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

Comparing trunk/tengDissertation/Introduction.tex (file contents):
Revision 2692 by tim, Tue Apr 4 21:32:58 2006 UTC vs.
Revision 2696 by tim, Thu Apr 6 22:06:50 2006 UTC

# Line 1 | Line 1
1   \chapter{\label{chapt:introduction}INTRODUCTION AND THEORETICAL BACKGROUND}
2  
3 < \section{\label{introSection:molecularDynamics}Molecular Dynamics}
3 > \section{\label{introSection:classicalMechanics}Classical
4 > Mechanics}
5  
5 As a special discipline of molecular modeling, Molecular dynamics
6 has proven to be a powerful tool for studying the functions of
7 biological systems, providing structural, thermodynamic and
8 dynamical information.
9
10 \subsection{\label{introSection:classicalMechanics}Classical Mechanics}
11
6   Closely related to Classical Mechanics, Molecular Dynamics
7   simulations are carried out by integrating the equations of motion
8   for a given system of particles. There are three fundamental ideas
# Line 20 | Line 14 | sufficient to predict the future behavior of the syste
14   when further combine with the laws of mechanics will also be
15   sufficient to predict the future behavior of the system.
16  
17 < \subsubsection{\label{introSection:newtonian}Newtonian Mechanics}
17 > \subsection{\label{introSection:newtonian}Newtonian Mechanics}
18 > The discovery of Newton's three laws of mechanics which govern the
19 > motion of particles is the foundation of the classical mechanics.
20 > Newton¡¯s first law defines a class of inertial frames. Inertial
21 > frames are reference frames where a particle not interacting with
22 > other bodies will move with constant speed in the same direction.
23 > With respect to inertial frames Newton¡¯s second law has the form
24 > \begin{equation}
25 > F = \frac {dp}{dt} = \frac {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 < \subsubsection{\label{introSection:lagrangian}Lagrangian Mechanics}
38 > Conservation laws of Newtonian Mechanics play very important roles
39 > in solving mechanics problems. The linear momentum of a particle is
40 > conserved if it is free or it experiences no force. The second
41 > conservation theorem concerns the angular momentum of a particle.
42 > The angular momentum $L$ of a particle with respect to an origin
43 > from which $r$ is measured is defined to be
44 > \begin{equation}
45 > L \equiv r \times p \label{introEquation:angularMomentumDefinition}
46 > \end{equation}
47 > The torque $\tau$ with respect to the same origin is defined to be
48 > \begin{equation}
49 > 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
# Line 35 | Line 83 | system, alternative procedures may be developed.
83   which arise in attempts to apply Newton's equation to complex
84   system, alternative procedures may be developed.
85  
86 < \subsubsubsection{\label{introSection:halmiltonPrinciple}Hamilton's
86 > \subsubsection{\label{introSection:halmiltonPrinciple}Hamilton's
87   Principle}
88  
89   Hamilton introduced the dynamical principle upon which it is
# Line 45 | Line 93 | the kinetic, $K$, and potential energies, $U$.
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, $U$.
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 < \lable{introEquation:halmitonianPrinciple1}
99 > \label{introEquation:halmitonianPrinciple1}
100   \end{equation}
101  
102   For simple mechanical systems, where the forces acting on the
# Line 62 | Line 110 | then Eq.~\ref{introEquation:halmitonianPrinciple1} bec
110   \end{equation}
111   then Eq.~\ref{introEquation:halmitonianPrinciple1} becomes
112   \begin{equation}
113 < \delta \int_{t_1 }^{t_2 } {K dt = 0} ,
114 < \lable{introEquation:halmitonianPrinciple2}
113 > \delta \int_{t_1 }^{t_2 } {L dt = 0} ,
114 > \label{introEquation:halmitonianPrinciple2}
115   \end{equation}
116  
117 < \subsubsubsection{\label{introSection:equationOfMotionLagrangian}The
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
# Line 74 | Line 122 | 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 < \lable{introEquation:eqMotionLagrangian}
125 > \label{introEquation:eqMotionLagrangian}
126   \end{equation}
127   where $q_{i}$ is generalized coordinate and $\dot{q_{i}}$ is
128   generalized velocity.
