ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/tengDissertation/Introduction.tex
Revision: 2703
Committed: Wed Apr 12 04:13:16 2006 UTC (18 years, 2 months ago) by tim
Content type: application/x-tex
File size: 45377 byte(s)
Log Message:
minor fix

File Contents

# Content
1 \chapter{\label{chapt:introduction}INTRODUCTION AND THEORETICAL BACKGROUND}
2
3 \section{\label{introSection:classicalMechanics}Classical
4 Mechanics}
5
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
10 a mechanical system at any time of interest; Secondly, all the
11 mechanical properties of the system at that time can be determined
12 by combining the knowledge of the properties of the system with the
13 specification of this state; Finally, the specification of the state
14 when further combine with the laws of mechanics will also be
15 sufficient to predict the future behavior of the system.
16
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 {mv}{dt}
26 \label{introEquation:newtonSecondLaw}
27 \end{equation}
28 A point mass interacting with other bodies moves with the
29 acceleration along the direction of the force acting on it. Let
30 $F_{ij}$ 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 Conservation laws of Newtonian Mechanics play very important roles
39 in solving mechanics problems. The linear momentum of a particle is
40 conserved if it is free or it experiences no force. The second
41 conservation theorem concerns the angular momentum of a particle.
42 The angular momentum $L$ of a particle with respect to an origin
43 from which $r$ is measured is defined to be
44 \begin{equation}
45 L \equiv r \times p \label{introEquation:angularMomentumDefinition}
46 \end{equation}
47 The torque $\tau$ with respect to the same origin is defined to be
48 \begin{equation}
49 N \equiv r \times F \label{introEquation:torqueDefinition}
50 \end{equation}
51 Differentiating Eq.~\ref{introEquation:angularMomentumDefinition},
52 \[
53 \dot L = \frac{d}{{dt}}(r \times p) = (\dot r \times p) + (r \times
54 \dot p)
55 \]
56 since
57 \[
58 \dot r \times p = \dot r \times mv = m\dot r \times \dot r \equiv 0
59 \]
60 thus,
61 \begin{equation}
62 \dot L = r \times \dot p = N
63 \end{equation}
64 If there are no external torques acting on a body, the angular
65 momentum of it is conserved. The last conservation theorem state
66 that if all forces are conservative, Energy
67 \begin{equation}E = T + V \label{introEquation:energyConservation}
68 \end{equation}
69 is conserved. All of these conserved quantities are
70 important factors to determine the quality of numerical integration
71 scheme for rigid body \cite{Dullweber1997}.
72
73 \subsection{\label{introSection:lagrangian}Lagrangian Mechanics}
74
75 Newtonian Mechanics suffers from two important limitations: it
76 describes their motion in special cartesian coordinate systems.
77 Another limitation of Newtonian mechanics becomes obvious when we
78 try to describe systems with large numbers of particles. It becomes
79 very difficult to predict the properties of the system by carrying
80 out calculations involving the each individual interaction between
81 all the particles, even if we know all of the details of the
82 interaction. In order to overcome some of the practical difficulties
83 which arise in attempts to apply Newton's equation to complex
84 system, alternative procedures may be developed.
85
86 \subsubsection{\label{introSection:halmiltonPrinciple}Hamilton's
87 Principle}
88
89 Hamilton introduced the dynamical principle upon which it is
90 possible to base all of mechanics and, indeed, most of classical
91 physics. Hamilton's Principle may be stated as follow,
92
93 The actual trajectory, along which a dynamical system may move from
94 one point to another within a specified time, is derived by finding
95 the path which minimizes the time integral of the difference between
96 the kinetic, $K$, and potential energies, $U$ \cite{tolman79}.
97 \begin{equation}
98 \delta \int_{t_1 }^{t_2 } {(K - U)dt = 0} ,
99 \label{introEquation:halmitonianPrinciple1}
100 \end{equation}
101
102 For simple mechanical systems, where the forces acting on the
103 different part are derivable from a potential and the velocities are
104 small compared with that of light, the Lagrangian function $L$ can
105 be define as the difference between the kinetic energy of the system
106 and its potential energy,
107 \begin{equation}
108 L \equiv K - U = L(q_i ,\dot q_i ) ,
109 \label{introEquation:lagrangianDef}
110 \end{equation}
111 then Eq.~\ref{introEquation:halmitonianPrinciple1} becomes
112 \begin{equation}
113 \delta \int_{t_1 }^{t_2 } {L dt = 0} ,
114 \label{introEquation:halmitonianPrinciple2}
115 \end{equation}
116
117 \subsubsection{\label{introSection:equationOfMotionLagrangian}The
118 Equations of Motion in Lagrangian Mechanics}
119
120 For a holonomic system of $f$ degrees of freedom, the equations of
121 motion in the Lagrangian form is
122 \begin{equation}
123 \frac{d}{{dt}}\frac{{\partial L}}{{\partial \dot q_i }} -
124 \frac{{\partial L}}{{\partial q_i }} = 0,{\rm{ }}i = 1, \ldots,f
125 \label{introEquation:eqMotionLagrangian}
126 \end{equation}
127 where $q_{i}$ is generalized coordinate and $\dot{q_{i}}$ is
128 generalized velocity.
129
130 \subsection{\label{introSection:hamiltonian}Hamiltonian Mechanics}
131
132 Arising from Lagrangian Mechanics, Hamiltonian Mechanics was
133 introduced by William Rowan Hamilton in 1833 as a re-formulation of
134 classical mechanics. If the potential energy of a system is
135 independent of generalized velocities, the generalized momenta can
136 be defined as
137 \begin{equation}
138 p_i = \frac{\partial L}{\partial \dot q_i}
139 \label{introEquation:generalizedMomenta}
140 \end{equation}
141 The Lagrange equations of motion are then expressed by
142 \begin{equation}
143 p_i = \frac{{\partial L}}{{\partial q_i }}
144 \label{introEquation:generalizedMomentaDot}
145 \end{equation}
146
147 With the help of the generalized momenta, we may now define a new
148 quantity $H$ by the equation
149 \begin{equation}
150 H = \sum\limits_k {p_k \dot q_k } - L ,
151 \label{introEquation:hamiltonianDefByLagrangian}
152 \end{equation}
153 where $ \dot q_1 \ldots \dot q_f $ are generalized velocities and
154 $L$ is the Lagrangian function for the system.
