecomposition}
694 > \end{equation}
695 > where each of the sub-flow is chosen such that each represent a
696 > simpler integration of the system.
697 >
698 > Suppose that a Hamiltonian system takes the form,
699 > \[
700 > H = H_1 + H_2.
701 > \]
702 > Here, $H_1$ and $H_2$ may represent different physical processes of
703 > the system. For instance, they may relate to kinetic and potential
704 > energy respectively, which is a natural decomposition of the
705 > problem. If $H_1$ and $H_2$ can be integrated using exact flows
706 > $\varphi_1(t)$ and $\varphi_2(t)$, respectively, a simple first
707 > order expression is then given by the Lie-Trotter formula
708 > \begin{equation}
709 > \varphi _h  = \varphi _{1,h}  \circ \varphi _{2,h},
710 > \label{introEquation:firstOrderSplitting}
711 > \end{equation}
712 > where $\varphi _h$ is the result of applying the corresponding
713 > continuous $\varphi _i$ over a time $h$. By definition, as
714 > $\varphi_i(t)$ is the exact solution of a Hamiltonian system, it
715 > must follow that each operator $\varphi_i(t)$ is a symplectic map.
716 > It is easy to show that any composition of symplectic flows yields a
717 > symplectic map,
718 > \begin{equation}
719 > (\varphi '\phi ')^T J\varphi '\phi ' = \phi '^T \varphi '^T J\varphi
720 > '\phi ' = \phi '^T J\phi ' = J,
721 > \label{introEquation:SymplecticFlowComposition}
722 > \end{equation}
723 > where $\phi$ and $\psi$ both are symplectic maps. Thus operator
724 > splitting in this context automatically generates a symplectic map.
725 >
726 > The Lie-Trotter splitting(\ref{introEquation:firstOrderSplitting})
727 > introduces local errors proportional to $h^2$, while Strang
728 > splitting gives a second-order decomposition,
729 > \begin{equation}
730 > \varphi _h  = \varphi _{1,h/2}  \circ \varphi _{2,h}  \circ \varphi
731 > _{1,h/2} , \label{introEquation:secondOrderSplitting}
732 > \end{equation}
733 > which has a local error proportional to $h^3$. The Sprang
734 > splitting's popularity in molecular simulation community attribute
735 > to its symmetric property,
736 > \begin{equation}
737 > \varphi _h^{ - 1} = \varphi _{ - h}.
738 > \label{introEquation:timeReversible}
739 > \end{equation}
740 >
741 > \subsubsection{\label{introSection:exampleSplittingMethod}\textbf{Examples of the Splitting Method}}
742 > The classical equation for a system consisting of interacting
743 > particles can be written in Hamiltonian form,
744 > \[
745 > H = T + V
746 > \]
747 > where $T$ is the kinetic energy and $V$ is the potential energy.
748 > Setting $H_1 = T, H_2 = V$ and applying the Strang splitting, one
749 > obtains the following:
750 > \begin{align}
751 > q(\Delta t) &= q(0) + \dot{q}(0)\Delta t +
752 >    \frac{F[q(0)]}{m}\frac{\Delta t^2}{2}, %
753 > \label{introEquation:Lp10a} \\%
754 > %
755 > \dot{q}(\Delta t) &= \dot{q}(0) + \frac{\Delta t}{2m}
756 >    \biggl [F[q(0)] + F[q(\Delta t)] \biggr]. %
757 > \label{introEquation:Lp10b}
758 > \end{align}
759 > where $F(t)$ is the force at time $t$. This integration scheme is
760 > known as \emph{velocity verlet} which is
761 > symplectic(\ref{introEquation:SymplecticFlowComposition}),
762 > time-reversible(\ref{introEquation:timeReversible}) and
763 > volume-preserving (\ref{introEquation:volumePreserving}). These
764 > geometric properties attribute to its long-time stability and its
765 > popularity in the community. However, the most commonly used
766 > velocity verlet integration scheme is written as below,
767 > \begin{align}
768 > \dot{q}\biggl (\frac{\Delta t}{2}\biggr ) &=
769 >    \dot{q}(0) + \frac{\Delta t}{2m}\, F[q(0)], \label{introEquation:Lp9a}\\%
770 > %
771 > q(\Delta t) &= q(0) + \Delta t\, \dot{q}\biggl (\frac{\Delta t}{2}\biggr ),%
772 >    \label{introEquation:Lp9b}\\%
773 > %
774 > \dot{q}(\Delta t) &= \dot{q}\biggl (\frac{\Delta t}{2}\biggr ) +
775 >    \frac{\Delta t}{2m}\, F[q(t)]. \label{introEquation:Lp9c}
776 > \end{align}
777 > From the preceding splitting, one can see that the integration of
778 > the equations of motion would follow:
779 > \begin{enumerate}
780 > \item calculate the velocities at the half step, $\frac{\Delta t}{2}$, from the forces calculated at the initial position.
781 >
782 > \item Use the half step velocities to move positions one whole step, $\Delta t$.
783 >
784 > \item Evaluate the forces at the new positions, $\mathbf{q}(\Delta t)$, and use the new forces to complete the velocity move.
785 >
786 > \item Repeat from step 1 with the new position, velocities, and forces assuming the roles of the initial values.
787 > \end{enumerate}
788 >
789 > By simply switching the order of the propagators in the splitting
790 > and composing a new integrator, the \emph{position verlet}
791 > integrator, can be generated,
792 > \begin{align}
793 > \dot q(\Delta t) &= \dot q(0) + \Delta tF(q(0))\left[ {q(0) +
794 > \frac{{\Delta t}}{{2m}}\dot q(0)} \right], %
795 > \label{introEquation:positionVerlet1} \\%
796 > %
797 > q(\Delta t) &= q(0) + \frac{{\Delta t}}{2}\left[ {\dot q(0) + \dot
798 > q(\Delta t)} \right]. %
799 > \label{introEquation:positionVerlet2}
800 > \end{align}
801 >
802 > \subsubsection{\label{introSection:errorAnalysis}\textbf{Error Analysis and Higher Order Methods}}
803 >
804 > The Baker-Campbell-Hausdorff formula can be used to determine the
805 > local error of splitting method in terms of the commutator of the
806 > operators(\ref{introEquation:exponentialOperator}) associated with
807 > the sub-flow. For operators $hX$ and $hY$ which are associated with
808 > $\varphi_1(t)$ and $\varphi_2(t)$ respectively , we have
809 > \begin{equation}
810 > \exp (hX + hY) = \exp (hZ)
811 > \end{equation}
812 > where
813 > \begin{equation}
814 > hZ = hX + hY + \frac{{h^2 }}{2}[X,Y] + \frac{{h^3 }}{2}\left(
815 > {[X,[X,Y]] + [Y,[Y,X]]} \right) +  \ldots .
816 > \end{equation}
817 > Here, $[X,Y]$ is the commutators of operator $X$ and $Y$ given by
818 > \[
819 > [X,Y] = XY - YX .
820 > \]
821 > Applying the Baker-Campbell-Hausdorff formula\cite{Varadarajan1974}
822 > to the Sprang splitting, we can obtain
823 > \begin{eqnarray*}
824 > \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 \\
825 >                                   &   & \mbox{} + h^2 [X,X]/8 + h^2 [Y,Y]/8 \\
826 >                                   &   & \mbox{} + h^3 [Y,[Y,X]]/12 - h^3[X,[X,Y]]/24 + \ldots )
827 > \end{eqnarray*}
828 > Since \[ [X,Y] + [Y,X] = 0\] and \[ [X,X] = 0,\] the dominant local
829 > error of Spring splitting is proportional to $h^3$. The same
830 > procedure can be applied to a general splitting,  of the form
831 > \begin{equation}
832 > \varphi _{b_m h}^2  \circ \varphi _{a_m h}^1  \circ \varphi _{b_{m -
833 > 1} h}^2  \circ  \ldots  \circ \varphi _{a_1 h}^1 .