129  
130 < \subsubsection{\label{introSection:hamiltonian}Hamiltonian Mechanics}
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
# Line 90 | Line 138 | With the help of these momenta, we may now define a ne
138   p_i = \frac{\partial L}{\partial \dot q_i}
139   \label{introEquation:generalizedMomenta}
140   \end{equation}
141 < With the help of these momenta, we may now define a new quantity $H$
94 < by the equation
141 > The Lagrange equations of motion are then expressed by
142   \begin{equation}
143 < H = p_1 \dot q_1  +  \ldots  + p_f \dot q_f  - L,
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
# Line 108 | Line 200 | equations.
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.
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 < \subsubsection{\label{introSection:canonicalTransformation}Canonical Transformation}
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 < \subsection{\label{introSection:statisticalMechanics}Statistical Mechanics}
240 > \section{\label{introSection:geometricIntegratos}Geometric Integrators}
241  
242 < The thermodynamic behaviors and properties  of Molecular Dynamics
242 > \subsection{\label{introSection:symplecticManifold}Symplectic Manifold}
243 > A \emph{manifold} is an abstract mathematical space. It locally
244 > looks like Euclidean space, but when viewed globally, it may have
245 > more complicate structure. A good example of manifold is the surface
246 > of Earth. It seems to be flat locally, but it is round if viewed as
247 > a whole. A \emph{differentiable manifold} (also known as
248 > \emph{smooth manifold}) is a manifold with an open cover in which
249 > the covering neighborhoods are all smoothly isomorphic to one
250 > another. In other words,it is possible to apply calculus on
251 > \emph{differentiable manifold}. A \emph{symplectic manifold} is
252 > defined as a pair $(M, \omega)$ consisting of a \emph{differentiable
253 > manifold} $M$ and a close, non-degenerated, bilinear symplectic
254 > form, $\omega$. One of the motivation to study \emph{symplectic
255 > manifold} in Hamiltonian Mechanics is that a symplectic manifold can
256 > represent all possible configurations of the system and the phase
257 > space of the system can be described by it's cotangent bundle. Every
258 > symplectic manifold is even dimensional. For instance, in Hamilton
259 > equations, coordinate and momentum always appear in pairs.
260 >
261 > A \emph{symplectomorphism} is also known as a \emph{canonical
262 > transformation}.
263 >
264 > Any real-valued differentiable function H on a symplectic manifold
265 > can serve as an energy function or Hamiltonian. Associated to any
266 > Hamiltonian is a Hamiltonian vector field; the integral curves of
267 > the Hamiltonian vector field are solutions to the Hamilton-Jacobi
268 > equations. The Hamiltonian vector field defines a flow on the
269 > symplectic manifold, called a Hamiltonian flow or symplectomorphism.
270 > By Liouville's theorem, Hamiltonian flows preserve the volume form
271 > on the phase space.
272 >
273 > \subsection{\label{Construction of Symplectic Methods}}
274 >
275 > \section{\label{introSection:statisticalMechanics}Statistical
276 > Mechanics}
277 >
278 > The thermodynamic behaviors and properties of Molecular Dynamics
279   simulation are governed by the principle of Statistical Mechanics.
