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 2697 by tim, Fri Apr 7 05:03:54 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}
216 + \label{introEquation:conserveHalmitonian}
217 + \end{equation}
218 +
219 + When studying Hamiltonian system, it is more convenient to use
220 + notation
221 + \begin{equation}
222 + r = r(q,p)^T
223 + \end{equation}
224 + and to introduce a $2n \times 2n$ canonical structure matrix $J$,
225 + \begin{equation}
226 + J = \left( {\begin{array}{*{20}c}
227 +   0 & I  \\
228 +   { - I} & 0  \\
229 + \end{array}} \right)
230 + \label{introEquation:canonicalMatrix}
231 + \end{equation}
232 + where $I$ is a $n \times n$ identity matrix and $J$ is a
233 + skew-symmetric matrix ($ J^T  =  - J $). Thus, Hamiltonian system
234 + can be rewritten as,
235 + \begin{equation}
236 + \frac{d}{{dt}}r = J\nabla _r H(r)
237 + \label{introEquation:compactHamiltonian}
238 + \end{equation}
239 +
240 + \section{\label{introSection:statisticalMechanics}Statistical
241 + Mechanics}
242 +
243 + The thermodynamic behaviors and properties of Molecular Dynamics
244 + simulation are governed by the principle of Statistical Mechanics.
245 + The following section will give a brief introduction to some of the
246 + Statistical Mechanics concepts presented in this dissertation.
247 +
248 + \subsection{\label{introSection:ensemble}Ensemble and Phase Space}
249 +
250 + \subsection{\label{introSection:ergodic}The Ergodic Hypothesis}
251 +
252 + Various thermodynamic properties can be calculated from Molecular
253 + Dynamics simulation. By comparing experimental values with the
254 + calculated properties, one can determine the accuracy of the
255 + simulation and the quality of the underlying model. However, both of
256 + experiment and computer simulation are usually performed during a
257 + certain time interval and the measurements are averaged over a
258 + period of them which is different from the average behavior of
259 + many-body system in Statistical Mechanics. Fortunately, Ergodic
260 + Hypothesis is proposed to make a connection between time average and
261 + ensemble average. It states that time average and average over the
262 + statistical ensemble are identical \cite{Frenkel1996, leach01:mm}.
263 + \begin{equation}
264 + \langle A \rangle_t = \mathop {\lim }\limits_{t \to \infty }
265 + \frac{1}{t}\int\limits_0^t {A(p(t),q(t))dt = \int\limits_\Gamma
266 + {A(p(t),q(t))} } \rho (p(t), q(t)) dpdq
267 + \end{equation}
268 + where $\langle A \rangle_t$ is an equilibrium value of a physical
269 + quantity and $\rho (p(t), q(t))$ is the equilibrium distribution
270 + function. If an observation is averaged over a sufficiently long
271 + time (longer than relaxation time), all accessible microstates in
272 + phase space are assumed to be equally probed, giving a properly
273 + weighted statistical average. This allows the researcher freedom of
274 + choice when deciding how best to measure a given observable. In case
275 + an ensemble averaged approach sounds most reasonable, the Monte
276 + Carlo techniques\cite{metropolis:1949} can be utilized. Or if the
277 + system lends itself to a time averaging approach, the Molecular
278 + Dynamics techniques in Sec.~\ref{introSection:molecularDynamics}
279 + will be the best choice\cite{Frenkel1996}.
280 +
281 + \section{\label{introSection:geometricIntegratos}Geometric Integrators}
282 + A variety of numerical integrators were proposed to simulate the
283 + motions. They usually begin with an initial conditionals and move
284 + the objects in the direction governed by the differential equations.
285 + However, most of them ignore the hidden physical law contained
286 + within the equations. Since 1990, geometric integrators, which
287 + preserve various phase-flow invariants such as symplectic structure,
288 + volume and time reversal symmetry, are developed to address this
289 + issue. The velocity verlet method, which happens to be a simple
290 + example of symplectic integrator, continues to gain its popularity
291 + in molecular dynamics community. This fact can be partly explained
292 + by its geometric nature.