834 > \end{equation}
835 > A careful choice of coefficient $a_1 \ldots b_m$ will lead to higher
836 > order methods. Yoshida proposed an elegant way to compose higher
837 > order methods based on symmetric splitting\cite{Yoshida1990}. Given
838 > a symmetric second order base method $ \varphi _h^{(2)} $, a
839 > fourth-order symmetric method can be constructed by composing,
840 > \[
841 > \varphi _h^{(4)}  = \varphi _{\alpha h}^{(2)}  \circ \varphi _{\beta
842 > h}^{(2)}  \circ \varphi _{\alpha h}^{(2)}
843 > \]
844 > where $ \alpha  =  - \frac{{2^{1/3} }}{{2 - 2^{1/3} }}$ and $ \beta
845 > = \frac{{2^{1/3} }}{{2 - 2^{1/3} }}$. Moreover, a symmetric
846 > integrator $ \varphi _h^{(2n + 2)}$ can be composed by
847 > \begin{equation}
848 > \varphi _h^{(2n + 2)}  = \varphi _{\alpha h}^{(2n)}  \circ \varphi
849 > _{\beta h}^{(2n)}  \circ \varphi _{\alpha h}^{(2n)},
850 > \end{equation}
851 > if the weights are chosen as
852 > \[
853 > \alpha  =  - \frac{{2^{1/(2n + 1)} }}{{2 - 2^{1/(2n + 1)} }},\beta =
854 > \frac{{2^{1/(2n + 1)} }}{{2 - 2^{1/(2n + 1)} }} .
855 > \]
856 >
857 > \section{\label{introSection:molecularDynamics}Molecular Dynamics}
858 >
859 > As one of the principal tools of molecular modeling, Molecular
860 > dynamics has proven to be a powerful tool for studying the functions
861 > of biological systems, providing structural, thermodynamic and
862 > dynamical information. The basic idea of molecular dynamics is that
863 > macroscopic properties are related to microscopic behavior and
864 > microscopic behavior can be calculated from the trajectories in
865 > simulations. For instance, instantaneous temperature of an
866 > Hamiltonian system of $N$ particle can be measured by
867 > \[
868 > T = \sum\limits_{i = 1}^N {\frac{{m_i v_i^2 }}{{fk_B }}}
869 > \]
870 > where $m_i$ and $v_i$ are the mass and velocity of $i$th particle
871 > respectively, $f$ is the number of degrees of freedom, and $k_B$ is
872 > the boltzman constant.
873 >
874 > A typical molecular dynamics run consists of three essential steps:
875 > \begin{enumerate}
876 >  \item Initialization
877 >    \begin{enumerate}
878 >    \item Preliminary preparation
879 >    \item Minimization
880 >    \item Heating
881 >    \item Equilibration
882 >    \end{enumerate}
883 >  \item Production
884 >  \item Analysis
885 > \end{enumerate}
886 > These three individual steps will be covered in the following
887 > sections. Sec.~\ref{introSec:initialSystemSettings} deals with the
888 > initialization of a simulation. Sec.~\ref{introSection:production}
889 > will discusse issues in production run.
890 > Sec.~\ref{introSection:Analysis} provides the theoretical tools for
891 > trajectory analysis.
892 >
893 > \subsection{\label{introSec:initialSystemSettings}Initialization}
894 >
895 > \subsubsection{\textbf{Preliminary preparation}}
896 >
897 > When selecting the starting structure of a molecule for molecular
898 > simulation, one may retrieve its Cartesian coordinates from public
899 > databases, such as RCSB Protein Data Bank \textit{etc}. Although
900 > thousands of crystal structures of molecules are discovered every
901 > year, many more remain unknown due to the difficulties of
902 > purification and crystallization. Even for molecules with known
903 > structure, some important information is missing. For example, a
904 > missing hydrogen atom which acts as donor in hydrogen bonding must
905 > be added. Moreover, in order to include electrostatic interaction,
906 > one may need to specify the partial charges for individual atoms.
907 > Under some circumstances, we may even need to prepare the system in
908 > a special configuration. For instance, when studying transport
909 > phenomenon in membrane systems, we may prepare the lipids in a
910 > bilayer structure instead of placing lipids randomly in solvent,
911 > since we are not interested in the slow self-aggregation process.
912 >
913 > \subsubsection{\textbf{Minimization}}
914 >
915 > It is quite possible that some of molecules in the system from
916 > preliminary preparation may be overlapping with each other. This
917 > close proximity leads to high initial potential energy which
918 > consequently jeopardizes any molecular dynamics simulations. To
919 > remove these steric overlaps, one typically performs energy
920 > minimization to find a more reasonable conformation. Several energy
921 > minimization methods have been developed to exploit the energy
922 > surface and to locate the local minimum. While converging slowly
923 > near the minimum, steepest descent method is extremely robust when
924 > systems are strongly anharmonic. Thus, it is often used to refine
925 > structure from crystallographic data. Relied on the gradient or
926 > hessian, advanced methods like Newton-Raphson converge rapidly to a
927 > local minimum, but become unstable if the energy surface is far from
928 > quadratic. Another factor that must be taken into account, when
929 > choosing energy minimization method, is the size of the system.
930 > Steepest descent and conjugate gradient can deal with models of any
931 > size. Because of the limits on computer memory to store the hessian
932 > matrix and the computing power needed to diagonalized these
933 > matrices, most Newton-Raphson methods can not be used with very
934 > large systems.
935 >
936 > \subsubsection{\textbf{Heating}}
937 >
938 > Typically, Heating is performed by assigning random velocities
939 > according to a Maxwell-Boltzman distribution for a desired
940 > temperature. Beginning at a lower temperature and gradually
941 > increasing the temperature by assigning larger random velocities, we
942 > end up with setting the temperature of the system to a final
943 > temperature at which the simulation will be conducted. In heating
944 > phase, we should also keep the system from drifting or rotating as a
945 > whole. To do this, the net linear momentum and angular momentum of
946 > the system is shifted to zero after each resampling from the Maxwell
947 > -Boltzman distribution.
948 >
949 > \subsubsection{\textbf{Equilibration}}
950 >
951 > The purpose of equilibration is to allow the system to evolve
952 > spontaneously for a period of time and reach equilibrium. The
953 > procedure is continued until various statistical properties, such as
954 > temperature, pressure, energy, volume and other structural
955 > properties \textit{etc}, become independent of time. Strictly
956 > speaking, minimization and heating are not necessary, provided the
957 > equilibration process is long enough. However, these steps can serve
958 > as a means to arrive at an equilibrated structure in an effective
959 > way.
960 >
961 > \subsection{\label{introSection:production}Production}
962 >
963 > The production run is the most important step of the simulation, in
964 > which the equilibrated structure is used as a starting point and the
965 > motions of the molecules are collected for later analysis. In order
966 > to capture the macroscopic properties of the system, the molecular
967 > dynamics simulation must be performed by sampling correctly and
968 > efficiently from the relevant thermodynamic ensemble.
969 >
970 > The most expensive part of a molecular dynamics simulation is the
971 > calculation of non-bonded forces, such as van der Waals force and
972 > Coulombic forces \textit{etc}. For a system of $N$ particles, the
973 > complexity of the algorithm for pair-wise interactions is $O(N^2 )$,
974 > which making large simulations prohibitive in the absence of any
975 > algorithmic tricks.