155
156 Differentiating Eq.~\ref{introEquation:hamiltonianDefByLagrangian},
157 one can obtain
158 \begin{equation}
159 dH = \sum\limits_k {\left( {p_k d\dot q_k + \dot q_k dp_k -
160 \frac{{\partial L}}{{\partial q_k }}dq_k - \frac{{\partial
161 L}}{{\partial \dot q_k }}d\dot q_k } \right)} - \frac{{\partial
162 L}}{{\partial t}}dt \label{introEquation:diffHamiltonian1}
163 \end{equation}
164 Making use of Eq.~\ref{introEquation:generalizedMomenta}, the
165 second and fourth terms in the parentheses cancel. Therefore,
166 Eq.~\ref{introEquation:diffHamiltonian1} can be rewritten as
167 \begin{equation}
168 dH = \sum\limits_k {\left( {\dot q_k dp_k - \dot p_k dq_k }
169 \right)} - \frac{{\partial L}}{{\partial t}}dt
170 \label{introEquation:diffHamiltonian2}
171 \end{equation}
172 By identifying the coefficients of $dq_k$, $dp_k$ and dt, we can
173 find
174 \begin{equation}
175 \frac{{\partial H}}{{\partial p_k }} = q_k
176 \label{introEquation:motionHamiltonianCoordinate}
177 \end{equation}
178 \begin{equation}
179 \frac{{\partial H}}{{\partial q_k }} = - p_k
180 \label{introEquation:motionHamiltonianMomentum}
181 \end{equation}
182 and
183 \begin{equation}
184 \frac{{\partial H}}{{\partial t}} = - \frac{{\partial L}}{{\partial
185 t}}
186 \label{introEquation:motionHamiltonianTime}
187 \end{equation}
188
189 Eq.~\ref{introEquation:motionHamiltonianCoordinate} and
190 Eq.~\ref{introEquation:motionHamiltonianMomentum} are Hamilton's
191 equation of motion. Due to their symmetrical formula, they are also
192 known as the canonical equations of motions \cite{Goldstein01}.
193
194 An important difference between Lagrangian approach and the
195 Hamiltonian approach is that the Lagrangian is considered to be a
196 function of the generalized velocities $\dot q_i$ and the
197 generalized coordinates $q_i$, while the Hamiltonian is considered
198 to be a function of the generalized momenta $p_i$ and the conjugate
199 generalized coordinate $q_i$. Hamiltonian Mechanics is more
200 appropriate for application to statistical mechanics and quantum
201 mechanics, since it treats the coordinate and its time derivative as
202 independent variables and it only works with 1st-order differential
203 equations\cite{Marion90}.
204
205 In Newtonian Mechanics, a system described by conservative forces
206 conserves the total energy \ref{introEquation:energyConservation}.
207 It follows that Hamilton's equations of motion conserve the total
208 Hamiltonian.
209 \begin{equation}
210 \frac{{dH}}{{dt}} = \sum\limits_i {\left( {\frac{{\partial
211 H}}{{\partial q_i }}\dot q_i + \frac{{\partial H}}{{\partial p_i
212 }}\dot p_i } \right)} = \sum\limits_i {\left( {\frac{{\partial
213 H}}{{\partial q_i }}\frac{{\partial H}}{{\partial p_i }} -
214 \frac{{\partial H}}{{\partial p_i }}\frac{{\partial H}}{{\partial
215 q_i }}} \right) = 0} \label{introEquation:conserveHalmitonian}
216 \end{equation}
217
218 \section{\label{introSection:statisticalMechanics}Statistical
219 Mechanics}
220
221 The thermodynamic behaviors and properties of Molecular Dynamics
222 simulation are governed by the principle of Statistical Mechanics.
223 The following section will give a brief introduction to some of the
224 Statistical Mechanics concepts and theorem presented in this
225 dissertation.
226
227 \subsection{\label{introSection:ensemble}Phase Space and Ensemble}
228
229 Mathematically, phase space is the space which represents all
230 possible states. Each possible state of the system corresponds to
231 one unique point in the phase space. For mechanical systems, the
232 phase space usually consists of all possible values of position and
233 momentum variables. Consider a dynamic system in a cartesian space,
234 where each of the $6f$ coordinates and momenta is assigned to one of
235 $6f$ mutually orthogonal axes, the phase space of this system is a
236 $6f$ dimensional space. A point, $x = (q_1 , \ldots ,q_f ,p_1 ,
237 \ldots ,p_f )$, with a unique set of values of $6f$ coordinates and
238 momenta is a phase space vector.
239
240 A microscopic state or microstate of a classical system is
241 specification of the complete phase space vector of a system at any
242 instant in time. An ensemble is defined as a collection of systems
243 sharing one or more macroscopic characteristics but each being in a
244 unique microstate. The complete ensemble is specified by giving all
245 systems or microstates consistent with the common macroscopic
246 characteristics of the ensemble. Although the state of each
247 individual system in the ensemble could be precisely described at
248 any instance in time by a suitable phase space vector, when using
249 ensembles for statistical purposes, there is no need to maintain
250 distinctions between individual systems, since the numbers of
251 systems at any time in the different states which correspond to
252 different regions of the phase space are more interesting. Moreover,
253 in the point of view of statistical mechanics, one would prefer to
254 use ensembles containing a large enough population of separate
255 members so that the numbers of systems in such different states can
256 be regarded as changing continuously as we traverse different
257 regions of the phase space. The condition of an ensemble at any time
258 can be regarded as appropriately specified by the density $\rho$
259 with which representative points are distributed over the phase
260 space. The density of distribution for an ensemble with $f$ degrees
261 of freedom is defined as,
262 \begin{equation}
263 \rho = \rho (q_1 , \ldots ,q_f ,p_1 , \ldots ,p_f ,t).
264 \label{introEquation:densityDistribution}
265 \end{equation}
266 Governed by the principles of mechanics, the phase points change
267 their value which would change the density at any time at phase
268 space. Hence, the density of distribution is also to be taken as a
269 function of the time.
270
271 The number of systems $\delta N$ at time $t$ can be determined by,
272 \begin{equation}
273 \delta N = \rho (q,p,t)dq_1 \ldots dq_f dp_1 \ldots dp_f.
274 \label{introEquation:deltaN}
275 \end{equation}
276 Assuming a large enough population of systems are exploited, we can
277 sufficiently approximate $\delta N$ without introducing
278 discontinuity when we go from one region in the phase space to
279 another. By integrating over the whole phase space,
280 \begin{equation}
281 N = \int { \ldots \int {\rho (q,p,t)dq_1 } ...dq_f dp_1 } ...dp_f
282 \label{introEquation:totalNumberSystem}
283 \end{equation}
284 gives us an expression for the total number of the systems. Hence,
285 the probability per unit in the phase space can be obtained by,
286 \begin{equation}
287 \frac{{\rho (q,p,t)}}{N} = \frac{{\rho (q,p,t)}}{{\int { \ldots \int
288 {\rho (q,p,t)dq_1 } ...dq_f dp_1 } ...dp_f }}.
289 \label{introEquation:unitProbability}
290 \end{equation}
291 With the help of Equation(\ref{introEquation:unitProbability}) and
292 the knowledge of the system, it is possible to calculate the average
293 value of any desired quantity which depends on the coordinates and
294 momenta of the system. Even when the dynamics of the real system is
295 complex, or stochastic, or even discontinuous, the average
296 properties of the ensemble of possibilities as a whole may still
297 remain well defined. For a classical system in thermal equilibrium
298 with its environment, the ensemble average of a mechanical quantity,
299 $\langle A(q , p) \rangle_t$, takes the form of an integral over the
300 phase space of the system,
301 \begin{equation}
302 \langle A(q , p) \rangle_t = \frac{{\int { \ldots \int {A(q,p)\rho
303 (q,p,t)dq_1 } ...dq_f dp_1 } ...dp_f }}{{\int { \ldots \int {\rho
304 (q,p,t)dq_1 } ...dq_f dp_1 } ...dp_f }}
305 \label{introEquation:ensembelAverage}
306 \end{equation}
307
308 There are several different types of ensembles with different
309 statistical characteristics. As a function of macroscopic
310 parameters, such as temperature \textit{etc}, partition function can
311 be used to describe the statistical properties of a system in
312 thermodynamic equilibrium.