293 +
294 + \subsection{\label{introSection:symplecticManifold}Symplectic Manifold}
295 + A \emph{manifold} is an abstract mathematical space. It locally
296 + looks like Euclidean space, but when viewed globally, it may have
297 + more complicate structure. A good example of manifold is the surface
298 + of Earth. It seems to be flat locally, but it is round if viewed as
299 + a whole. A \emph{differentiable manifold} (also known as
300 + \emph{smooth manifold}) is a manifold with an open cover in which
301 + the covering neighborhoods are all smoothly isomorphic to one
302 + another. In other words,it is possible to apply calculus on
303 + \emph{differentiable manifold}. A \emph{symplectic manifold} is
304 + defined as a pair $(M, \omega)$ which consisting of a
305 + \emph{differentiable manifold} $M$ and a close, non-degenerated,
306 + bilinear symplectic form, $\omega$. A symplectic form on a vector
307 + space $V$ is a function $\omega(x, y)$ which satisfies
308 + $\omega(\lambda_1x_1+\lambda_2x_2, y) = \lambda_1\omega(x_1, y)+
309 + \lambda_2\omega(x_2, y)$, $\omega(x, y) = - \omega(y, x)$ and
310 + $\omega(x, x) = 0$. Cross product operation in vector field is an
311 + example of symplectic form.
312 +
313 + One of the motivations to study \emph{symplectic manifold} in
314 + Hamiltonian Mechanics is that a symplectic manifold can represent
315 + all possible configurations of the system and the phase space of the
316 + system can be described by it's cotangent bundle. Every symplectic
317 + manifold is even dimensional. For instance, in Hamilton equations,
318 + coordinate and momentum always appear in pairs.
319 +
320 + Let  $(M,\omega)$ and $(N, \eta)$ be symplectic manifolds. A map
321 + \[
322 + f : M \rightarrow N
323 + \]
324 + is a \emph{symplectomorphism} if it is a \emph{diffeomorphims} and
325 + the \emph{pullback} of $\eta$ under f is equal to $\omega$.
326 + Canonical transformation is an example of symplectomorphism in
327 + classical mechanics. According to Liouville's theorem, the
328 + Hamiltonian \emph{flow} or \emph{symplectomorphism} generated by the
329 + Hamiltonian vector filed preserves the volume form on the phase
330 + space, which is the basis of classical statistical mechanics.
331 +
332 + \subsection{\label{introSection:exactFlow}The Exact Flow of ODE}
333 +
334 + \subsection{\label{introSection:hamiltonianSplitting}Hamiltonian Splitting}
335 +
336   \section{\label{introSection:molecularDynamics}Molecular Dynamics}
337  
338 + As a special discipline of molecular modeling, Molecular dynamics
339 + has proven to be a powerful tool for studying the functions of
340 + biological systems, providing structural, thermodynamic and
341 + dynamical information.
342 +
343 + \subsection{\label{introSec:mdInit}Initialization}
344 +
345 + \subsection{\label{introSection:mdIntegration} Integration of the Equations of Motion}
346 +
347 + \section{\label{introSection:rigidBody}Dynamics of Rigid Bodies}
348 +
349 + A rigid body is a body in which the distance between any two given
350 + points of a rigid body remains constant regardless of external
351 + forces exerted on it. A rigid body therefore conserves its shape
352 + during its motion.
353 +
354 + Applications of dynamics of rigid bodies.
