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 2685 by tim, Mon Apr 3 18:07:54 2006 UTC vs.
Revision 2699 by tim, Mon Apr 10 05:35:55 2006 UTC

# Line 1 | Line 1
1   \chapter{\label{chapt:introduction}INTRODUCTION AND THEORETICAL BACKGROUND}
2  
3 < \section{\label{introSection:classicalMechanics}Classical Mechanics}
3 > \section{\label{introSection:classicalMechanics}Classical
4 > Mechanics}
5  
6 < \section{\label{introSection:rigidBody}Dynamics of Rigid Bodies}
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 < \section{\label{introSection:statisticalMechanics}Statistical 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 {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 presented in this dissertation.
225 +
226 + \subsection{\label{introSection:ensemble}Ensemble and Phase Space}
227 +
228 + \subsection{\label{introSection:ergodic}The Ergodic Hypothesis}
229 +
230 + Various thermodynamic properties can be calculated from Molecular
231 + Dynamics simulation. By comparing experimental values with the
232 + calculated properties, one can determine the accuracy of the
233 + simulation and the quality of the underlying model. However, both of
234 + experiment and computer simulation are usually performed during a
235 + certain time interval and the measurements are averaged over a
236 + period of them which is different from the average behavior of
237 + many-body system in Statistical Mechanics. Fortunately, Ergodic
238 + Hypothesis is proposed to make a connection between time average and
239 + ensemble average. It states that time average and average over the
240 + statistical ensemble are identical \cite{Frenkel1996, leach01:mm}.
241 + \begin{equation}
242 + \langle A \rangle_t = \mathop {\lim }\limits_{t \to \infty }
243 + \frac{1}{t}\int\limits_0^t {A(p(t),q(t))dt = \int\limits_\Gamma
244 + {A(p(t),q(t))} } \rho (p(t), q(t)) dpdq
245 + \end{equation}
246 + where $\langle A \rangle_t$ is an equilibrium value of a physical
247 + quantity and $\rho (p(t), q(t))$ is the equilibrium distribution
248 + function. If an observation is averaged over a sufficiently long
249 + time (longer than relaxation time), all accessible microstates in
250 + phase space are assumed to be equally probed, giving a properly
251 + weighted statistical average. This allows the researcher freedom of
252 + choice when deciding how best to measure a given observable. In case
253 + an ensemble averaged approach sounds most reasonable, the Monte
254 + Carlo techniques\cite{metropolis:1949} can be utilized. Or if the
255 + system lends itself to a time averaging approach, the Molecular
256 + Dynamics techniques in Sec.~\ref{introSection:molecularDynamics}
257 + will be the best choice\cite{Frenkel1996}.
258 +
259 + \section{\label{introSection:geometricIntegratos}Geometric Integrators}
260 + A variety of numerical integrators were proposed to simulate the
261 + motions. They usually begin with an initial conditionals and move
262 + the objects in the direction governed by the differential equations.
263 + However, most of them ignore the hidden physical law contained
264 + within the equations. Since 1990, geometric integrators, which
265 + preserve various phase-flow invariants such as symplectic structure,
266 + volume and time reversal symmetry, are developed to address this
267 + issue. The velocity verlet method, which happens to be a simple
268 + example of symplectic integrator, continues to gain its popularity
269 + in molecular dynamics community. This fact can be partly explained
270 + by its geometric nature.
271 +
272 + \subsection{\label{introSection:symplecticManifold}Symplectic Manifold}
273 + A \emph{manifold} is an abstract mathematical space. It locally
274 + looks like Euclidean space, but when viewed globally, it may have
275 + more complicate structure. A good example of manifold is the surface
276 + of Earth. It seems to be flat locally, but it is round if viewed as
277 + a whole. A \emph{differentiable manifold} (also known as
278 + \emph{smooth manifold}) is a manifold with an open cover in which
279 + the covering neighborhoods are all smoothly isomorphic to one
280 + another. In other words,it is possible to apply calculus on
281 + \emph{differentiable manifold}. A \emph{symplectic manifold} is
282 + defined as a pair $(M, \omega)$ which consisting of a
283 + \emph{differentiable manifold} $M$ and a close, non-degenerated,
284 + bilinear symplectic form, $\omega$. A symplectic form on a vector
285 + space $V$ is a function $\omega(x, y)$ which satisfies
286 + $\omega(\lambda_1x_1+\lambda_2x_2, y) = \lambda_1\omega(x_1, y)+
287 + \lambda_2\omega(x_2, y)$, $\omega(x, y) = - \omega(y, x)$ and
288 + $\omega(x, x) = 0$. Cross product operation in vector field is an
289 + example of symplectic form.