313
314 As an ensemble of systems, each of which is known to be thermally
315 isolated and conserve energy, Microcanonical ensemble(NVE) has a
316 partition function like,
317 \begin{equation}
318 \Omega (N,V,E) = e^{\beta TS}
319 \label{introEqaution:NVEPartition}.
320 \end{equation}
321 A canonical ensemble(NVT)is an ensemble of systems, each of which
322 can share its energy with a large heat reservoir. The distribution
323 of the total energy amongst the possible dynamical states is given
324 by the partition function,
325 \begin{equation}
326 \Omega (N,V,T) = e^{ - \beta A}
327 \label{introEquation:NVTPartition}
328 \end{equation}
329 Here, $A$ is the Helmholtz free energy which is defined as $ A = U -
330 TS$. Since most experiment are carried out under constant pressure
331 condition, isothermal-isobaric ensemble(NPT) play a very important
332 role in molecular simulation. The isothermal-isobaric ensemble allow
333 the system to exchange energy with a heat bath of temperature $T$
334 and to change the volume as well. Its partition function is given as
335 \begin{equation}
336 \Delta (N,P,T) = - e^{\beta G}.
337 \label{introEquation:NPTPartition}
338 \end{equation}
339 Here, $G = U - TS + PV$, and $G$ is called Gibbs free energy.
340
341 \subsection{\label{introSection:liouville}Liouville's theorem}
342
343 The Liouville's theorem is the foundation on which statistical
344 mechanics rests. It describes the time evolution of phase space
345 distribution function. In order to calculate the rate of change of
346 $\rho$, we begin from Equation(\ref{introEquation:deltaN}). If we
347 consider the two faces perpendicular to the $q_1$ axis, which are
348 located at $q_1$ and $q_1 + \delta q_1$, the number of phase points
349 leaving the opposite face is given by the expression,
350 \begin{equation}
351 \left( {\rho + \frac{{\partial \rho }}{{\partial q_1 }}\delta q_1 }
352 \right)\left( {\dot q_1 + \frac{{\partial \dot q_1 }}{{\partial q_1
353 }}\delta q_1 } \right)\delta q_2 \ldots \delta q_f \delta p_1
354 \ldots \delta p_f .
355 \end{equation}
356 Summing all over the phase space, we obtain
357 \begin{equation}
358 \frac{{d(\delta N)}}{{dt}} = - \sum\limits_{i = 1}^f {\left[ {\rho
359 \left( {\frac{{\partial \dot q_i }}{{\partial q_i }} +
360 \frac{{\partial \dot p_i }}{{\partial p_i }}} \right) + \left(
361 {\frac{{\partial \rho }}{{\partial q_i }}\dot q_i + \frac{{\partial
362 \rho }}{{\partial p_i }}\dot p_i } \right)} \right]} \delta q_1
363 \ldots \delta q_f \delta p_1 \ldots \delta p_f .
364 \end{equation}
365 Differentiating the equations of motion in Hamiltonian formalism
366 (\ref{introEquation:motionHamiltonianCoordinate},
367 \ref{introEquation:motionHamiltonianMomentum}), we can show,
368 \begin{equation}
369 \sum\limits_i {\left( {\frac{{\partial \dot q_i }}{{\partial q_i }}
370 + \frac{{\partial \dot p_i }}{{\partial p_i }}} \right)} = 0 ,
371 \end{equation}
372 which cancels the first terms of the right hand side. Furthermore,
373 divining $ \delta q_1 \ldots \delta q_f \delta p_1 \ldots \delta
374 p_f $ in both sides, we can write out Liouville's theorem in a
375 simple form,
376 \begin{equation}
377 \frac{{\partial \rho }}{{\partial t}} + \sum\limits_{i = 1}^f
378 {\left( {\frac{{\partial \rho }}{{\partial q_i }}\dot q_i +
379 \frac{{\partial \rho }}{{\partial p_i }}\dot p_i } \right)} = 0 .
380 \label{introEquation:liouvilleTheorem}
381 \end{equation}
382
383 Liouville's theorem states that the distribution function is
384 constant along any trajectory in phase space. In classical
385 statistical mechanics, since the number of particles in the system
386 is huge, we may be able to believe the system is stationary,
387 \begin{equation}
388 \frac{{\partial \rho }}{{\partial t}} = 0.
389 \label{introEquation:stationary}
390 \end{equation}
391 In such stationary system, the density of distribution $\rho$ can be
392 connected to the Hamiltonian $H$ through Maxwell-Boltzmann
393 distribution,
394 \begin{equation}
395 \rho \propto e^{ - \beta H}
396 \label{introEquation:densityAndHamiltonian}
397 \end{equation}
398
399 \subsubsection{\label{introSection:phaseSpaceConservation}Conservation of Phase Space}
400 Lets consider a region in the phase space,
401 \begin{equation}
402 \delta v = \int { \ldots \int {dq_1 } ...dq_f dp_1 } ..dp_f .
403 \end{equation}
404 If this region is small enough, the density $\rho$ can be regarded
405 as uniform over the whole phase space. Thus, the number of phase
406 points inside this region is given by,
407 \begin{equation}
408 \delta N = \rho \delta v = \rho \int { \ldots \int {dq_1 } ...dq_f
409 dp_1 } ..dp_f.
410 \end{equation}
411
412 \begin{equation}
413 \frac{{d(\delta N)}}{{dt}} = \frac{{d\rho }}{{dt}}\delta v + \rho
414 \frac{d}{{dt}}(\delta v) = 0.
415 \end{equation}
416 With the help of stationary assumption
417 (\ref{introEquation:stationary}), we obtain the principle of the
418 \emph{conservation of extension in phase space},
419 \begin{equation}
420 \frac{d}{{dt}}(\delta v) = \frac{d}{{dt}}\int { \ldots \int {dq_1 }
421 ...dq_f dp_1 } ..dp_f = 0.
422 \label{introEquation:volumePreserving}
423 \end{equation}
424
425 \subsubsection{\label{introSection:liouvilleInOtherForms}Liouville's Theorem in Other Forms}
426
427 Liouville's theorem can be expresses in a variety of different forms
428 which are convenient within different contexts. For any two function
429 $F$ and $G$ of the coordinates and momenta of a system, the Poisson
430 bracket ${F, G}$ is defined as
431 \begin{equation}
432 \left\{ {F,G} \right\} = \sum\limits_i {\left( {\frac{{\partial
433 F}}{{\partial q_i }}\frac{{\partial G}}{{\partial p_i }} -
434 \frac{{\partial F}}{{\partial p_i }}\frac{{\partial G}}{{\partial
435 q_i }}} \right)}.