355 +
356 + \subsection{\label{introSection:lieAlgebra}Lie Algebra}
357 +
358 + \subsection{\label{introSection:DLMMotionEquation}The Euler Equations of Rigid Body Motion}
359 +
360 + \subsection{\label{introSection:otherRBMotionEquation}Other Formulations for Rigid Body Motion}
361 +
362 + %\subsection{\label{introSection:poissonBrackets}Poisson Brackets}
363 +
364 + \section{\label{introSection:correlationFunctions}Correlation Functions}
365 +
366   \section{\label{introSection:langevinDynamics}Langevin Dynamics}
367  
368 + \subsection{\label{introSection:LDIntroduction}Introduction and application of Langevin Dynamics}
369 +
370 + \subsection{\label{introSection:generalizedLangevinDynamics}Generalized Langevin Dynamics}
371 +
372 + \begin{equation}
373 + H = \frac{{p^2 }}{{2m}} + U(x) + H_B  + \Delta U(x,x_1 , \ldots x_N)
374 + \label{introEquation:bathGLE}
375 + \end{equation}
376 + where $H_B$ is harmonic bath Hamiltonian,
377 + \[
378 + H_B  =\sum\limits_{\alpha  = 1}^N {\left\{ {\frac{{p_\alpha ^2
379 + }}{{2m_\alpha  }} + \frac{1}{2}m_\alpha  w_\alpha ^2 } \right\}}
380 + \]
381 + and $\Delta U$ is bilinear system-bath coupling,
382 + \[
383 + \Delta U =  - \sum\limits_{\alpha  = 1}^N {g_\alpha  x_\alpha  x}
384 + \]
385 + Completing the square,
386 + \[
387 + H_B  + \Delta U = \sum\limits_{\alpha  = 1}^N {\left\{
388 + {\frac{{p_\alpha ^2 }}{{2m_\alpha  }} + \frac{1}{2}m_\alpha
389 + w_\alpha ^2 \left( {x_\alpha   - \frac{{g_\alpha  }}{{m_\alpha
390 + w_\alpha ^2 }}x} \right)^2 } \right\}}  - \sum\limits_{\alpha  =
391 + 1}^N {\frac{{g_\alpha ^2 }}{{2m_\alpha  w_\alpha ^2 }}} x^2
392 + \]
393 + and putting it back into Eq.~\ref{introEquation:bathGLE},
394 + \[
395 + H = \frac{{p^2 }}{{2m}} + W(x) + \sum\limits_{\alpha  = 1}^N
396 + {\left\{ {\frac{{p_\alpha ^2 }}{{2m_\alpha  }} + \frac{1}{2}m_\alpha
397 + w_\alpha ^2 \left( {x_\alpha   - \frac{{g_\alpha  }}{{m_\alpha
398 + w_\alpha ^2 }}x} \right)^2 } \right\}}
399 + \]
400 + where
401 + \[
402 + W(x) = U(x) - \sum\limits_{\alpha  = 1}^N {\frac{{g_\alpha ^2
403 + }}{{2m_\alpha  w_\alpha ^2 }}} x^2
404 + \]
405 + Since the first two terms of the new Hamiltonian depend only on the
406 + system coordinates, we can get the equations of motion for
407 + Generalized Langevin Dynamics by Hamilton's equations
408 + \ref{introEquation:motionHamiltonianCoordinate,
409 + introEquation:motionHamiltonianMomentum},
410 + \begin{align}
411 + \dot p &=  - \frac{{\partial H}}{{\partial x}}
412 +       &= m\ddot x
413 +       &= - \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)}
414 + \label{introEq:Lp5}
415 + \end{align}
416 + , and
417 + \begin{align}
418 + \dot p_\alpha   &=  - \frac{{\partial H}}{{\partial x_\alpha  }}
419 +                &= m\ddot x_\alpha
420 +                &= \- m_\alpha  w_\alpha ^2 \left( {x_\alpha   - \frac{{g_\alpha}}{{m_\alpha  w_\alpha ^2 }}x} \right)
421 + \end{align}
422 +
423 + \subsection{\label{introSection:laplaceTransform}The Laplace Transform}
424 +
425 + \[
426 + L(x) = \int_0^\infty  {x(t)e^{ - pt} dt}
427 + \]
428 +
429 + \[
430 + L(x + y) = L(x) + L(y)
431 + \]
432 +
433 + \[
434 + L(ax) = aL(x)
435 + \]
436 +
437 + \[
438 + L(\dot x) = pL(x) - px(0)
439 + \]
440 +
441 + \[