290 +
291 + One of the motivations to study \emph{symplectic manifold} in
292 + Hamiltonian Mechanics is that a symplectic manifold can represent
293 + all possible configurations of the system and the phase space of the
294 + system can be described by it's cotangent bundle. Every symplectic
295 + manifold is even dimensional. For instance, in Hamilton equations,
296 + coordinate and momentum always appear in pairs.
297 +
298 + Let  $(M,\omega)$ and $(N, \eta)$ be symplectic manifolds. A map
299 + \[
300 + f : M \rightarrow N
301 + \]
302 + is a \emph{symplectomorphism} if it is a \emph{diffeomorphims} and
303 + the \emph{pullback} of $\eta$ under f is equal to $\omega$.
304 + Canonical transformation is an example of symplectomorphism in
305 + classical mechanics.
306 +
307 + \subsection{\label{introSection:ODE}Ordinary Differential Equations}
308 +
309 + For a ordinary differential system defined as
310 + \begin{equation}
311 + \dot x = f(x)
312 + \end{equation}
313 + where $x = x(q,p)^T$, this system is canonical Hamiltonian, if
314 + \begin{equation}
315 + f(r) = J\nabla _x H(r).
316 + \end{equation}
317 + $H = H (q, p)$ is Hamiltonian function and $J$ is the skew-symmetric
318 + matrix
319 + \begin{equation}
320 + J = \left( {\begin{array}{*{20}c}
321 +   0 & I  \\
322 +   { - I} & 0  \\
323 + \end{array}} \right)
324 + \label{introEquation:canonicalMatrix}
325 + \end{equation}
326 + where $I$ is an identity matrix. Using this notation, Hamiltonian
327 + system can be rewritten as,
328 + \begin{equation}
329 + \frac{d}{{dt}}x = J\nabla _x H(x)
330 + \label{introEquation:compactHamiltonian}
331 + \end{equation}In this case, $f$ is
332 + called a \emph{Hamiltonian vector field}.
333 +
334 + Another generalization of Hamiltonian dynamics is Poisson Dynamics,
335 + \begin{equation}
336 + \dot x = J(x)\nabla _x H \label{introEquation:poissonHamiltonian}
337 + \end{equation}
338 + The most obvious change being that matrix $J$ now depends on $x$.
339 + The free rigid body is an example of Poisson system (actually a
340 + Lie-Poisson system) with Hamiltonian function of angular kinetic
341 + energy.
342 + \begin{equation}
343 + J(\pi ) = \left( {\begin{array}{*{20}c}
344 +   0 & {\pi _3 } & { - \pi _2 }  \\
345 +   { - \pi _3 } & 0 & {\pi _1 }  \\
346 +   {\pi _2 } & { - \pi _1 } & 0  \\
347 + \end{array}} \right)
348 + \end{equation}
349 +
350 + \begin{equation}
351 + H = \frac{1}{2}\left( {\frac{{\pi _1^2 }}{{I_1 }} + \frac{{\pi _2^2
352 + }}{{I_2 }} + \frac{{\pi _3^2 }}{{I_3 }}} \right)
353 + \end{equation}
354 +
355 + \subsection{\label{introSection:geometricProperties}Geometric Properties}
356 + Let $x(t)$ be the exact solution of the ODE system,
357 + \begin{equation}
358 + \frac{{dx}}{{dt}} = f(x) \label{introEquation:ODE}
359 + \end{equation}
360 + The exact flow(solution) $\varphi_\tau$ is defined by
361 + \[
362 + x(t+\tau) =\varphi_\tau(x(t))
363 + \]
364 + where $\tau$ is a fixed time step and $\varphi$ is a map from phase
365 + space to itself. In most cases, it is not easy to find the exact
366 + flow $\varphi_\tau$. Instead, we use a approximate map, $\psi_\tau$,
367 + which is usually called integrator. The order of an integrator
368 + $\psi_\tau$ is $p$, if the Taylor series of $\psi_\tau$ agree to
369 + order $p$,
370 + \begin{equation}
371 + \psi_tau(x) = x + \tau f(x) + O(\tau^{p+1})
372 + \end{equation}
373 +
374 + The hidden geometric properties of ODE and its flow play important
375 + roles in numerical studies. Let $\varphi$ be the flow of Hamiltonian
376 + vector field, $\varphi$ is a \emph{symplectic} flow if it satisfies,
377 + \begin{equation}
378 + '\varphi^T J '\varphi = J.