436 \label{introEquation:poissonBracket}
437 \end{equation}
438 Substituting equations of motion in Hamiltonian formalism(
439 \ref{introEquation:motionHamiltonianCoordinate} ,
440 \ref{introEquation:motionHamiltonianMomentum} ) into
441 (\ref{introEquation:liouvilleTheorem}), we can rewrite Liouville's
442 theorem using Poisson bracket notion,
443 \begin{equation}
444 \left( {\frac{{\partial \rho }}{{\partial t}}} \right) = - \left\{
445 {\rho ,H} \right\}.
446 \label{introEquation:liouvilleTheromInPoissin}
447 \end{equation}
448 Moreover, the Liouville operator is defined as
449 \begin{equation}
450 iL = \sum\limits_{i = 1}^f {\left( {\frac{{\partial H}}{{\partial
451 p_i }}\frac{\partial }{{\partial q_i }} - \frac{{\partial
452 H}}{{\partial q_i }}\frac{\partial }{{\partial p_i }}} \right)}
453 \label{introEquation:liouvilleOperator}
454 \end{equation}
455 In terms of Liouville operator, Liouville's equation can also be
456 expressed as
457 \begin{equation}
458 \left( {\frac{{\partial \rho }}{{\partial t}}} \right) = - iL\rho
459 \label{introEquation:liouvilleTheoremInOperator}
460 \end{equation}
461
462 \subsection{\label{introSection:ergodic}The Ergodic Hypothesis}
463
464 Various thermodynamic properties can be calculated from Molecular
465 Dynamics simulation. By comparing experimental values with the
466 calculated properties, one can determine the accuracy of the
467 simulation and the quality of the underlying model. However, both of
468 experiment and computer simulation are usually performed during a
469 certain time interval and the measurements are averaged over a
470 period of them which is different from the average behavior of
471 many-body system in Statistical Mechanics. Fortunately, Ergodic
472 Hypothesis is proposed to make a connection between time average and
473 ensemble average. It states that time average and average over the
474 statistical ensemble are identical \cite{Frenkel1996, leach01:mm}.
475 \begin{equation}
476 \langle A(q , p) \rangle_t = \mathop {\lim }\limits_{t \to \infty }
477 \frac{1}{t}\int\limits_0^t {A(q(t),p(t))dt = \int\limits_\Gamma
478 {A(q(t),p(t))} } \rho (q(t), p(t)) dqdp
479 \end{equation}
480 where $\langle A(q , p) \rangle_t$ is an equilibrium value of a
481 physical quantity and $\rho (p(t), q(t))$ is the equilibrium
482 distribution function. If an observation is averaged over a
483 sufficiently long time (longer than relaxation time), all accessible
484 microstates in phase space are assumed to be equally probed, giving
485 a properly weighted statistical average. This allows the researcher
486 freedom of choice when deciding how best to measure a given
487 observable. In case an ensemble averaged approach sounds most
488 reasonable, the Monte Carlo techniques\cite{metropolis:1949} can be
489 utilized. Or if the system lends itself to a time averaging
490 approach, the Molecular Dynamics techniques in
491 Sec.~\ref{introSection:molecularDynamics} will be the best
492 choice\cite{Frenkel1996}.
493
494 \section{\label{introSection:geometricIntegratos}Geometric Integrators}
495 A variety of numerical integrators were proposed to simulate the
496 motions. They usually begin with an initial conditionals and move
497 the objects in the direction governed by the differential equations.
498 However, most of them ignore the hidden physical law contained
499 within the equations. Since 1990, geometric integrators, which
500 preserve various phase-flow invariants such as symplectic structure,
501 volume and time reversal symmetry, are developed to address this
502 issue. The velocity verlet method, which happens to be a simple
503 example of symplectic integrator, continues to gain its popularity
504 in molecular dynamics community. This fact can be partly explained
505 by its geometric nature.
506
507 \subsection{\label{introSection:symplecticManifold}Symplectic Manifold}
508 A \emph{manifold} is an abstract mathematical space. It locally
509 looks like Euclidean space, but when viewed globally, it may have
510 more complicate structure. A good example of manifold is the surface
511 of Earth. It seems to be flat locally, but it is round if viewed as
512 a whole. A \emph{differentiable manifold} (also known as
513 \emph{smooth manifold}) is a manifold with an open cover in which
514 the covering neighborhoods are all smoothly isomorphic to one
515 another. In other words,it is possible to apply calculus on
516 \emph{differentiable manifold}. A \emph{symplectic manifold} is
517 defined as a pair $(M, \omega)$ which consisting of a
518 \emph{differentiable manifold} $M$ and a close, non-degenerated,
519 bilinear symplectic form, $\omega$. A symplectic form on a vector
520 space $V$ is a function $\omega(x, y)$ which satisfies
521 $\omega(\lambda_1x_1+\lambda_2x_2, y) = \lambda_1\omega(x_1, y)+
522 \lambda_2\omega(x_2, y)$, $\omega(x, y) = - \omega(y, x)$ and
523 $\omega(x, x) = 0$. Cross product operation in vector field is an
524 example of symplectic form.
525
526 One of the motivations to study \emph{symplectic manifold} in
527 Hamiltonian Mechanics is that a symplectic manifold can represent
528 all possible configurations of the system and the phase space of the
529 system can be described by it's cotangent bundle. Every symplectic
530 manifold is even dimensional. For instance, in Hamilton equations,
531 coordinate and momentum always appear in pairs.
532
533 Let $(M,\omega)$ and $(N, \eta)$ be symplectic manifolds. A map
534 \[
535 f : M \rightarrow N
536 \]
537 is a \emph{symplectomorphism} if it is a \emph{diffeomorphims} and
538 the \emph{pullback} of $\eta$ under f is equal to $\omega$.
539 Canonical transformation is an example of symplectomorphism in
540 classical mechanics.
541
542 \subsection{\label{introSection:ODE}Ordinary Differential Equations}
543
544 For a ordinary differential system defined as
545 \begin{equation}
546 \dot x = f(x)
547 \end{equation}
548 where $x = x(q,p)^T$, this system is canonical Hamiltonian, if
549 \begin{equation}
550 f(r) = J\nabla _x H(r).
551 \end{equation}
552 $H = H (q, p)$ is Hamiltonian function and $J$ is the skew-symmetric
553 matrix
554 \begin{equation}
555 J = \left( {\begin{array}{*{20}c}
556 0 & I \\
557 { - I} & 0 \\
558 \end{array}} \right)
559 \label{introEquation:canonicalMatrix}
560 \end{equation}
561 where $I$ is an identity matrix. Using this notation, Hamiltonian
562 system can be rewritten as,
563 \begin{equation}
564 \frac{d}{{dt}}x = J\nabla _x H(x)
565 \label{introEquation:compactHamiltonian}
566 \end{equation}In this case, $f$ is
567 called a \emph{Hamiltonian vector field}.