976 >
977 > A natural approach to avoid system size issues is to represent the
978 > bulk behavior by a finite number of the particles. However, this
979 > approach will suffer from the surface effect at the edges of the
980 > simulation. To offset this, \textit{Periodic boundary conditions}
981 > (see Fig.~\ref{introFig:pbc}) is developed to simulate bulk
982 > properties with a relatively small number of particles. In this
983 > method, the simulation box is replicated throughout space to form an
984 > infinite lattice. During the simulation, when a particle moves in
985 > the primary cell, its image in other cells move in exactly the same
986 > direction with exactly the same orientation. Thus, as a particle
987 > leaves the primary cell, one of its images will enter through the
988 > opposite face.
989 > \begin{figure}
990 > \centering
991 > \includegraphics[width=\linewidth]{pbc.eps}
992 > \caption[An illustration of periodic boundary conditions]{A 2-D
993 > illustration of periodic boundary conditions. As one particle leaves
994 > the left of the simulation box, an image of it enters the right.}
995 > \label{introFig:pbc}
996 > \end{figure}
997 >
998 > %cutoff and minimum image convention
999 > Another important technique to improve the efficiency of force
1000 > evaluation is to apply spherical cutoff where particles farther than
1001 > a predetermined distance are not included in the calculation
1002 > \cite{Frenkel1996}. The use of a cutoff radius will cause a
1003 > discontinuity in the potential energy curve. Fortunately, one can
1004 > shift simple radial potential to ensure the potential curve go
1005 > smoothly to zero at the cutoff radius. The cutoff strategy works
1006 > well for Lennard-Jones interaction because of its short range
1007 > nature. However, simply truncating the electrostatic interaction
1008 > with the use of cutoffs has been shown to lead to severe artifacts
1009 > in simulations. The Ewald summation, in which the slowly decaying
1010 > Coulomb potential is transformed into direct and reciprocal sums
1011 > with rapid and absolute convergence, has proved to minimize the
1012 > periodicity artifacts in liquid simulations. Taking the advantages
1013 > of the fast Fourier transform (FFT) for calculating discrete Fourier
1014 > transforms, the particle mesh-based
1015 > methods\cite{Hockney1981,Shimada1993, Luty1994} are accelerated from
1016 > $O(N^{3/2})$ to $O(N logN)$. An alternative approach is the
1017 > \emph{fast multipole method}\cite{Greengard1987, Greengard1994},
1018 > which treats Coulombic interactions exactly at short range, and
1019 > approximate the potential at long range through multipolar
1020 > expansion. In spite of their wide acceptance at the molecular
1021 > simulation community, these two methods are difficult to implement
1022 > correctly and efficiently. Instead, we use a damped and
1023 > charge-neutralized Coulomb potential method developed by Wolf and
1024 > his coworkers\cite{Wolf1999}. The shifted Coulomb potential for
1025 > particle $i$ and particle $j$ at distance $r_{rj}$ is given by:
1026 > \begin{equation}
1027 > V(r_{ij})= \frac{q_i q_j \textrm{erfc}(\alpha
1028 > r_{ij})}{r_{ij}}-\lim_{r_{ij}\rightarrow
1029 > R_\textrm{c}}\left\{\frac{q_iq_j \textrm{erfc}(\alpha
1030 > r_{ij})}{r_{ij}}\right\}. \label{introEquation:shiftedCoulomb}
1031 > \end{equation}
1032 > where $\alpha$ is the convergence parameter. Due to the lack of
1033 > inherent periodicity and rapid convergence,this method is extremely
1034 > efficient and easy to implement.
1035 > \begin{figure}
1036 > \centering
1037 > \includegraphics[width=\linewidth]{shifted_coulomb.eps}
1038 > \caption[An illustration of shifted Coulomb potential]{An
1039 > illustration of shifted Coulomb potential.}
1040 > \label{introFigure:shiftedCoulomb}
1041 > \end{figure}
1042 >
1043 > %multiple time step
1044 >
1045 > \subsection{\label{introSection:Analysis} Analysis}
1046 >
1047 > Recently, advanced visualization technique have become applied to
1048 > monitor the motions of molecules. Although the dynamics of the
1049 > system can be described qualitatively from animation, quantitative
1050 > trajectory analysis are more useful. According to the principles of
1051 > Statistical Mechanics, Sec.~\ref{introSection:statisticalMechanics},
1052 > one can compute thermodynamic properties, analyze fluctuations of
1053 > structural parameters, and investigate time-dependent processes of
1054 > the molecule from the trajectories.
1055 >
1056 > \subsubsection{\label{introSection:thermodynamicsProperties}\textbf{Thermodynamic Properties}}
1057 >
1058 > Thermodynamic properties, which can be expressed in terms of some
1059 > function of the coordinates and momenta of all particles in the
1060 > system, can be directly computed from molecular dynamics. The usual
1061 > way to measure the pressure is based on virial theorem of Clausius
1062 > which states that the virial is equal to $-3Nk_BT$. For a system
1063 > with forces between particles, the total virial, $W$, contains the
1064 > contribution from external pressure and interaction between the
1065 > particles:
1066 > \[
1067 > W =  - 3PV + \left\langle {\sum\limits_{i < j} {r{}_{ij} \cdot
1068 > f_{ij} } } \right\rangle
1069 > \]
1070 > where $f_{ij}$ is the force between particle $i$ and $j$ at a
1071 > distance $r_{ij}$. Thus, the expression for the pressure is given
1072 > by:
1073 > \begin{equation}
1074 > P = \frac{{Nk_B T}}{V} - \frac{1}{{3V}}\left\langle {\sum\limits_{i
1075 > < j} {r{}_{ij} \cdot f_{ij} } } \right\rangle
1076 > \end{equation}
1077 >
1078 > \subsubsection{\label{introSection:structuralProperties}\textbf{Structural Properties}}
1079 >
1080 > Structural Properties of a simple fluid can be described by a set of
1081 > distribution functions. Among these functions,the \emph{pair
1082 > distribution function}, also known as \emph{radial distribution
1083 > function}, is of most fundamental importance to liquid theory.
1084 > Experimentally, pair distribution function can be gathered by
1085 > Fourier transforming raw data from a series of neutron diffraction
1086 > experiments and integrating over the surface factor
1087 > \cite{Powles1973}. The experimental results can serve as a criterion
1088 > to justify the correctness of a liquid model. Moreover, various
1089 > equilibrium thermodynamic and structural properties can also be
1090 > expressed in terms of radial distribution function \cite{Allen1987}.
1091 >
1092 > The pair distribution functions $g(r)$ gives the probability that a
1093 > particle $i$ will be located at a distance $r$ from a another
1094 > particle $j$ in the system
1095 > \[
1096 > g(r) = \frac{V}{{N^2 }}\left\langle {\sum\limits_i {\sum\limits_{j
1097 > \ne i} {\delta (r - r_{ij} )} } } \right\rangle = \frac{\rho
1098 > (r)}{\rho}.
1099 > \]
1100 > Note that the delta function can be replaced by a histogram in
1101 > computer simulation. Peaks in $g(r)$ represent solvent shells, and
1102 > the height of these peaks gradually decreases to 1 as the liquid of
1103 > large distance approaches the bulk density.
1104 >
1105 >
1106 > \subsubsection{\label{introSection:timeDependentProperties}\textbf{Time-dependent
1107 > Properties}}
1108 >
1109 > Time-dependent properties are usually calculated using \emph{time
1110 > correlation functions}, which correlate random variables $A$ and $B$
1111 > at two different times,
1112 > \begin{equation}
1113 > C_{AB} (t) = \left\langle {A(t)B(0)} \right\rangle.