280   The following section will give a brief introduction to some of the
281   Statistical Mechanics concepts presented in this dissertation.
282  
283 < \subsubsection{\label{introSection::ensemble}Ensemble}
283 > \subsection{\label{introSection:ensemble}Ensemble and Phase Space}
284  
285 < \subsubsection{\label{introSection:ergodic}The Ergodic Hypothesis}
285 > \subsection{\label{introSection:ergodic}The Ergodic Hypothesis}
286  
287 < \subsection{\label{introSection:rigidBody}Dynamics of Rigid Bodies}
288 <
289 < \subsection{\label{introSection:correlationFunctions}Correlation Functions}
287 > Various thermodynamic properties can be calculated from Molecular
288 > Dynamics simulation. By comparing experimental values with the
289 > calculated properties, one can determine the accuracy of the
290 > simulation and the quality of the underlying model. However, both of
291 > experiment and computer simulation are usually performed during a
292 > certain time interval and the measurements are averaged over a
293 > period of them which is different from the average behavior of
294 > many-body system in Statistical Mechanics. Fortunately, Ergodic
295 > Hypothesis is proposed to make a connection between time average and
296 > ensemble average. It states that time average and average over the
297 > statistical ensemble are identical \cite{Frenkel1996, leach01:mm}.
298 > \begin{equation}
299 > \langle A \rangle_t = \mathop {\lim }\limits_{t \to \infty }
300 > \frac{1}{t}\int\limits_0^t {A(p(t),q(t))dt = \int\limits_\Gamma
301 > {A(p(t),q(t))} } \rho (p(t), q(t)) dpdq
302 > \end{equation}
303 > where $\langle A \rangle_t$ is an equilibrium value of a physical
304 > quantity and $\rho (p(t), q(t))$ is the equilibrium distribution
305 > function. If an observation is averaged over a sufficiently long
306 > time (longer than relaxation time), all accessible microstates in
307 > phase space are assumed to be equally probed, giving a properly
308 > weighted statistical average. This allows the researcher freedom of
309 > choice when deciding how best to measure a given observable. In case
310 > an ensemble averaged approach sounds most reasonable, the Monte
311 > Carlo techniques\cite{metropolis:1949} can be utilized. Or if the
312 > system lends itself to a time averaging approach, the Molecular
313 > Dynamics techniques in Sec.~\ref{introSection:molecularDynamics}
314 > will be the best choice\cite{Frenkel1996}.
315  
316 + \section{\label{introSection:molecularDynamics}Molecular Dynamics}
317 +
318 + As a special discipline of molecular modeling, Molecular dynamics
319 + has proven to be a powerful tool for studying the functions of
320 + biological systems, providing structural, thermodynamic and
321 + dynamical information.
322 +
323 + \subsection{\label{introSec:mdInit}Initialization}
324 +
325 + \subsection{\label{introSection:mdIntegration} Integration of the Equations of Motion}
326 +
327 + \section{\label{introSection:rigidBody}Dynamics of Rigid Bodies}
328 +
329 + A rigid body is a body in which the distance between any two given
330 + points of a rigid body remains constant regardless of external