568
569 Another generalization of Hamiltonian dynamics is Poisson Dynamics,
570 \begin{equation}
571 \dot x = J(x)\nabla _x H \label{introEquation:poissonHamiltonian}
572 \end{equation}
573 The most obvious change being that matrix $J$ now depends on $x$.
574 The free rigid body is an example of Poisson system (actually a
575 Lie-Poisson system) with Hamiltonian function of angular kinetic
576 energy.
577 \begin{equation}
578 J(\pi ) = \left( {\begin{array}{*{20}c}
579 0 & {\pi _3 } & { - \pi _2 } \\
580 { - \pi _3 } & 0 & {\pi _1 } \\
581 {\pi _2 } & { - \pi _1 } & 0 \\
582 \end{array}} \right)
583 \end{equation}
584
585 \begin{equation}
586 H = \frac{1}{2}\left( {\frac{{\pi _1^2 }}{{I_1 }} + \frac{{\pi _2^2
587 }}{{I_2 }} + \frac{{\pi _3^2 }}{{I_3 }}} \right)
588 \end{equation}
589
590 \subsection{\label{introSection:exactFlow}Exact Flow}
591
592 Let $x(t)$ be the exact solution of the ODE system,
593 \begin{equation}
594 \frac{{dx}}{{dt}} = f(x) \label{introEquation:ODE}
595 \end{equation}
596 The exact flow(solution) $\varphi_\tau$ is defined by
597 \[
598 x(t+\tau) =\varphi_\tau(x(t))
599 \]
600 where $\tau$ is a fixed time step and $\varphi$ is a map from phase
601 space to itself. The flow has the continuous group property,
602 \begin{equation}
603 \varphi _{\tau _1 } \circ \varphi _{\tau _2 } = \varphi _{\tau _1
604 + \tau _2 } .
605 \end{equation}
606 In particular,
607 \begin{equation}
608 \varphi _\tau \circ \varphi _{ - \tau } = I
609 \end{equation}
610 Therefore, the exact flow is self-adjoint,
611 \begin{equation}
612 \varphi _\tau = \varphi _{ - \tau }^{ - 1}.
613 \end{equation}
614 The exact flow can also be written in terms of the of an operator,
615 \begin{equation}
616 \varphi _\tau (x) = e^{\tau \sum\limits_i {f_i (x)\frac{\partial
617 }{{\partial x_i }}} } (x) \equiv \exp (\tau f)(x).
618 \label{introEquation:exponentialOperator}
619 \end{equation}
620
621 In most cases, it is not easy to find the exact flow $\varphi_\tau$.
622 Instead, we use a approximate map, $\psi_\tau$, which is usually
623 called integrator. The order of an integrator $\psi_\tau$ is $p$, if
624 the Taylor series of $\psi_\tau$ agree to order $p$,
625 \begin{equation}
626 \psi_tau(x) = x + \tau f(x) + O(\tau^{p+1})
627 \end{equation}
628
629 \subsection{\label{introSection:geometricProperties}Geometric Properties}
630
631 The hidden geometric properties of ODE and its flow play important
632 roles in numerical studies. Many of them can be found in systems
633 which occur naturally in applications.
634
635 Let $\varphi$ be the flow of Hamiltonian vector field, $\varphi$ is
636 a \emph{symplectic} flow if it satisfies,
637 \begin{equation}
638 {\varphi '}^T J \varphi ' = J.
639 \end{equation}
640 According to Liouville's theorem, the symplectic volume is invariant
641 under a Hamiltonian flow, which is the basis for classical
642 statistical mechanics. Furthermore, the flow of a Hamiltonian vector
643 field on a symplectic manifold can be shown to be a
644 symplectomorphism. As to the Poisson system,
645 \begin{equation}
646 {\varphi '}^T J \varphi ' = J \circ \varphi
647 \end{equation}
648 is the property must be preserved by the integrator.
649
650 It is possible to construct a \emph{volume-preserving} flow for a
651 source free($ \nabla \cdot f = 0 $) ODE, if the flow satisfies $
652 \det d\varphi = 1$. One can show easily that a symplectic flow will
653 be volume-preserving.
654
655 Changing the variables $y = h(x)$ in a ODE\ref{introEquation:ODE}
656 will result in a new system,
657 \[
658 \dot y = \tilde f(y) = ((dh \cdot f)h^{ - 1} )(y).
659 \]
660 The vector filed $f$ has reversing symmetry $h$ if $f = - \tilde f$.
661 In other words, the flow of this vector field is reversible if and
662 only if $ h \circ \varphi ^{ - 1} = \varphi \circ h $.
663
664 When designing any numerical methods, one should always try to
665 preserve the structural properties of the original ODE and its flow.
666
667 \subsection{\label{introSection:constructionSymplectic}Construction of Symplectic Methods}
668 A lot of well established and very effective numerical methods have
669 been successful precisely because of their symplecticities even
670 though this fact was not recognized when they were first
671 constructed. The most famous example is leapfrog methods in
672 molecular dynamics. In general, symplectic integrators can be
673 constructed using one of four different methods.
674 \begin{enumerate}
675 \item Generating functions
676 \item Variational methods
677 \item Runge-Kutta methods
678 \item Splitting methods
679 \end{enumerate}
680
681 Generating function tends to lead to methods which are cumbersome
682 and difficult to use. In dissipative systems, variational methods
683 can capture the decay of energy accurately. Since their
684 geometrically unstable nature against non-Hamiltonian perturbations,
685 ordinary implicit Runge-Kutta methods are not suitable for
686 Hamiltonian system. Recently, various high-order explicit
687 Runge--Kutta methods have been developed to overcome this
688 instability. However, due to computational penalty involved in
689 implementing the Runge-Kutta methods, they do not attract too much
690 attention from Molecular Dynamics community. Instead, splitting have
691 been widely accepted since they exploit natural decompositions of
692 the system\cite{Tuckerman92}.
693
694 \subsubsection{\label{introSection:splittingMethod}Splitting Method}
695
696 The main idea behind splitting methods is to decompose the discrete
697 $\varphi_h$ as a composition of simpler flows,
698 \begin{equation}
699 \varphi _h = \varphi _{h_1 } \circ \varphi _{h_2 } \ldots \circ
700 \varphi _{h_n }
701 \label{introEquation:FlowDecomposition}
702 \end{equation}
703 where each of the sub-flow is chosen such that each represent a
704 simpler integration of the system.
705
706 Suppose that a Hamiltonian system takes the form,
707 \[
708 H = H_1 + H_2.
709 \]
710 Here, $H_1$ and $H_2$ may represent different physical processes of
711 the system. For instance, they may relate to kinetic and potential
712 energy respectively, which is a natural decomposition of the
713 problem. If $H_1$ and $H_2$ can be integrated using exact flows
714 $\varphi_1(t)$ and $\varphi_2(t)$, respectively, a simple first
715 order is then given by the Lie-Trotter formula
716 \begin{equation}
717 \varphi _h = \varphi _{1,h} \circ \varphi _{2,h},
718 \label{introEquation:firstOrderSplitting}
719 \end{equation}
720 where $\varphi _h$ is the result of applying the corresponding
721 continuous $\varphi _i$ over a time $h$. By definition, as
722 $\varphi_i(t)$ is the exact solution of a Hamiltonian system, it
723 must follow that each operator $\varphi_i(t)$ is a symplectic map.