1114 > \label{introEquation:timeCorrelationFunction}
1115 > \end{equation}
1116 > If $A$ and $B$ refer to same variable, this kind of correlation
1117 > function is called an \emph{autocorrelation function}. One example
1118 > of an auto correlation function is the velocity auto-correlation
1119 > function which is directly related to transport properties of
1120 > molecular liquids:
1121 > \[
1122 > D = \frac{1}{3}\int\limits_0^\infty  {\left\langle {v(t) \cdot v(0)}
1123 > \right\rangle } dt
1124 > \]
1125 > where $D$ is diffusion constant. Unlike the velocity autocorrelation
1126 > function, which is averaging over time origins and over all the
1127 > atoms, the dipole autocorrelation functions are calculated for the
1128 > entire system. The dipole autocorrelation function is given by:
1129 > \[
1130 > c_{dipole}  = \left\langle {u_{tot} (t) \cdot u_{tot} (t)}
1131 > \right\rangle
1132 > \]
1133 > Here $u_{tot}$ is the net dipole of the entire system and is given
1134 > by
1135 > \[
1136 > u_{tot} (t) = \sum\limits_i {u_i (t)}
1137 > \]
1138 > In principle, many time correlation functions can be related with
1139 > Fourier transforms of the infrared, Raman, and inelastic neutron
1140 > scattering spectra of molecular liquids. In practice, one can
1141 > extract the IR spectrum from the intensity of dipole fluctuation at
1142 > each frequency using the following relationship:
1143 > \[
1144 > \hat c_{dipole} (v) = \int_{ - \infty }^\infty  {c_{dipole} (t)e^{ -
1145 > i2\pi vt} dt}
1146 > \]
1147 >
1148 > \section{\label{introSection:rigidBody}Dynamics of Rigid Bodies}
1149 >
1150 > Rigid bodies are frequently involved in the modeling of different
1151 > areas, from engineering, physics, to chemistry. For example,
1152 > missiles and vehicle are usually modeled by rigid bodies.  The
1153 > movement of the objects in 3D gaming engine or other physics
1154 > simulator is governed by rigid body dynamics. In molecular
1155 > simulations, rigid bodies are used to simplify protein-protein
1156 > docking studies\cite{Gray2003}.
1157 >
1158 > It is very important to develop stable and efficient methods to
1159 > integrate the equations of motion for orientational degrees of
1160 > freedom. Euler angles are the natural choice to describe the
1161 > rotational degrees of freedom. However, due to $\frac {1}{sin
1162 > \theta}$ singularities, the numerical integration of corresponding
1163 > equations of motion is very inefficient and inaccurate. Although an
1164 > alternative integrator using multiple sets of Euler angles can
1165 > overcome this difficulty\cite{Barojas1973}, the computational
1166 > penalty and the loss of angular momentum conservation still remain.
1167 > A singularity-free representation utilizing quaternions was
1168 > developed by Evans in 1977\cite{Evans1977}. Unfortunately, this
1169 > approach uses a nonseparable Hamiltonian resulting from the
1170 > quaternion representation, which prevents the symplectic algorithm
1171 > to be utilized. Another different approach is to apply holonomic
1172 > constraints to the atoms belonging to the rigid body. Each atom
1173 > moves independently under the normal forces deriving from potential
1174 > energy and constraint forces which are used to guarantee the
1175 > rigidness. However, due to their iterative nature, the SHAKE and
1176 > Rattle algorithms also converge very slowly when the number of
1177 > constraints increases\cite{Ryckaert1977, Andersen1983}.
1178 >
1179 > A break-through in geometric literature suggests that, in order to
1180 > develop a long-term integration scheme, one should preserve the
1181 > symplectic structure of the flow. By introducing a conjugate
1182 > momentum to the rotation matrix $Q$ and re-formulating Hamiltonian's
1183 > equation, a symplectic integrator, RSHAKE\cite{Kol1997}, was
1184 > proposed to evolve the Hamiltonian system in a constraint manifold
1185 > by iteratively satisfying the orthogonality constraint $Q^T Q = 1$.
1186 > An alternative method using the quaternion representation was
1187 > developed by Omelyan\cite{Omelyan1998}. However, both of these
1188 > methods are iterative and inefficient. In this section, we descibe a
1189 > symplectic Lie-Poisson integrator for rigid body developed by
1190 > Dullweber and his coworkers\cite{Dullweber1997} in depth.
1191 >
1192 > \subsection{\label{introSection:constrainedHamiltonianRB}Constrained Hamiltonian for Rigid Bodies}
1193 > The motion of a rigid body is Hamiltonian with the Hamiltonian
1194 > function
1195 > \begin{equation}
1196 > H = \frac{1}{2}(p^T m^{ - 1} p) + \frac{1}{2}tr(PJ^{ - 1} P) +
1197 > V(q,Q) + \frac{1}{2}tr[(QQ^T  - 1)\Lambda ].
1198 > \label{introEquation:RBHamiltonian}
1199 > \end{equation}
1200 > Here, $q$ and $Q$  are the position and rotation matrix for the
1201 > rigid-body, $p$ and $P$  are conjugate momenta to $q$  and $Q$ , and
1202 > $J$, a diagonal matrix, is defined by
1203 > \[
1204 > I_{ii}^{ - 1}  = \frac{1}{2}\sum\limits_{i \ne j} {J_{jj}^{ - 1} }
1205 > \]
1206 > where $I_{ii}$ is the diagonal element of the inertia tensor. This
1207 > constrained Hamiltonian equation is subjected to a holonomic
1208 > constraint,
1209 > \begin{equation}
1210 > Q^T Q = 1, \label{introEquation:orthogonalConstraint}
1211 > \end{equation}
1212 > which is used to ensure rotation matrix's unitarity. Differentiating
1213 > \ref{introEquation:orthogonalConstraint} and using Equation
1214 > \ref{introEquation:RBMotionMomentum}, one may obtain,
1215 > \begin{equation}
1216 > Q^T PJ^{ - 1}  + J^{ - 1} P^T Q = 0 . \\
1217 > \label{introEquation:RBFirstOrderConstraint}
1218 > \end{equation}
1219 >
1220 > Using Equation (\ref{introEquation:motionHamiltonianCoordinate},
1221 > \ref{introEquation:motionHamiltonianMomentum}), one can write down
1222 > the equations of motion,
1223 >
1224 > \begin{eqnarray}
1225 > \frac{{dq}}{{dt}} & = & \frac{p}{m} \label{introEquation:RBMotionPosition}\\
1226 > \frac{{dp}}{{dt}} & = & - \nabla _q V(q,Q) \label{introEquation:RBMotionMomentum}\\
1227 > \frac{{dQ}}{{dt}} & = & PJ^{ - 1}  \label{introEquation:RBMotionRotation}\\
1228 > \frac{{dP}}{{dt}} & = & - \nabla _Q V(q,Q) - 2Q\Lambda . \label{introEquation:RBMotionP}
1229 > \end{eqnarray}
1230 >
1231 > In general, there are two ways to satisfy the holonomic constraints.
1232 > We can use a constraint force provided by a Lagrange multiplier on
1233 > the normal manifold to keep the motion on constraint space. Or we
1234 > can simply evolve the system on the constraint manifold. These two
1235 > methods have been proved to be equivalent. The holonomic constraint
1236 > and equations of motions define a constraint manifold for rigid
1237 > bodies
1238 > \[
1239 > M = \left\{ {(Q,P):Q^T Q = 1,Q^T PJ^{ - 1}  + J^{ - 1} P^T Q = 0}
1240 > \right\}.