379 + \end{equation}
380 + According to Liouville's theorem, the symplectic volume is invariant
381 + under a Hamiltonian flow, which is the basis for classical
382 + statistical mechanics. Furthermore, the flow of a Hamiltonian vector
383 + field on a symplectic manifold can be shown to be a
384 + symplectomorphism. As to the Poisson system,
385 + \begin{equation}
386 + '\varphi ^T J '\varphi  = J \circ \varphi
387 + \end{equation}
388 + is the property must be preserved by the integrator. It is possible
389 + to construct a \emph{volume-preserving} flow for a source free($
390 + \nabla \cdot f = 0 $) ODE, if the flow satisfies $ \det d\varphi  =
391 + 1$. Changing the variables $y = h(x)$ in a
392 + ODE\ref{introEquation:ODE} will result in a new system,
393 + \[
394 + \dot y = \tilde f(y) = ((dh \cdot f)h^{ - 1} )(y).
395 + \]
396 + The vector filed $f$ has reversing symmetry $h$ if $f = - \tilde f$.
397 + In other words, the flow of this vector field is reversible if and
398 + only if $ h \circ \varphi ^{ - 1}  = \varphi  \circ h $. When
399 + designing any numerical methods, one should always try to preserve
400 + the structural properties of the original ODE and its flow.
401 +
402 + \subsection{\label{introSection:constructionSymplectic}Construction of Symplectic Methods}
403 + A lot of well established and very effective numerical methods have
404 + been successful precisely because of their symplecticities even
405 + though this fact was not recognized when they were first
406 + constructed. The most famous example is leapfrog methods in
407 + molecular dynamics. In general, symplectic integrators can be
408 + constructed using one of four different methods.
409 + \begin{enumerate}
410 + \item Generating functions
411 + \item Variational methods
412 + \item Runge-Kutta methods
413 + \item Splitting methods
414 + \end{enumerate}
415 +
416 + Generating function tends to lead to methods which are cumbersome
417 + and difficult to use\cite{}. In dissipative systems, variational
418 + methods can capture the decay of energy accurately\cite{}. Since
419 + their geometrically unstable nature against non-Hamiltonian
420 + perturbations, ordinary implicit Runge-Kutta methods are not
421 + suitable for Hamiltonian system. Recently, various high-order
422 + explicit Runge--Kutta methods have been developed to overcome this
423 + instability \cite{}. However, due to computational penalty involved
424 + in implementing the Runge-Kutta methods, they do not attract too
425 + much attention from Molecular Dynamics community. Instead, splitting
426 + have been widely accepted since they exploit natural decompositions
427 + of the system\cite{Tuckerman92}. The main idea behind splitting
428 + methods is to decompose the discrete $\varphi_h$ as a composition of
429 + simpler flows,
430 + \begin{equation}
431 + \varphi _h  = \varphi _{h_1 }  \circ \varphi _{h_2 }  \ldots  \circ
432 + \varphi _{h_n }
433 + \label{introEquation:FlowDecomposition}
434 + \end{equation}
435 + where each of the sub-flow is chosen such that each represent a
436 + simpler integration of the system. Let $\phi$ and $\psi$ both be
437 + symplectic maps, it is easy to show that any composition of
438 + symplectic flows yields a symplectic map,
439 + \begin{equation}
440 + (\varphi '\phi ')^T J\varphi '\phi ' = \phi '^T \varphi '^T J\varphi
441 + '\phi ' = \phi '^T J\phi ' = J.
442 + \label{introEquation:SymplecticFlowComposition}
443 + \end{equation}
444 + Suppose that a Hamiltonian system has a form with $H = T + V$
445 +
446 +
447 +
448   \section{\label{introSection:molecularDynamics}Molecular Dynamics}
449  
450 + As a special discipline of molecular modeling, Molecular dynamics
451 + has proven to be a powerful tool for studying the functions of
452 + biological systems, providing structural, thermodynamic and
453 + dynamical information.