724 It is easy to show that any composition of symplectic flows yields a
725 symplectic map,
726 \begin{equation}
727 (\varphi '\phi ')^T J\varphi '\phi ' = \phi '^T \varphi '^T J\varphi
728 '\phi ' = \phi '^T J\phi ' = J,
729 \label{introEquation:SymplecticFlowComposition}
730 \end{equation}
731 where $\phi$ and $\psi$ both are symplectic maps. Thus operator
732 splitting in this context automatically generates a symplectic map.
733
734 The Lie-Trotter splitting(\ref{introEquation:firstOrderSplitting})
735 introduces local errors proportional to $h^2$, while Strang
736 splitting gives a second-order decomposition,
737 \begin{equation}
738 \varphi _h = \varphi _{1,h/2} \circ \varphi _{2,h} \circ \varphi
739 _{1,h/2} ,
740 \label{introEqaution:secondOrderSplitting}
741 \end{equation}
742 which has a local error proportional to $h^3$. Sprang splitting's
743 popularity in molecular simulation community attribute to its
744 symmetric property,
745 \begin{equation}
746 \varphi _h^{ - 1} = \varphi _{ - h}.
747 \label{introEquation:timeReversible}
748 \end{equation}
749
750 \subsubsection{\label{introSection:exampleSplittingMethod}Example of Splitting Method}
751 The classical equation for a system consisting of interacting
752 particles can be written in Hamiltonian form,
753 \[
754 H = T + V
755 \]
756 where $T$ is the kinetic energy and $V$ is the potential energy.
757 Setting $H_1 = T, H_2 = V$ and applying Strang splitting, one
758 obtains the following:
759 \begin{align}
760 q(\Delta t) &= q(0) + \dot{q}(0)\Delta t +
761 \frac{F[q(0)]}{m}\frac{\Delta t^2}{2}, %
762 \label{introEquation:Lp10a} \\%
763 %
764 \dot{q}(\Delta t) &= \dot{q}(0) + \frac{\Delta t}{2m}
765 \biggl [F[q(0)] + F[q(\Delta t)] \biggr]. %
766 \label{introEquation:Lp10b}
767 \end{align}
768 where $F(t)$ is the force at time $t$. This integration scheme is
769 known as \emph{velocity verlet} which is
770 symplectic(\ref{introEquation:SymplecticFlowComposition}),
771 time-reversible(\ref{introEquation:timeReversible}) and
772 volume-preserving (\ref{introEquation:volumePreserving}). These
773 geometric properties attribute to its long-time stability and its
774 popularity in the community. However, the most commonly used
775 velocity verlet integration scheme is written as below,
776 \begin{align}
777 \dot{q}\biggl (\frac{\Delta t}{2}\biggr ) &=
778 \dot{q}(0) + \frac{\Delta t}{2m}\, F[q(0)], \label{introEquation:Lp9a}\\%
779 %
780 q(\Delta t) &= q(0) + \Delta t\, \dot{q}\biggl (\frac{\Delta t}{2}\biggr ),%
781 \label{introEquation:Lp9b}\\%
782 %
783 \dot{q}(\Delta t) &= \dot{q}\biggl (\frac{\Delta t}{2}\biggr ) +
784 \frac{\Delta t}{2m}\, F[q(0)]. \label{introEquation:Lp9c}
785 \end{align}
786 From the preceding splitting, one can see that the integration of
787 the equations of motion would follow:
788 \begin{enumerate}
789 \item calculate the velocities at the half step, $\frac{\Delta t}{2}$, from the forces calculated at the initial position.
790
791 \item Use the half step velocities to move positions one whole step, $\Delta t$.
792
793 \item Evaluate the forces at the new positions, $\mathbf{r}(\Delta t)$, and use the new forces to complete the velocity move.
794
795 \item Repeat from step 1 with the new position, velocities, and forces assuming the roles of the initial values.
796 \end{enumerate}
797
798 Simply switching the order of splitting and composing, a new
799 integrator, the \emph{position verlet} integrator, can be generated,
800 \begin{align}
801 \dot q(\Delta t) &= \dot q(0) + \Delta tF(q(0))\left[ {q(0) +
802 \frac{{\Delta t}}{{2m}}\dot q(0)} \right], %
803 \label{introEquation:positionVerlet1} \\%
804 %
805 q(\Delta t) &= q(0) + \frac{{\Delta t}}{2}\left[ {\dot q(0) + \dot
806 q(\Delta t)} \right]. %
807 \label{introEquation:positionVerlet1}
808 \end{align}
809
810 \subsubsection{\label{introSection:errorAnalysis}Error Analysis and Higher Order Methods}
811
812 Baker-Campbell-Hausdorff formula can be used to determine the local
813 error of splitting method in terms of commutator of the
814 operators(\ref{introEquation:exponentialOperator}) associated with
815 the sub-flow. For operators $hX$ and $hY$ which are associate to
816 $\varphi_1(t)$ and $\varphi_2(t$ respectively , we have
817 \begin{equation}
818 \exp (hX + hY) = \exp (hZ)
819 \end{equation}
820 where
821 \begin{equation}
822 hZ = hX + hY + \frac{{h^2 }}{2}[X,Y] + \frac{{h^3 }}{2}\left(
823 {[X,[X,Y]] + [Y,[Y,X]]} \right) + \ldots .
824 \end{equation}
825 Here, $[X,Y]$ is the commutators of operator $X$ and $Y$ given by
826 \[
827 [X,Y] = XY - YX .
828 \]
829 Applying Baker-Campbell-Hausdorff formula to Sprang splitting, we
830 can obtain
831 \begin{eqnarray*}
832 \exp (h X/2)\exp (h Y)\exp (h X/2) & = & \exp (h X + h Y + h^2
833 [X,Y]/4 + h^2 [Y,X]/4 \\ & & \mbox{} + h^2 [X,X]/8 + h^2 [Y,Y]/8 \\
834 & & \mbox{} + h^3 [Y,[Y,X]]/12 - h^3 [X,[X,Y]]/24 & & \mbox{} +
835 \ldots )
836 \end{eqnarray*}
837 Since \[ [X,Y] + [Y,X] = 0\] and \[ [X,X] = 0\], the dominant local
838 error of Spring splitting is proportional to $h^3$. The same
839 procedure can be applied to general splitting, of the form
840 \begin{equation}
841 \varphi _{b_m h}^2 \circ \varphi _{a_m h}^1 \circ \varphi _{b_{m -
842 1} h}^2 \circ \ldots \circ \varphi _{a_1 h}^1 .