1241 > \]
1242 >
1243 > Unfortunately, this constraint manifold is not the cotangent bundle
1244 > $T^* SO(3)$ which can be consider as a symplectic manifold on Lie
1245 > rotation group $SO(3)$. However, it turns out that under symplectic
1246 > transformation, the cotangent space and the phase space are
1247 > diffeomorphic. By introducing
1248 > \[
1249 > \tilde Q = Q,\tilde P = \frac{1}{2}\left( {P - QP^T Q} \right),
1250 > \]
1251 > the mechanical system subject to a holonomic constraint manifold $M$
1252 > can be re-formulated as a Hamiltonian system on the cotangent space
1253 > \[
1254 > T^* SO(3) = \left\{ {(\tilde Q,\tilde P):\tilde Q^T \tilde Q =
1255 > 1,\tilde Q^T \tilde PJ^{ - 1}  + J^{ - 1} P^T \tilde Q = 0} \right\}
1256 > \]
1257 >
1258 > For a body fixed vector $X_i$ with respect to the center of mass of
1259 > the rigid body, its corresponding lab fixed vector $X_0^{lab}$  is
1260 > given as
1261 > \begin{equation}
1262 > X_i^{lab} = Q X_i + q.
1263 > \end{equation}
1264 > Therefore, potential energy $V(q,Q)$ is defined by
1265 > \[
1266 > V(q,Q) = V(Q X_0 + q).
1267 > \]
1268 > Hence, the force and torque are given by
1269 > \[
1270 > \nabla _q V(q,Q) = F(q,Q) = \sum\limits_i {F_i (q,Q)},
1271 > \]
1272 > and
1273 > \[
1274 > \nabla _Q V(q,Q) = F(q,Q)X_i^t
1275 > \]
1276 > respectively.
1277 >
1278 > As a common choice to describe the rotation dynamics of the rigid
1279 > body, the angular momentum on the body fixed frame $\Pi  = Q^t P$ is
1280 > introduced to rewrite the equations of motion,
1281 > \begin{equation}
1282 > \begin{array}{l}
1283 > \dot \Pi  = J^{ - 1} \Pi ^T \Pi  + Q^T \sum\limits_i {F_i (q,Q)X_i^T }  - \Lambda  \\
1284 > \dot Q  = Q\Pi {\rm{ }}J^{ - 1}  \\
1285 > \end{array}
1286 > \label{introEqaution:RBMotionPI}
1287 > \end{equation}
1288 > , as well as holonomic constraints,
1289 > \[
1290 > \begin{array}{l}
1291 > \Pi J^{ - 1}  + J^{ - 1} \Pi ^t  = 0 \\
1292 > Q^T Q = 1 \\
1293 > \end{array}
1294 > \]
1295 >
1296 > For a vector $v(v_1 ,v_2 ,v_3 ) \in R^3$ and a matrix $\hat v \in
1297 > so(3)^ \star$, the hat-map isomorphism,
1298 > \begin{equation}
1299 > v(v_1 ,v_2 ,v_3 ) \Leftrightarrow \hat v = \left(
1300 > {\begin{array}{*{20}c}
1301 >   0 & { - v_3 } & {v_2 }  \\
1302 >   {v_3 } & 0 & { - v_1 }  \\
1303 >   { - v_2 } & {v_1 } & 0  \\
1304 > \end{array}} \right),
1305 > \label{introEquation:hatmapIsomorphism}
1306 > \end{equation}
1307 > will let us associate the matrix products with traditional vector
1308 > operations
1309 > \[
1310 > \hat vu = v \times u
1311 > \]
1312 > Using \ref{introEqaution:RBMotionPI}, one can construct a skew
1313 > matrix,
1314 >
1315 > \begin{eqnarray*}
1316 > (\dot \Pi  - \dot \Pi ^T ){\rm{ }} = {\rm{ }}(\Pi  - \Pi ^T ){\rm{
1317 > }}(J^{ - 1} \Pi  + \Pi J^{ - 1} ) + \sum\limits_i {[Q^T F_i
1318 > (r,Q)X_i^T  - X_i F_i (r,Q)^T Q]}  - (\Lambda  - \Lambda ^T ).
1319 > \label{introEquation:skewMatrixPI}
1320 > \end{eqnarray*}
1321 >
1322 > Since $\Lambda$ is symmetric, the last term of Equation
1323 > \ref{introEquation:skewMatrixPI} is zero, which implies the Lagrange
1324 > multiplier $\Lambda$ is absent from the equations of motion. This
1325 > unique property eliminates the requirement of iterations which can
1326 > not be avoided in other methods\cite{Kol1997, Omelyan1998}.
1327 >
1328 > Applying the hat-map isomorphism, we obtain the equation of motion
1329 > for angular momentum on body frame
1330 > \begin{equation}
1331 > \dot \pi  = \pi  \times I^{ - 1} \pi  + \sum\limits_i {\left( {Q^T
1332 > F_i (r,Q)} \right) \times X_i }.
1333 > \label{introEquation:bodyAngularMotion}
1334 > \end{equation}
1335 > In the same manner, the equation of motion for rotation matrix is
1336 > given by
1337 > \[
1338 > \dot Q = Qskew(I^{ - 1} \pi )
1339 > \]
1340 >
1341 > \subsection{\label{introSection:SymplecticFreeRB}Symplectic
1342 > Lie-Poisson Integrator for Free Rigid Body}
1343 >
1344 > If there are no external forces exerted on the rigid body, the only
1345 > contribution to the rotational motion is from the kinetic energy
1346 > (the first term of \ref{introEquation:bodyAngularMotion}). The free
1347 > rigid body is an example of a Lie-Poisson system with Hamiltonian
1348 > function
1349 > \begin{equation}
1350 > T^r (\pi ) = T_1 ^r (\pi _1 ) + T_2^r (\pi _2 ) + T_3^r (\pi _3 )
1351 > \label{introEquation:rotationalKineticRB}
1352 > \end{equation}
1353 > where $T_i^r (\pi _i ) = \frac{{\pi _i ^2 }}{{2I_i }}$ and
1354 > Lie-Poisson structure matrix,
1355 > \begin{equation}
1356 > J(\pi ) = \left( {\begin{array}{*{20}c}
1357 >   0 & {\pi _3 } & { - \pi _2 }  \\
1358 >   { - \pi _3 } & 0 & {\pi _1 }  \\
1359 >   {\pi _2 } & { - \pi _1 } & 0  \\
1360 > \end{array}} \right)
1361 > \end{equation}
1362 > Thus, the dynamics of free rigid body is governed by
1363 > \begin{equation}
1364 > \frac{d}{{dt}}\pi  = J(\pi )\nabla _\pi  T^r (\pi )
1365 > \end{equation}
1366 >
1367 > One may notice that each $T_i^r$ in Equation
1368 > \ref{introEquation:rotationalKineticRB} can be solved exactly. For
1369 > instance, the equations of motion due to $T_1^r$ are given by
1370 > \begin{equation}
1371 > \frac{d}{{dt}}\pi  = R_1 \pi ,\frac{d}{{dt}}Q = QR_1
1372 > \label{introEqaution:RBMotionSingleTerm}
1373 > \end{equation}
1374 > where
1375 > \[ R_1  = \left( {\begin{array}{*{20}c}
1376 >   0 & 0 & 0  \\
1377 >   0 & 0 & {\pi _1 }  \\
1378 >   0 & { - \pi _1 } & 0  \\
1379 > \end{array}} \right).
1380 > \]
1381 > The solutions of Equation \ref{introEqaution:RBMotionSingleTerm} is
1382 > \[
1383 > \pi (\Delta t) = e^{\Delta tR_1 } \pi (0),Q(\Delta t) =
1384 > Q(0)e^{\Delta tR_1 }
1385 > \]
1386 > with
1387 > \[
1388 > e^{\Delta tR_1 }  = \left( {\begin{array}{*{20}c}
1389 >   0 & 0 & 0  \\
1390 >   0 & {\cos \theta _1 } & {\sin \theta _1 }  \\
1391 >   0 & { - \sin \theta _1 } & {\cos \theta _1 }  \\
1392 > \end{array}} \right),\theta _1  = \frac{{\pi _1 }}{{I_1 }}\Delta t.