442 + L(\ddot x) = p^2 L(x) - px(0) - \dot x(0)
443 + \]
444 +
445 + \[
446 + L\left( {\int_0^t {g(t - \tau )h(\tau )d\tau } } \right) = G(p)H(p)
447 + \]
448 +
449 + Some relatively important transformation,
450 + \[
451 + L(\cos at) = \frac{p}{{p^2  + a^2 }}
452 + \]
453 +
454 + \[
455 + L(\sin at) = \frac{a}{{p^2  + a^2 }}
456 + \]
457 +
458 + \[
459 + L(1) = \frac{1}{p}
460 + \]
461 +
462 + First, the bath coordinates,
463 + \[
464 + p^2 L(x_\alpha  ) - px_\alpha  (0) - \dot x_\alpha  (0) =  - \omega
465 + _\alpha ^2 L(x_\alpha  ) + \frac{{g_\alpha  }}{{\omega _\alpha
466 + }}L(x)
467 + \]
468 + \[
469 + L(x_\alpha  ) = \frac{{\frac{{g_\alpha  }}{{\omega _\alpha  }}L(x) +
470 + px_\alpha  (0) + \dot x_\alpha  (0)}}{{p^2  + \omega _\alpha ^2 }}
471 + \]
472 + Then, the system coordinates,
473 + \begin{align}
474 + mL(\ddot x) &=  - \frac{1}{p}\frac{{\partial W(x)}}{{\partial x}} -
475 + \sum\limits_{\alpha  = 1}^N {\left\{ {\frac{{\frac{{g_\alpha
476 + }}{{\omega _\alpha  }}L(x) + px_\alpha  (0) + \dot x_\alpha
477 + (0)}}{{p^2  + \omega _\alpha ^2 }} - \frac{{g_\alpha ^2 }}{{m_\alpha
478 + }}\omega _\alpha ^2 L(x)} \right\}}
479 + %
480 + &= - \frac{1}{p}\frac{{\partial W(x)}}{{\partial x}} -
481 + \sum\limits_{\alpha  = 1}^N {\left\{ { - \frac{{g_\alpha ^2 }}{{m_\alpha  \omega _\alpha ^2 }}\frac{p}{{p^2  + \omega _\alpha ^2 }}pL(x)
482 + - \frac{p}{{p^2  + \omega _\alpha ^2 }}g_\alpha  x_\alpha  (0)
483 + - \frac{1}{{p^2  + \omega _\alpha ^2 }}g_\alpha  \dot x_\alpha  (0)} \right\}}
484 + \end{align}
485 + Then, the inverse transform,
486 +
487 + \begin{align}
488 + m\ddot x &=  - \frac{{\partial W(x)}}{{\partial x}} -
489 + \sum\limits_{\alpha  = 1}^N {\left\{ {\left( { - \frac{{g_\alpha ^2
490 + }}{{m_\alpha  \omega _\alpha ^2 }}} \right)\int_0^t {\cos (\omega
491 + _\alpha  t)\dot x(t - \tau )d\tau  - \left[ {g_\alpha  x_\alpha  (0)
492 + - \frac{{g_\alpha  }}{{m_\alpha  \omega _\alpha  }}} \right]\cos
493 + (\omega _\alpha  t) - \frac{{g_\alpha  \dot x_\alpha  (0)}}{{\omega
494 + _\alpha  }}\sin (\omega _\alpha  t)} } \right\}}
495 + %
496 + &= - \frac{{\partial W(x)}}{{\partial x}} - \int_0^t
497 + {\sum\limits_{\alpha  = 1}^N {\left( { - \frac{{g_\alpha ^2
498 + }}{{m_\alpha  \omega _\alpha ^2 }}} \right)\cos (\omega _\alpha
499 + t)\dot x(t - \tau )d} \tau }  + \sum\limits_{\alpha  = 1}^N {\left\{
500 + {\left[ {g_\alpha  x_\alpha  (0) - \frac{{g_\alpha  }}{{m_\alpha
501 + \omega _\alpha  }}} \right]\cos (\omega _\alpha  t) +
502 + \frac{{g_\alpha  \dot x_\alpha  (0)}}{{\omega _\alpha  }}\sin
503 + (\omega _\alpha  t)} \right\}}
504 + \end{align}
505 +
506 + \begin{equation}
507 + m\ddot x =  - \frac{{\partial W}}{{\partial x}} - \int_0^t {\xi
508 + (t)\dot x(t - \tau )d\tau }  + R(t)
509 + \label{introEuqation:GeneralizedLangevinDynamics}
510 + \end{equation}
511 + %where $ {\xi (t)}$ is friction kernel, $R(t)$ is random force and
512 + %$W$ is the potential of mean force. $W(x) =  - kT\ln p(x)$
513 + \[
514 + \xi (t) = \sum\limits_{\alpha  = 1}^N {\left( { - \frac{{g_\alpha ^2
515 + }}{{m_\alpha  \omega _\alpha ^2 }}} \right)\cos (\omega _\alpha  t)}
516 + \]
517 + For an infinite harmonic bath, we can use the spectral density and
518 + an integral over frequencies.