331 + forces exerted on it. A rigid body therefore conserves its shape
332 + during its motion.
333 +
334 + Applications of dynamics of rigid bodies.
335 +
336 + \subsection{\label{introSection:lieAlgebra}Lie Algebra}
337 +
338 + \subsection{\label{introSection:DLMMotionEquation}The Euler Equations of Rigid Body Motion}
339 +
340 + \subsection{\label{introSection:otherRBMotionEquation}Other Formulations for Rigid Body Motion}
341 +
342 + %\subsection{\label{introSection:poissonBrackets}Poisson Brackets}
343 +
344 + \section{\label{introSection:correlationFunctions}Correlation Functions}
345 +
346   \section{\label{introSection:langevinDynamics}Langevin Dynamics}
347  
348 + \subsection{\label{introSection:LDIntroduction}Introduction and application of Langevin Dynamics}
349 +
350   \subsection{\label{introSection:generalizedLangevinDynamics}Generalized Langevin Dynamics}
351  
352 < \subsection{\label{introSection:hydroynamics}Hydrodynamics}
352 > \begin{equation}
353 > H = \frac{{p^2 }}{{2m}} + U(x) + H_B  + \Delta U(x,x_1 , \ldots x_N)
354 > \label{introEquation:bathGLE}
355 > \end{equation}
356 > where $H_B$ is harmonic bath Hamiltonian,
357 > \[
358 > H_B  =\sum\limits_{\alpha  = 1}^N {\left\{ {\frac{{p_\alpha ^2
359 > }}{{2m_\alpha  }} + \frac{1}{2}m_\alpha  w_\alpha ^2 } \right\}}
360 > \]
361 > and $\Delta U$ is bilinear system-bath coupling,
362 > \[
363 > \Delta U =  - \sum\limits_{\alpha  = 1}^N {g_\alpha  x_\alpha  x}
364 > \]
365 > Completing the square,
366 > \[
367 > H_B  + \Delta U = \sum\limits_{\alpha  = 1}^N {\left\{
368 > {\frac{{p_\alpha ^2 }}{{2m_\alpha  }} + \frac{1}{2}m_\alpha
369 > w_\alpha ^2 \left( {x_\alpha   - \frac{{g_\alpha  }}{{m_\alpha
370 > w_\alpha ^2 }}x} \right)^2 } \right\}}  - \sum\limits_{\alpha  =
371 > 1}^N {\frac{{g_\alpha ^2 }}{{2m_\alpha  w_\alpha ^2 }}} x^2
372 > \]
373 > and putting it back into Eq.~\ref{introEquation:bathGLE},
374 > \[
375 > H = \frac{{p^2 }}{{2m}} + W(x) + \sum\limits_{\alpha  = 1}^N
376 > {\left\{ {\frac{{p_\alpha ^2 }}{{2m_\alpha  }} + \frac{1}{2}m_\alpha
377 > w_\alpha ^2 \left( {x_\alpha   - \frac{{g_\alpha  }}{{m_\alpha
378 > w_\alpha ^2 }}x} \right)^2 } \right\}}
379 > \]
380 > where
381 > \[
382 > W(x) = U(x) - \sum\limits_{\alpha  = 1}^N {\frac{{g_\alpha ^2
383 > }}{{2m_\alpha  w_\alpha ^2 }}} x^2
384 > \]
385 > Since the first two terms of the new Hamiltonian depend only on the
386 > system coordinates, we can get the equations of motion for
387 > Generalized Langevin Dynamics by Hamilton's equations
388 > \ref{introEquation:motionHamiltonianCoordinate,
389 > introEquation:motionHamiltonianMomentum},
390 > \begin{align}
391 > \dot p &=  - \frac{{\partial H}}{{\partial x}}
392 >       &= m\ddot x
393 >       &= - \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)}
394 > \label{introEq:Lp5}
395 > \end{align}
396 > , and
397 > \begin{align}
398 > \dot p_\alpha   &=  - \frac{{\partial H}}{{\partial x_\alpha  }}
399 >                &= m\ddot x_\alpha
400 >                &= \- m_\alpha  w_\alpha ^2 \left( {x_\alpha   - \frac{{g_\alpha}}{{m_\alpha  w_\alpha ^2 }}x} \right)
401 > \end{align}
402 >
403 > \subsection{\label{introSection:laplaceTransform}The Laplace Transform}
404 >
405 > \[