1393 > \]
1394 > To reduce the cost of computing expensive functions in $e^{\Delta
1395 > tR_1 }$, we can use Cayley transformation to obtain a single-aixs
1396 > propagator,
1397 > \[
1398 > e^{\Delta tR_1 }  \approx (1 - \Delta tR_1 )^{ - 1} (1 + \Delta tR_1
1399 > )
1400 > \]
1401 > The flow maps for $T_2^r$ and $T_3^r$ can be found in the same
1402 > manner. In order to construct a second-order symplectic method, we
1403 > split the angular kinetic Hamiltonian function can into five terms
1404 > \[
1405 > T^r (\pi ) = \frac{1}{2}T_1 ^r (\pi _1 ) + \frac{1}{2}T_2^r (\pi _2
1406 > ) + T_3^r (\pi _3 ) + \frac{1}{2}T_2^r (\pi _2 ) + \frac{1}{2}T_1 ^r
1407 > (\pi _1 ).
1408 > \]
1409 > By concatenating the propagators corresponding to these five terms,
1410 > we can obtain an symplectic integrator,
1411 > \[
1412 > \varphi _{\Delta t,T^r }  = \varphi _{\Delta t/2,\pi _1 }  \circ
1413 > \varphi _{\Delta t/2,\pi _2 }  \circ \varphi _{\Delta t,\pi _3 }
1414 > \circ \varphi _{\Delta t/2,\pi _2 }  \circ \varphi _{\Delta t/2,\pi
1415 > _1 }.
1416 > \]
1417 >
1418 > The non-canonical Lie-Poisson bracket ${F, G}$ of two function
1419 > $F(\pi )$ and $G(\pi )$ is defined by
1420 > \[
1421 > \{ F,G\} (\pi ) = [\nabla _\pi  F(\pi )]^T J(\pi )\nabla _\pi  G(\pi
1422 > )
1423 > \]
1424 > If the Poisson bracket of a function $F$ with an arbitrary smooth
1425 > function $G$ is zero, $F$ is a \emph{Casimir}, which is the
1426 > conserved quantity in Poisson system. We can easily verify that the
1427 > norm of the angular momentum, $\parallel \pi
1428 > \parallel$, is a \emph{Casimir}. Let$ F(\pi ) = S(\frac{{\parallel
1429 > \pi \parallel ^2 }}{2})$ for an arbitrary function $ S:R \to R$ ,
1430 > then by the chain rule
1431 > \[
1432 > \nabla _\pi  F(\pi ) = S'(\frac{{\parallel \pi \parallel ^2
1433 > }}{2})\pi
1434 > \]
1435 > Thus $ [\nabla _\pi  F(\pi )]^T J(\pi ) =  - S'(\frac{{\parallel \pi
1436 > \parallel ^2 }}{2})\pi  \times \pi  = 0 $. This explicit
1437 > Lie-Poisson integrator is found to be both extremely efficient and
1438 > stable. These properties can be explained by the fact the small
1439 > angle approximation is used and the norm of the angular momentum is
1440 > conserved.
1441 >
1442 > \subsection{\label{introSection:RBHamiltonianSplitting} Hamiltonian
1443 > Splitting for Rigid Body}
1444 >
1445 > The Hamiltonian of rigid body can be separated in terms of kinetic
1446 > energy and potential energy,
1447 > \[
1448 > H = T(p,\pi ) + V(q,Q)
1449 > \]
1450 > The equations of motion corresponding to potential energy and
1451 > kinetic energy are listed in the below table,
1452 > \begin{table}
1453 > \caption{EQUATIONS OF MOTION DUE TO POTENTIAL AND KINETIC ENERGIES}
1454 > \begin{center}
1455 > \begin{tabular}{|l|l|}
1456 >  \hline
1457 >  % after \\: \hline or \cline{col1-col2} \cline{col3-col4} ...
1458 >  Potential & Kinetic \\
1459 >  $\frac{{dq}}{{dt}} = \frac{p}{m}$ & $\frac{d}{{dt}}q = p$ \\
1460 >  $\frac{d}{{dt}}p =  - \frac{{\partial V}}{{\partial q}}$ & $ \frac{d}{{dt}}p = 0$ \\
1461 >  $\frac{d}{{dt}}Q = 0$ & $ \frac{d}{{dt}}Q = Qskew(I^{ - 1} j)$ \\
1462 >  $ \frac{d}{{dt}}\pi  = \sum\limits_i {\left( {Q^T F_i (r,Q)} \right) \times X_i }$ & $\frac{d}{{dt}}\pi  = \pi  \times I^{ - 1} \pi$\\
1463 >  \hline
1464 > \end{tabular}
1465 > \end{center}
1466 > \end{table}
1467 > A second-order symplectic method is now obtained by the composition
1468 > of the position and velocity propagators,
1469 > \[
1470 > \varphi _{\Delta t}  = \varphi _{\Delta t/2,V}  \circ \varphi
1471 > _{\Delta t,T}  \circ \varphi _{\Delta t/2,V}.
1472 > \]
1473 > Moreover, $\varphi _{\Delta t/2,V}$ can be divided into two
1474 > sub-propagators which corresponding to force and torque
1475 > respectively,
1476 > \[
1477 > \varphi _{\Delta t/2,V}  = \varphi _{\Delta t/2,F}  \circ \varphi
1478 > _{\Delta t/2,\tau }.
1479 > \]
1480 > Since the associated operators of $\varphi _{\Delta t/2,F} $ and
1481 > $\circ \varphi _{\Delta t/2,\tau }$ commute, the composition order
1482 > inside $\varphi _{\Delta t/2,V}$ does not matter. Furthermore, the
1483 > kinetic energy can be separated to translational kinetic term, $T^t
1484 > (p)$, and rotational kinetic term, $T^r (\pi )$,
1485 > \begin{equation}
1486 > T(p,\pi ) =T^t (p) + T^r (\pi ).
1487 > \end{equation}
1488 > where $ T^t (p) = \frac{1}{2}p^T m^{ - 1} p $ and $T^r (\pi )$ is
1489 > defined by \ref{introEquation:rotationalKineticRB}. Therefore, the
1490 > corresponding propagators are given by
1491 > \[
1492 > \varphi _{\Delta t,T}  = \varphi _{\Delta t,T^t }  \circ \varphi
1493 > _{\Delta t,T^r }.
1494 > \]
1495 > Finally, we obtain the overall symplectic propagators for freely
1496 > moving rigid bodies
1497 > \begin{equation}
1498 > \begin{array}{c}
1499 > \varphi _{\Delta t}  = \varphi _{\Delta t/2,F}  \circ \varphi _{\Delta t/2,\tau }  \\
1500 >  \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 }  \\
1501 >  \circ \varphi _{\Delta t/2,\tau }  \circ \varphi _{\Delta t/2,F}  .\\
1502 > \end{array}
1503 > \label{introEquation:overallRBFlowMaps}
1504 > \end{equation}
1505 >
1506 > \section{\label{introSection:langevinDynamics}Langevin Dynamics}
1507 > As an alternative to newtonian dynamics, Langevin dynamics, which
1508 > mimics a simple heat bath with stochastic and dissipative forces,
1509 > has been applied in a variety of studies. This section will review
1510 > the theory of Langevin dynamics. A brief derivation of generalized
1511 > Langevin equation will be given first. Following that, we will
1512 > discuss the physical meaning of the terms appearing in the equation
1513 > as well as the calculation of friction tensor from hydrodynamics
1514 > theory.