843 \end{equation}
844 Careful choice of coefficient $a_1 ,\ldot , b_m$ will lead to higher
845 order method. Yoshida proposed an elegant way to compose higher
846 order methods based on symmetric splitting. Given a symmetric second
847 order base method $ \varphi _h^{(2)} $, a fourth-order symmetric
848 method can be constructed by composing,
849 \[
850 \varphi _h^{(4)} = \varphi _{\alpha h}^{(2)} \circ \varphi _{\beta
851 h}^{(2)} \circ \varphi _{\alpha h}^{(2)}
852 \]
853 where $ \alpha = - \frac{{2^{1/3} }}{{2 - 2^{1/3} }}$ and $ \beta
854 = \frac{{2^{1/3} }}{{2 - 2^{1/3} }}$. Moreover, a symmetric
855 integrator $ \varphi _h^{(2n + 2)}$ can be composed by
856 \begin{equation}
857 \varphi _h^{(2n + 2)} = \varphi _{\alpha h}^{(2n)} \circ \varphi
858 _{\beta h}^{(2n)} \circ \varphi _{\alpha h}^{(2n)}
859 \end{equation}
860 , if the weights are chosen as
861 \[
862 \alpha = - \frac{{2^{1/(2n + 1)} }}{{2 - 2^{1/(2n + 1)} }},\beta =
863 \frac{{2^{1/(2n + 1)} }}{{2 - 2^{1/(2n + 1)} }} .
864 \]
865
866 \section{\label{introSection:molecularDynamics}Molecular Dynamics}
867
868 As a special discipline of molecular modeling, Molecular dynamics
869 has proven to be a powerful tool for studying the functions of
870 biological systems, providing structural, thermodynamic and
871 dynamical information.
872
873 \subsection{\label{introSec:mdInit}Initialization}
874
875 \subsection{\label{introSection:mdIntegration} Integration of the Equations of Motion}
876
877 \section{\label{introSection:rigidBody}Dynamics of Rigid Bodies}
878
879 A rigid body is a body in which the distance between any two given
880 points of a rigid body remains constant regardless of external
881 forces exerted on it. A rigid body therefore conserves its shape
882 during its motion.
883
884 Applications of dynamics of rigid bodies.
885
886 \subsection{\label{introSection:lieAlgebra}Lie Algebra}
887
888 \subsection{\label{introSection:DLMMotionEquation}The Euler Equations of Rigid Body Motion}
889
890 \subsection{\label{introSection:otherRBMotionEquation}Other Formulations for Rigid Body Motion}
891
892 \section{\label{introSection:correlationFunctions}Correlation Functions}
893
894 \section{\label{introSection:langevinDynamics}Langevin Dynamics}
895
896 \subsection{\label{introSection:LDIntroduction}Introduction and application of Langevin Dynamics}
897
898 \subsection{\label{introSection:generalizedLangevinDynamics}Generalized Langevin Dynamics}
899
900 \begin{equation}
901 H = \frac{{p^2 }}{{2m}} + U(x) + H_B + \Delta U(x,x_1 , \ldots x_N)
902 \label{introEquation:bathGLE}
903 \end{equation}
904 where $H_B$ is harmonic bath Hamiltonian,
905 \[
906 H_B =\sum\limits_{\alpha = 1}^N {\left\{ {\frac{{p_\alpha ^2
907 }}{{2m_\alpha }} + \frac{1}{2}m_\alpha w_\alpha ^2 } \right\}}
908 \]
909 and $\Delta U$ is bilinear system-bath coupling,
910 \[
911 \Delta U = - \sum\limits_{\alpha = 1}^N {g_\alpha x_\alpha x}
912 \]
913 Completing the square,
914 \[
915 H_B + \Delta U = \sum\limits_{\alpha = 1}^N {\left\{
916 {\frac{{p_\alpha ^2 }}{{2m_\alpha }} + \frac{1}{2}m_\alpha
917 w_\alpha ^2 \left( {x_\alpha - \frac{{g_\alpha }}{{m_\alpha
918 w_\alpha ^2 }}x} \right)^2 } \right\}} - \sum\limits_{\alpha =
919 1}^N {\frac{{g_\alpha ^2 }}{{2m_\alpha w_\alpha ^2 }}} x^2
920 \]
921 and putting it back into Eq.~\ref{introEquation:bathGLE},
922 \[
923 H = \frac{{p^2 }}{{2m}} + W(x) + \sum\limits_{\alpha = 1}^N
924 {\left\{ {\frac{{p_\alpha ^2 }}{{2m_\alpha }} + \frac{1}{2}m_\alpha
925 w_\alpha ^2 \left( {x_\alpha - \frac{{g_\alpha }}{{m_\alpha
926 w_\alpha ^2 }}x} \right)^2 } \right\}}
927 \]
928 where
929 \[
930 W(x) = U(x) - \sum\limits_{\alpha = 1}^N {\frac{{g_\alpha ^2
931 }}{{2m_\alpha w_\alpha ^2 }}} x^2
932 \]
933 Since the first two terms of the new Hamiltonian depend only on the
934 system coordinates, we can get the equations of motion for
935 Generalized Langevin Dynamics by Hamilton's equations
936 \ref{introEquation:motionHamiltonianCoordinate,
937 introEquation:motionHamiltonianMomentum},
938 \begin{align}
939 \dot p &= - \frac{{\partial H}}{{\partial x}}
940 &= m\ddot x
941 &= - \frac{{\partial W(x)}}{{\partial x}} - \sum\limits_{\alpha = 1}^N {g_\alpha \left( {x_\alpha - \frac{{g_\alpha }}{{m_\alpha w_\alpha ^2 }}x} \right)}
942 \label{introEquation:Lp5}
943 \end{align}
944 , and
945 \begin{align}
946 \dot p_\alpha &= - \frac{{\partial H}}{{\partial x_\alpha }}
947 &= m\ddot x_\alpha
948 &= \- m_\alpha w_\alpha ^2 \left( {x_\alpha - \frac{{g_\alpha}}{{m_\alpha w_\alpha ^2 }}x} \right)
949 \end{align}
950
951 \subsection{\label{introSection:laplaceTransform}The Laplace Transform}
952
953 \[
954 L(x) = \int_0^\infty {x(t)e^{ - pt} dt}
955 \]
956
957 \[
958 L(x + y) = L(x) + L(y)
959 \]
960
961 \[
962 L(ax) = aL(x)
963 \]
964
965 \[
966 L(\dot x) = pL(x) - px(0)
967 \]
968
969 \[
970 L(\ddot x) = p^2 L(x) - px(0) - \dot x(0)
971 \]
972
973 \[
974 L\left( {\int_0^t {g(t - \tau )h(\tau )d\tau } } \right) = G(p)H(p)