454 +
455 + \subsection{\label{introSec:mdInit}Initialization}
456 +
457 + \subsection{\label{introSection:mdIntegration} Integration of the Equations of Motion}
458 +
459 + \section{\label{introSection:rigidBody}Dynamics of Rigid Bodies}
460 +
461 + A rigid body is a body in which the distance between any two given
462 + points of a rigid body remains constant regardless of external
463 + forces exerted on it. A rigid body therefore conserves its shape
464 + during its motion.
465 +
466 + Applications of dynamics of rigid bodies.
467 +
468 + \subsection{\label{introSection:lieAlgebra}Lie Algebra}
469 +
470 + \subsection{\label{introSection:DLMMotionEquation}The Euler Equations of Rigid Body Motion}
471 +
472 + \subsection{\label{introSection:otherRBMotionEquation}Other Formulations for Rigid Body Motion}
473 +
474 + %\subsection{\label{introSection:poissonBrackets}Poisson Brackets}
475 +
476 + \section{\label{introSection:correlationFunctions}Correlation Functions}
477 +
478   \section{\label{introSection:langevinDynamics}Langevin Dynamics}
479  
480 + \subsection{\label{introSection:LDIntroduction}Introduction and application of Langevin Dynamics}
481 +
482 + \subsection{\label{introSection:generalizedLangevinDynamics}Generalized Langevin Dynamics}
483 +
484 + \begin{equation}
485 + H = \frac{{p^2 }}{{2m}} + U(x) + H_B  + \Delta U(x,x_1 , \ldots x_N)
486 + \label{introEquation:bathGLE}
487 + \end{equation}
488 + where $H_B$ is harmonic bath Hamiltonian,
489 + \[
490 + H_B  =\sum\limits_{\alpha  = 1}^N {\left\{ {\frac{{p_\alpha ^2
491 + }}{{2m_\alpha  }} + \frac{1}{2}m_\alpha  w_\alpha ^2 } \right\}}
492 + \]
493 + and $\Delta U$ is bilinear system-bath coupling,
494 + \[
495 + \Delta U =  - \sum\limits_{\alpha  = 1}^N {g_\alpha  x_\alpha  x}
496 + \]
497 + Completing the square,
498 + \[
499 + H_B  + \Delta U = \sum\limits_{\alpha  = 1}^N {\left\{
500 + {\frac{{p_\alpha ^2 }}{{2m_\alpha  }} + \frac{1}{2}m_\alpha
501 + w_\alpha ^2 \left( {x_\alpha   - \frac{{g_\alpha  }}{{m_\alpha
502 + w_\alpha ^2 }}x} \right)^2 } \right\}}  - \sum\limits_{\alpha  =
503 + 1}^N {\frac{{g_\alpha ^2 }}{{2m_\alpha  w_\alpha ^2 }}} x^2
504 + \]
505 + and putting it back into Eq.~\ref{introEquation:bathGLE},
506 + \[
507 + H = \frac{{p^2 }}{{2m}} + W(x) + \sum\limits_{\alpha  = 1}^N
508 + {\left\{ {\frac{{p_\alpha ^2 }}{{2m_\alpha  }} + \frac{1}{2}m_\alpha
509 + w_\alpha ^2 \left( {x_\alpha   - \frac{{g_\alpha  }}{{m_\alpha
510 + w_\alpha ^2 }}x} \right)^2 } \right\}}
511 + \]
512 + where
513 + \[
514 + W(x) = U(x) - \sum\limits_{\alpha  = 1}^N {\frac{{g_\alpha ^2
515 + }}{{2m_\alpha  w_\alpha ^2 }}} x^2
516 + \]
517 + Since the first two terms of the new Hamiltonian depend only on the
518 + system coordinates, we can get the equations of motion for
519 + Generalized Langevin Dynamics by Hamilton's equations
520 + \ref{introEquation:motionHamiltonianCoordinate,
521 + introEquation:motionHamiltonianMomentum},
522 + \begin{align}
523 + \dot p &=  - \frac{{\partial H}}{{\partial x}}
524 +       &= m\ddot x
525 +       &= - \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)}