519 +
520 + \[
521 + R(t) = \sum\limits_{\alpha  = 1}^N {\left( {g_\alpha  x_\alpha  (0)
522 + - \frac{{g_\alpha ^2 }}{{m_\alpha  \omega _\alpha ^2 }}x(0)}
523 + \right)\cos (\omega _\alpha  t)}  + \frac{{\dot x_\alpha
524 + (0)}}{{\omega _\alpha  }}\sin (\omega _\alpha  t)
525 + \]
526 + The random forces depend only on initial conditions.
527 +
528 + \subsubsection{\label{introSection:secondFluctuationDissipation}The Second Fluctuation Dissipation Theorem}
529 + So we can define a new set of coordinates,
530 + \[
531 + q_\alpha  (t) = x_\alpha  (t) - \frac{1}{{m_\alpha  \omega _\alpha
532 + ^2 }}x(0)
533 + \]
534 + This makes
535 + \[
536 + R(t) = \sum\limits_{\alpha  = 1}^N {g_\alpha  q_\alpha  (t)}
537 + \]
538 + And since the $q$ coordinates are harmonic oscillators,
539 + \[
540 + \begin{array}{l}
541 + \left\langle {q_\alpha  (t)q_\alpha  (0)} \right\rangle  = \left\langle {q_\alpha ^2 (0)} \right\rangle \cos (\omega _\alpha  t) \\
542 + \left\langle {q_\alpha  (t)q_\beta  (0)} \right\rangle  = \delta _{\alpha \beta } \left\langle {q_\alpha  (t)q_\alpha  (0)} \right\rangle  \\
543 + \end{array}
544 + \]
545 +
546 + \begin{align}
547 + \left\langle {R(t)R(0)} \right\rangle  &= \sum\limits_\alpha
548 + {\sum\limits_\beta  {g_\alpha  g_\beta  \left\langle {q_\alpha
549 + (t)q_\beta  (0)} \right\rangle } }
550 + %
551 + &= \sum\limits_\alpha  {g_\alpha ^2 \left\langle {q_\alpha ^2 (0)}
552 + \right\rangle \cos (\omega _\alpha  t)}
553 + %
554 + &= kT\xi (t)
555 + \end{align}
556 +
557 + \begin{equation}
558 + \xi (t) = \left\langle {R(t)R(0)} \right\rangle
559 + \label{introEquation:secondFluctuationDissipation}
560 + \end{equation}
561 +
562   \section{\label{introSection:hydroynamics}Hydrodynamics}
563  
564 < \section{\label{introSection:correlationFunctions}Correlation Functions}
564 > \subsection{\label{introSection:frictionTensor} Friction Tensor}
565 > \subsection{\label{introSection:analyticalApproach}Analytical
566 > Approach}
567 >
568 > \subsection{\label{introSection:approximationApproach}Approximation
569 > Approach}
570 >
571 > \subsection{\label{introSection:centersRigidBody}Centers of Rigid
572 > Body}

Diff Legend

Removed lines
+ Added lines
< Changed lines
> Changed lines