1515 >
1516 > \subsection{\label{introSection:generalizedLangevinDynamics}Derivation of Generalized Langevin Equation}
1517 >
1518 > A harmonic bath model, in which an effective set of harmonic
1519 > oscillators are used to mimic the effect of a linearly responding
1520 > environment, has been widely used in quantum chemistry and
1521 > statistical mechanics. One of the successful applications of
1522 > Harmonic bath model is the derivation of the Generalized Langevin
1523 > Dynamics (GLE). Lets consider a system, in which the degree of
1524 > freedom $x$ is assumed to couple to the bath linearly, giving a
1525 > Hamiltonian of the form
1526 > \begin{equation}
1527 > H = \frac{{p^2 }}{{2m}} + U(x) + H_B  + \Delta U(x,x_1 , \ldots x_N)
1528 > \label{introEquation:bathGLE}.
1529 > \end{equation}
1530 > Here $p$ is a momentum conjugate to $x$, $m$ is the mass associated
1531 > with this degree of freedom, $H_B$ is a harmonic bath Hamiltonian,
1532 > \[
1533 > H_B  = \sum\limits_{\alpha  = 1}^N {\left\{ {\frac{{p_\alpha ^2
1534 > }}{{2m_\alpha  }} + \frac{1}{2}m_\alpha  \omega _\alpha ^2 }
1535 > \right\}}
1536 > \]
1537 > where the index $\alpha$ runs over all the bath degrees of freedom,
1538 > $\omega _\alpha$ are the harmonic bath frequencies, $m_\alpha$ are
1539 > the harmonic bath masses, and $\Delta U$ is a bilinear system-bath
1540 > coupling,
1541 > \[
1542 > \Delta U =  - \sum\limits_{\alpha  = 1}^N {g_\alpha  x_\alpha  x}
1543 > \]
1544 > where $g_\alpha$ are the coupling constants between the bath
1545 > coordinates ($x_ \alpha$) and the system coordinate ($x$).
1546 > Introducing
1547 > \[
1548 > W(x) = U(x) - \sum\limits_{\alpha  = 1}^N {\frac{{g_\alpha ^2
1549 > }}{{2m_\alpha  w_\alpha ^2 }}} x^2
1550 > \] and combining the last two terms in Equation
1551 > \ref{introEquation:bathGLE}, we may rewrite the Harmonic bath
1552 > Hamiltonian as
1553 > \[
1554 > H = \frac{{p^2 }}{{2m}} + W(x) + \sum\limits_{\alpha  = 1}^N
1555 > {\left\{ {\frac{{p_\alpha ^2 }}{{2m_\alpha  }} + \frac{1}{2}m_\alpha
1556 > w_\alpha ^2 \left( {x_\alpha   - \frac{{g_\alpha  }}{{m_\alpha
1557 > w_\alpha ^2 }}x} \right)^2 } \right\}}
1558 > \]
1559 > Since the first two terms of the new Hamiltonian depend only on the
1560 > system coordinates, we can get the equations of motion for
1561 > Generalized Langevin Dynamics by Hamilton's equations,
1562 > \begin{equation}
1563 > m\ddot x =  - \frac{{\partial W(x)}}{{\partial x}} -
1564 > \sum\limits_{\alpha  = 1}^N {g_\alpha  \left( {x_\alpha   -
1565 > \frac{{g_\alpha  }}{{m_\alpha  w_\alpha ^2 }}x} \right)},
1566 > \label{introEquation:coorMotionGLE}
1567 > \end{equation}
1568 > and
1569 > \begin{equation}
1570 > m\ddot x_\alpha   =  - m_\alpha  w_\alpha ^2 \left( {x_\alpha   -
1571 > \frac{{g_\alpha  }}{{m_\alpha  w_\alpha ^2 }}x} \right).
1572 > \label{introEquation:bathMotionGLE}
1573 > \end{equation}
1574 >
1575 > In order to derive an equation for $x$, the dynamics of the bath
1576 > variables $x_\alpha$ must be solved exactly first. As an integral
1577 > transform which is particularly useful in solving linear ordinary
1578 > differential equations,the Laplace transform is the appropriate tool
1579 > to solve this problem. The basic idea is to transform the difficult
1580 > differential equations into simple algebra problems which can be
1581 > solved easily. Then, by applying the inverse Laplace transform, also
1582 > known as the Bromwich integral, we can retrieve the solutions of the
1583 > original problems.
1584 >
1585 > Let $f(t)$ be a function defined on $ [0,\infty ) $. The Laplace
1586 > transform of f(t) is a new function defined as
1587 > \[
1588 > L(f(t)) \equiv F(p) = \int_0^\infty  {f(t)e^{ - pt} dt}
1589 > \]
1590 > where  $p$ is real and  $L$ is called the Laplace Transform
1591 > Operator. Below are some important properties of Laplace transform
1592 >
1593 > \begin{eqnarray*}
1594 > L(x + y)  & = & L(x) + L(y) \\
1595 > L(ax)     & = & aL(x) \\
1596 > L(\dot x) & = & pL(x) - px(0) \\
1597 > L(\ddot x)& = & p^2 L(x) - px(0) - \dot x(0) \\
1598 > L\left( {\int_0^t {g(t - \tau )h(\tau )d\tau } } \right)& = & G(p)H(p) \\
1599 > \end{eqnarray*}
1600 >
1601 >
1602 > Applying the Laplace transform to the bath coordinates, we obtain
1603 > \begin{eqnarray*}
1604 > 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) \\
1605 > L(x_\alpha  ) & = & \frac{{\frac{{g_\alpha  }}{{\omega _\alpha  }}L(x) + px_\alpha  (0) + \dot x_\alpha  (0)}}{{p^2  + \omega _\alpha ^2 }} \\
1606 > \end{eqnarray*}
1607 >
1608 > By the same way, the system coordinates become
1609 > \begin{eqnarray*}
1610 > mL(\ddot x) & = & - \frac{1}{p}\frac{{\partial W(x)}}{{\partial x}} \\
1611 >  & & \mbox{} - \sum\limits_{\alpha  = 1}^N {\left\{ { - \frac{{g_\alpha ^2 }}{{m_\alpha  \omega _\alpha ^2 }}\frac{p}{{p^2  + \omega _\alpha ^2 }}pL(x) - \frac{p}{{p^2  + \omega _\alpha ^2 }}g_\alpha  x_\alpha  (0) - \frac{1}{{p^2  + \omega _\alpha ^2 }}g_\alpha  \dot x_\alpha  (0)} \right\}}  \\
1612 > \end{eqnarray*}
1613 >
1614 > With the help of some relatively important inverse Laplace
1615 > transformations:
1616 > \[
1617 > \begin{array}{c}
1618 > L(\cos at) = \frac{p}{{p^2  + a^2 }} \\
1619 > L(\sin at) = \frac{a}{{p^2  + a^2 }} \\
1620 > L(1) = \frac{1}{p} \\
1621 > \end{array}
1622 > \]
1623 > , we obtain
1624 > \begin{eqnarray*}
1625 > m\ddot x & =  & - \frac{{\partial W(x)}}{{\partial x}} -
1626 > \sum\limits_{\alpha  = 1}^N {\left\{ {\left( { - \frac{{g_\alpha ^2
1627 > }}{{m_\alpha  \omega _\alpha ^2 }}} \right)\int_0^t {\cos (\omega
1628 > _\alpha  t)\dot x(t - \tau )d\tau } } \right\}}  \\
1629 > & & + \sum\limits_{\alpha  = 1}^N {\left\{ {\left[ {g_\alpha
1630 > x_\alpha (0) - \frac{{g_\alpha  }}{{m_\alpha  \omega _\alpha  }}}
1631 > \right]\cos (\omega _\alpha  t) + \frac{{g_\alpha  \dot x_\alpha
1632 > (0)}}{{\omega _\alpha  }}\sin (\omega _\alpha  t)} \right\}}
1633 > \end{eqnarray*}
1634 > \begin{eqnarray*}