526 + \label{introEq:Lp5}
527 + \end{align}
528 + , and
529 + \begin{align}
530 + \dot p_\alpha   &=  - \frac{{\partial H}}{{\partial x_\alpha  }}
531 +                &= m\ddot x_\alpha
532 +                &= \- m_\alpha  w_\alpha ^2 \left( {x_\alpha   - \frac{{g_\alpha}}{{m_\alpha  w_\alpha ^2 }}x} \right)
533 + \end{align}
534 +
535 + \subsection{\label{introSection:laplaceTransform}The Laplace Transform}
536 +
537 + \[
538 + L(x) = \int_0^\infty  {x(t)e^{ - pt} dt}
539 + \]
540 +
541 + \[
542 + L(x + y) = L(x) + L(y)
543 + \]
544 +
545 + \[
546 + L(ax) = aL(x)
547 + \]
548 +
549 + \[
550 + L(\dot x) = pL(x) - px(0)
551 + \]
552 +
553 + \[
554 + L(\ddot x) = p^2 L(x) - px(0) - \dot x(0)
555 + \]
556 +
557 + \[
558 + L\left( {\int_0^t {g(t - \tau )h(\tau )d\tau } } \right) = G(p)H(p)
559 + \]
560 +
561 + Some relatively important transformation,
562 + \[
563 + L(\cos at) = \frac{p}{{p^2  + a^2 }}
564 + \]
565 +
566 + \[
567 + L(\sin at) = \frac{a}{{p^2  + a^2 }}
568 + \]
569 +
570 + \[
571 + L(1) = \frac{1}{p}
572 + \]
573 +
574 + First, the bath coordinates,
575 + \[
576 + p^2 L(x_\alpha  ) - px_\alpha  (0) - \dot x_\alpha  (0) =  - \omega
577 + _\alpha ^2 L(x_\alpha  ) + \frac{{g_\alpha  }}{{\omega _\alpha
578 + }}L(x)
579 + \]
580 + \[
581 + L(x_\alpha  ) = \frac{{\frac{{g_\alpha  }}{{\omega _\alpha  }}L(x) +
582 + px_\alpha  (0) + \dot x_\alpha  (0)}}{{p^2  + \omega _\alpha ^2 }}
583 + \]
584 + Then, the system coordinates,
585 + \begin{align}
586 + mL(\ddot x) &=  - \frac{1}{p}\frac{{\partial W(x)}}{{\partial x}} -
587 + \sum\limits_{\alpha  = 1}^N {\left\{ {\frac{{\frac{{g_\alpha
588 + }}{{\omega _\alpha  }}L(x) + px_\alpha  (0) + \dot x_\alpha
589 + (0)}}{{p^2  + \omega _\alpha ^2 }} - \frac{{g_\alpha ^2 }}{{m_\alpha
590 + }}\omega _\alpha ^2 L(x)} \right\}}
591 + %
592 + &= - \frac{1}{p}\frac{{\partial W(x)}}{{\partial x}} -
593 + \sum\limits_{\alpha  = 1}^N {\left\{ { - \frac{{g_\alpha ^2 }}{{m_\alpha  \omega _\alpha ^2 }}\frac{p}{{p^2  + \omega _\alpha ^2 }}pL(x)
594 + - \frac{p}{{p^2  + \omega _\alpha ^2 }}g_\alpha  x_\alpha  (0)
595 + - \frac{1}{{p^2  + \omega _\alpha ^2 }}g_\alpha  \dot x_\alpha  (0)} \right\}}
596 + \end{align}
597 + Then, the inverse transform,
598 +
599 + \begin{align}
600 + m\ddot x &=  - \frac{{\partial W(x)}}{{\partial x}} -
601 + \sum\limits_{\alpha  = 1}^N {\left\{ {\left( { - \frac{{g_\alpha ^2
602 + }}{{m_\alpha  \omega _\alpha ^2 }}} \right)\int_0^t {\cos (\omega
603 + _\alpha  t)\dot x(t - \tau )d\tau  - \left[ {g_\alpha  x_\alpha  (0)
604 + - \frac{{g_\alpha  }}{{m_\alpha  \omega _\alpha  }}} \right]\cos
605 + (\omega _\alpha  t) - \frac{{g_\alpha  \dot x_\alpha  (0)}}{{\omega
606 + _\alpha  }}\sin (\omega _\alpha  t)} } \right\}}
607 + %
608 + &= - \frac{{\partial W(x)}}{{\partial x}} - \int_0^t
609 + {\sum\limits_{\alpha  = 1}^N {\left( { - \frac{{g_\alpha ^2
610 + }}{{m_\alpha  \omega _\alpha ^2 }}} \right)\cos (\omega _\alpha
611 + t)\dot x(t - \tau )d} \tau }  + \sum\limits_{\alpha  = 1}^N {\left\{
612 + {\left[ {g_\alpha  x_\alpha  (0) - \frac{{g_\alpha  }}{{m_\alpha