975 \]
976
977 Some relatively important transformation,
978 \[
979 L(\cos at) = \frac{p}{{p^2 + a^2 }}
980 \]
981
982 \[
983 L(\sin at) = \frac{a}{{p^2 + a^2 }}
984 \]
985
986 \[
987 L(1) = \frac{1}{p}
988 \]
989
990 First, the bath coordinates,
991 \[
992 p^2 L(x_\alpha ) - px_\alpha (0) - \dot x_\alpha (0) = - \omega
993 _\alpha ^2 L(x_\alpha ) + \frac{{g_\alpha }}{{\omega _\alpha
994 }}L(x)
995 \]
996 \[
997 L(x_\alpha ) = \frac{{\frac{{g_\alpha }}{{\omega _\alpha }}L(x) +
998 px_\alpha (0) + \dot x_\alpha (0)}}{{p^2 + \omega _\alpha ^2 }}
999 \]
1000 Then, the system coordinates,
1001 \begin{align}
1002 mL(\ddot x) &= - \frac{1}{p}\frac{{\partial W(x)}}{{\partial x}} -
1003 \sum\limits_{\alpha = 1}^N {\left\{ {\frac{{\frac{{g_\alpha
1004 }}{{\omega _\alpha }}L(x) + px_\alpha (0) + \dot x_\alpha
1005 (0)}}{{p^2 + \omega _\alpha ^2 }} - \frac{{g_\alpha ^2 }}{{m_\alpha
1006 }}\omega _\alpha ^2 L(x)} \right\}}
1007 %
1008 &= - \frac{1}{p}\frac{{\partial W(x)}}{{\partial x}} -
1009 \sum\limits_{\alpha = 1}^N {\left\{ { - \frac{{g_\alpha ^2 }}{{m_\alpha \omega _\alpha ^2 }}\frac{p}{{p^2 + \omega _\alpha ^2 }}pL(x)
1010 - \frac{p}{{p^2 + \omega _\alpha ^2 }}g_\alpha x_\alpha (0)
1011 - \frac{1}{{p^2 + \omega _\alpha ^2 }}g_\alpha \dot x_\alpha (0)} \right\}}
1012 \end{align}
1013 Then, the inverse transform,
1014
1015 \begin{align}
1016 m\ddot x &= - \frac{{\partial W(x)}}{{\partial x}} -
1017 \sum\limits_{\alpha = 1}^N {\left\{ {\left( { - \frac{{g_\alpha ^2
1018 }}{{m_\alpha \omega _\alpha ^2 }}} \right)\int_0^t {\cos (\omega
1019 _\alpha t)\dot x(t - \tau )d\tau - \left[ {g_\alpha x_\alpha (0)
1020 - \frac{{g_\alpha }}{{m_\alpha \omega _\alpha }}} \right]\cos
1021 (\omega _\alpha t) - \frac{{g_\alpha \dot x_\alpha (0)}}{{\omega
1022 _\alpha }}\sin (\omega _\alpha t)} } \right\}}
1023 %
1024 &= - \frac{{\partial W(x)}}{{\partial x}} - \int_0^t
1025 {\sum\limits_{\alpha = 1}^N {\left( { - \frac{{g_\alpha ^2
1026 }}{{m_\alpha \omega _\alpha ^2 }}} \right)\cos (\omega _\alpha
1027 t)\dot x(t - \tau )d} \tau } + \sum\limits_{\alpha = 1}^N {\left\{
1028 {\left[ {g_\alpha x_\alpha (0) - \frac{{g_\alpha }}{{m_\alpha
1029 \omega _\alpha }}} \right]\cos (\omega _\alpha t) +
1030 \frac{{g_\alpha \dot x_\alpha (0)}}{{\omega _\alpha }}\sin
1031 (\omega _\alpha t)} \right\}}
1032 \end{align}
1033
1034 \begin{equation}
1035 m\ddot x = - \frac{{\partial W}}{{\partial x}} - \int_0^t {\xi
1036 (t)\dot x(t - \tau )d\tau } + R(t)
1037 \label{introEuqation:GeneralizedLangevinDynamics}
1038 \end{equation}
1039 %where $ {\xi (t)}$ is friction kernel, $R(t)$ is random force and
1040 %$W$ is the potential of mean force. $W(x) = - kT\ln p(x)$
1041 \[
1042 \xi (t) = \sum\limits_{\alpha = 1}^N {\left( { - \frac{{g_\alpha ^2
1043 }}{{m_\alpha \omega _\alpha ^2 }}} \right)\cos (\omega _\alpha t)}
1044 \]
1045 For an infinite harmonic bath, we can use the spectral density and
1046 an integral over frequencies.
1047
1048 \[
1049 R(t) = \sum\limits_{\alpha = 1}^N {\left( {g_\alpha x_\alpha (0)
1050 - \frac{{g_\alpha ^2 }}{{m_\alpha \omega _\alpha ^2 }}x(0)}
1051 \right)\cos (\omega _\alpha t)} + \frac{{\dot x_\alpha
1052 (0)}}{{\omega _\alpha }}\sin (\omega _\alpha t)
1053 \]
1054 The random forces depend only on initial conditions.
1055
1056 \subsubsection{\label{introSection:secondFluctuationDissipation}The Second Fluctuation Dissipation Theorem}
1057 So we can define a new set of coordinates,
1058 \[
1059 q_\alpha (t) = x_\alpha (t) - \frac{1}{{m_\alpha \omega _\alpha
1060 ^2 }}x(0)
1061 \]
1062 This makes
1063 \[
1064 R(t) = \sum\limits_{\alpha = 1}^N {g_\alpha q_\alpha (t)}
1065 \]
1066 And since the $q$ coordinates are harmonic oscillators,
1067 \[
1068 \begin{array}{l}
1069 \left\langle {q_\alpha (t)q_\alpha (0)} \right\rangle = \left\langle {q_\alpha ^2 (0)} \right\rangle \cos (\omega _\alpha t) \\
1070 \left\langle {q_\alpha (t)q_\beta (0)} \right\rangle = \delta _{\alpha \beta } \left\langle {q_\alpha (t)q_\alpha (0)} \right\rangle \\
1071 \end{array}
1072 \]
1073
1074 \begin{align}
1075 \left\langle {R(t)R(0)} \right\rangle &= \sum\limits_\alpha
1076 {\sum\limits_\beta {g_\alpha g_\beta \left\langle {q_\alpha
1077 (t)q_\beta (0)} \right\rangle } }
1078 %
1079 &= \sum\limits_\alpha {g_\alpha ^2 \left\langle {q_\alpha ^2 (0)}
1080 \right\rangle \cos (\omega _\alpha t)}
1081 %
1082 &= kT\xi (t)
1083 \end{align}
1084
1085 \begin{equation}
1086 \xi (t) = \left\langle {R(t)R(0)} \right\rangle
1087 \label{introEquation:secondFluctuationDissipation}
1088 \end{equation}
1089
1090 \section{\label{introSection:hydroynamics}Hydrodynamics}
1091
1092 \subsection{\label{introSection:frictionTensor} Friction Tensor}
1093 \subsection{\label{introSection:analyticalApproach}Analytical
1094 Approach}
1095
1096 \subsection{\label{introSection:approximationApproach}Approximation
1097 Approach}
1098
1099 \subsection{\label{introSection:centersRigidBody}Centers of Rigid
1100 Body}