1635 > m\ddot x & = & - \frac{{\partial W(x)}}{{\partial x}} - \int_0^t
1636 > {\sum\limits_{\alpha  = 1}^N {\left( { - \frac{{g_\alpha ^2
1637 > }}{{m_\alpha  \omega _\alpha ^2 }}} \right)\cos (\omega _\alpha
1638 > t)\dot x(t - \tau )d} \tau }  \\
1639 > & & + \sum\limits_{\alpha  = 1}^N {\left\{ {\left[ {g_\alpha
1640 > x_\alpha (0) - \frac{{g_\alpha }}{{m_\alpha \omega _\alpha  }}}
1641 > \right]\cos (\omega _\alpha  t) + \frac{{g_\alpha  \dot x_\alpha
1642 > (0)}}{{\omega _\alpha  }}\sin (\omega _\alpha  t)} \right\}}
1643 > \end{eqnarray*}
1644 > Introducing a \emph{dynamic friction kernel}
1645 > \begin{equation}
1646 > \xi (t) = \sum\limits_{\alpha  = 1}^N {\left( { - \frac{{g_\alpha ^2
1647 > }}{{m_\alpha  \omega _\alpha ^2 }}} \right)\cos (\omega _\alpha  t)}
1648 > \label{introEquation:dynamicFrictionKernelDefinition}
1649 > \end{equation}
1650 > and \emph{a random force}
1651 > \begin{equation}
1652 > R(t) = \sum\limits_{\alpha  = 1}^N {\left( {g_\alpha  x_\alpha  (0)
1653 > - \frac{{g_\alpha ^2 }}{{m_\alpha  \omega _\alpha ^2 }}x(0)}
1654 > \right)\cos (\omega _\alpha  t)}  + \frac{{\dot x_\alpha
1655 > (0)}}{{\omega _\alpha  }}\sin (\omega _\alpha  t),
1656 > \label{introEquation:randomForceDefinition}
1657 > \end{equation}
1658 > the equation of motion can be rewritten as
1659 > \begin{equation}
1660 > m\ddot x =  - \frac{{\partial W}}{{\partial x}} - \int_0^t {\xi
1661 > (t)\dot x(t - \tau )d\tau }  + R(t)
1662 > \label{introEuqation:GeneralizedLangevinDynamics}
1663 > \end{equation}
1664 > which is known as the \emph{generalized Langevin equation}.
1665 >
1666 > \subsubsection{\label{introSection:randomForceDynamicFrictionKernel}\textbf{Random Force and Dynamic Friction Kernel}}
1667 >
1668 > One may notice that $R(t)$ depends only on initial conditions, which
1669 > implies it is completely deterministic within the context of a
1670 > harmonic bath. However, it is easy to verify that $R(t)$ is totally
1671 > uncorrelated to $x$ and $\dot x$,
1672 > \[
1673 > \begin{array}{l}
1674 > \left\langle {x(t)R(t)} \right\rangle  = 0, \\
1675 > \left\langle {\dot x(t)R(t)} \right\rangle  = 0. \\
1676 > \end{array}
1677 > \]
1678 > This property is what we expect from a truly random process. As long
1679 > as the model chosen for $R(t)$ was a gaussian distribution in
1680 > general, the stochastic nature of the GLE still remains.
1681 >
1682 > %dynamic friction kernel
1683 > The convolution integral
1684 > \[
1685 > \int_0^t {\xi (t)\dot x(t - \tau )d\tau }
1686 > \]
1687 > depends on the entire history of the evolution of $x$, which implies
1688 > that the bath retains memory of previous motions. In other words,
1689 > the bath requires a finite time to respond to change in the motion
1690 > of the system. For a sluggish bath which responds slowly to changes
1691 > in the system coordinate, we may regard $\xi(t)$ as a constant
1692 > $\xi(t) = \Xi_0$. Hence, the convolution integral becomes
1693 > \[
1694 > \int_0^t {\xi (t)\dot x(t - \tau )d\tau }  = \xi _0 (x(t) - x(0))
1695 > \]
1696 > and Equation \ref{introEuqation:GeneralizedLangevinDynamics} becomes
1697 > \[
1698 > m\ddot x =  - \frac{\partial }{{\partial x}}\left( {W(x) +
1699 > \frac{1}{2}\xi _0 (x - x_0 )^2 } \right) + R(t),
1700 > \]
1701 > which can be used to describe the effect of dynamic caging in
1702 > viscous solvents. The other extreme is the bath that responds
1703 > infinitely quickly to motions in the system. Thus, $\xi (t)$ can be
1704 > taken as a $delta$ function in time:
1705 > \[
1706 > \xi (t) = 2\xi _0 \delta (t)
1707 > \]
1708 > Hence, the convolution integral becomes
1709 > \[
1710 > \int_0^t {\xi (t)\dot x(t - \tau )d\tau }  = 2\xi _0 \int_0^t
1711 > {\delta (t)\dot x(t - \tau )d\tau }  = \xi _0 \dot x(t),
1712 > \]
1713 > and Equation \ref{introEuqation:GeneralizedLangevinDynamics} becomes
1714 > \begin{equation}
1715 > m\ddot x =  - \frac{{\partial W(x)}}{{\partial x}} - \xi _0 \dot
1716 > x(t) + R(t) \label{introEquation:LangevinEquation}
1717 > \end{equation}
1718 > which is known as the Langevin equation. The static friction
1719 > coefficient $\xi _0$ can either be calculated from spectral density
1720 > or be determined by Stokes' law for regular shaped particles. A
1721 > briefly review on calculating friction tensor for arbitrary shaped
1722 > particles is given in Sec.~\ref{introSection:frictionTensor}.
1723 >
1724 > \subsubsection{\label{introSection:secondFluctuationDissipation}\textbf{The Second Fluctuation Dissipation Theorem}}
1725 >
1726 > Defining a new set of coordinates,
1727 > \[
1728 > q_\alpha  (t) = x_\alpha  (t) - \frac{1}{{m_\alpha  \omega _\alpha
1729 > ^2 }}x(0)
1730 > \],
1731 > we can rewrite $R(T)$ as
1732 > \[
1733 > R(t) = \sum\limits_{\alpha  = 1}^N {g_\alpha  q_\alpha  (t)}.
1734 > \]
1735 > And since the $q$ coordinates are harmonic oscillators,
1736 >
1737 > \begin{eqnarray*}
1738 > \left\langle {q_\alpha ^2 } \right\rangle  & = & \frac{{kT}}{{m_\alpha  \omega _\alpha ^2 }} \\
1739 > \left\langle {q_\alpha  (t)q_\alpha  (0)} \right\rangle & = & \left\langle {q_\alpha ^2 (0)} \right\rangle \cos (\omega _\alpha  t) \\
1740 > \left\langle {q_\alpha  (t)q_\beta  (0)} \right\rangle & = &\delta _{\alpha \beta } \left\langle {q_\alpha  (t)q_\alpha  (0)} \right\rangle  \\
1741 > \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 } }  \\
1742 >  & = &\sum\limits_\alpha  {g_\alpha ^2 \left\langle {q_\alpha ^2 (0)} \right\rangle \cos (\omega _\alpha  t)}  \\
1743 >  & = &kT\xi (t) \\
1744 > \end{eqnarray*}
1745 >
1746 > Thus, we recover the \emph{second fluctuation dissipation theorem}
1747 > \begin{equation}
1748 > \xi (t) = \left\langle {R(t)R(0)} \right\rangle
1749 > \label{introEquation:secondFluctuationDissipation}.
1750 > \end{equation}
1751 > In effect, it acts as a constraint on the possible ways in which one
1752 > can model the random force and friction kernel.

Diff Legend

Removed lines
+ Added lines
< Changed lines
> Changed lines