613 + \omega _\alpha  }}} \right]\cos (\omega _\alpha  t) +
614 + \frac{{g_\alpha  \dot x_\alpha  (0)}}{{\omega _\alpha  }}\sin
615 + (\omega _\alpha  t)} \right\}}
616 + \end{align}
617 +
618 + \begin{equation}
619 + m\ddot x =  - \frac{{\partial W}}{{\partial x}} - \int_0^t {\xi
620 + (t)\dot x(t - \tau )d\tau }  + R(t)
621 + \label{introEuqation:GeneralizedLangevinDynamics}
622 + \end{equation}
623 + %where $ {\xi (t)}$ is friction kernel, $R(t)$ is random force and
624 + %$W$ is the potential of mean force. $W(x) =  - kT\ln p(x)$
625 + \[
626 + \xi (t) = \sum\limits_{\alpha  = 1}^N {\left( { - \frac{{g_\alpha ^2
627 + }}{{m_\alpha  \omega _\alpha ^2 }}} \right)\cos (\omega _\alpha  t)}
628 + \]
629 + For an infinite harmonic bath, we can use the spectral density and
630 + an integral over frequencies.
631 +
632 + \[
633 + R(t) = \sum\limits_{\alpha  = 1}^N {\left( {g_\alpha  x_\alpha  (0)
634 + - \frac{{g_\alpha ^2 }}{{m_\alpha  \omega _\alpha ^2 }}x(0)}
635 + \right)\cos (\omega _\alpha  t)}  + \frac{{\dot x_\alpha
636 + (0)}}{{\omega _\alpha  }}\sin (\omega _\alpha  t)
637 + \]
638 + The random forces depend only on initial conditions.
639 +
640 + \subsubsection{\label{introSection:secondFluctuationDissipation}The Second Fluctuation Dissipation Theorem}
641 + So we can define a new set of coordinates,
642 + \[
643 + q_\alpha  (t) = x_\alpha  (t) - \frac{1}{{m_\alpha  \omega _\alpha
644 + ^2 }}x(0)
645 + \]
646 + This makes
647 + \[
648 + R(t) = \sum\limits_{\alpha  = 1}^N {g_\alpha  q_\alpha  (t)}
649 + \]
650 + And since the $q$ coordinates are harmonic oscillators,
651 + \[
652 + \begin{array}{l}
653 + \left\langle {q_\alpha  (t)q_\alpha  (0)} \right\rangle  = \left\langle {q_\alpha ^2 (0)} \right\rangle \cos (\omega _\alpha  t) \\
654 + \left\langle {q_\alpha  (t)q_\beta  (0)} \right\rangle  = \delta _{\alpha \beta } \left\langle {q_\alpha  (t)q_\alpha  (0)} \right\rangle  \\
655 + \end{array}
656 + \]
657 +
658 + \begin{align}
659 + \left\langle {R(t)R(0)} \right\rangle  &= \sum\limits_\alpha
660 + {\sum\limits_\beta  {g_\alpha  g_\beta  \left\langle {q_\alpha
661 + (t)q_\beta  (0)} \right\rangle } }
662 + %
663 + &= \sum\limits_\alpha  {g_\alpha ^2 \left\langle {q_\alpha ^2 (0)}
664 + \right\rangle \cos (\omega _\alpha  t)}
665 + %
666 + &= kT\xi (t)
667 + \end{align}
668 +
669 + \begin{equation}
670 + \xi (t) = \left\langle {R(t)R(0)} \right\rangle
671 + \label{introEquation:secondFluctuationDissipation}
672 + \end{equation}
673 +
674   \section{\label{introSection:hydroynamics}Hydrodynamics}
675  
676 < \section{\label{introSection:correlationFunctions}Correlation Functions}
676 > \subsection{\label{introSection:frictionTensor} Friction Tensor}
677 > \subsection{\label{introSection:analyticalApproach}Analytical
678 > Approach}
679 >
680 > \subsection{\label{introSection:approximationApproach}Approximation
681 > Approach}
682 >
683 > \subsection{\label{introSection:centersRigidBody}Centers of Rigid
684 > Body}

Diff Legend

Removed lines
+ Added lines
< Changed lines
> Changed lines