406 > L(x) = \int_0^\infty  {x(t)e^{ - pt} dt}
407 > \]
408 >
409 > \[
410 > L(x + y) = L(x) + L(y)
411 > \]
412 >
413 > \[
414 > L(ax) = aL(x)
415 > \]
416 >
417 > \[
418 > L(\dot x) = pL(x) - px(0)
419 > \]
420 >
421 > \[
422 > L(\ddot x) = p^2 L(x) - px(0) - \dot x(0)
423 > \]
424 >
425 > \[
426 > L\left( {\int_0^t {g(t - \tau )h(\tau )d\tau } } \right) = G(p)H(p)
427 > \]
428 >
429 > Some relatively important transformation,
430 > \[
431 > L(\cos at) = \frac{p}{{p^2  + a^2 }}
432 > \]
433 >
434 > \[
435 > L(\sin at) = \frac{a}{{p^2  + a^2 }}
436 > \]
437 >
438 > \[
439 > L(1) = \frac{1}{p}
440 > \]
441 >
442 > First, the bath coordinates,
443 > \[
444 > p^2 L(x_\alpha  ) - px_\alpha  (0) - \dot x_\alpha  (0) =  - \omega
445 > _\alpha ^2 L(x_\alpha  ) + \frac{{g_\alpha  }}{{\omega _\alpha
446 > }}L(x)
447 > \]
448 > \[
449 > L(x_\alpha  ) = \frac{{\frac{{g_\alpha  }}{{\omega _\alpha  }}L(x) +
450 > px_\alpha  (0) + \dot x_\alpha  (0)}}{{p^2  + \omega _\alpha ^2 }}
451 > \]
452 > Then, the system coordinates,
453 > \begin{align}
454 > mL(\ddot x) &=  - \frac{1}{p}\frac{{\partial W(x)}}{{\partial x}} -
455 > \sum\limits_{\alpha  = 1}^N {\left\{ {\frac{{\frac{{g_\alpha
456 > }}{{\omega _\alpha  }}L(x) + px_\alpha  (0) + \dot x_\alpha
457 > (0)}}{{p^2  + \omega _\alpha ^2 }} - \frac{{g_\alpha ^2 }}{{m_\alpha
458 > }}\omega _\alpha ^2 L(x)} \right\}}
459 > %
460 > &= - \frac{1}{p}\frac{{\partial W(x)}}{{\partial x}} -
461 > \sum\limits_{\alpha  = 1}^N {\left\{ { - \frac{{g_\alpha ^2 }}{{m_\alpha  \omega _\alpha ^2 }}\frac{p}{{p^2  + \omega _\alpha ^2 }}pL(x)
462 > - \frac{p}{{p^2  + \omega _\alpha ^2 }}g_\alpha  x_\alpha  (0)
463 > - \frac{1}{{p^2  + \omega _\alpha ^2 }}g_\alpha  \dot x_\alpha  (0)} \right\}}
464 > \end{align}
465 > Then, the inverse transform,
466 >
467 > \begin{align}
468 > m\ddot x &=  - \frac{{\partial W(x)}}{{\partial x}} -
469 > \sum\limits_{\alpha  = 1}^N {\left\{ {\left( { - \frac{{g_\alpha ^2
470 > }}{{m_\alpha  \omega _\alpha ^2 }}} \right)\int_0^t {\cos (\omega
471 > _\alpha  t)\dot x(t - \tau )d\tau  - \left[ {g_\alpha  x_\alpha  (0)
472 > - \frac{{g_\alpha  }}{{m_\alpha  \omega _\alpha  }}} \right]\cos
473 > (\omega _\alpha  t) - \frac{{g_\alpha  \dot x_\alpha  (0)}}{{\omega
474 > _\alpha  }}\sin (\omega _\alpha  t)} } \right\}}
475 > %
476 > &= - \frac{{\partial W(x)}}{{\partial x}} - \int_0^t
477 > {\sum\limits_{\alpha  = 1}^N {\left( { - \frac{{g_\alpha ^2
478 > }}{{m_\alpha  \omega _\alpha ^2 }}} \right)\cos (\omega _\alpha
479 > t)\dot x(t - \tau )d} \tau }  + \sum\limits_{\alpha  = 1}^N {\left\{
480 > {\left[ {g_\alpha  x_\alpha  (0) - \frac{{g_\alpha  }}{{m_\alpha
481 > \omega _\alpha  }}} \right]\cos (\omega _\alpha  t) +
482 > \frac{{g_\alpha  \dot x_\alpha  (0)}}{{\omega _\alpha  }}\sin
483 > (\omega _\alpha  t)} \right\}}
484 > \end{align}
485 >
486 > \begin{equation}
487 > m\ddot x =  - \frac{{\partial W}}{{\partial x}} - \int_0^t {\xi
488 > (t)\dot x(t - \tau )d\tau }  + R(t)
489 > \label{introEuqation:GeneralizedLangevinDynamics}
490 > \end{equation}
491 > %where $ {\xi (t)}$ is friction kernel, $R(t)$ is random force and
492 > %$W$ is the potential of mean force. $W(x) =  - kT\ln p(x)$
493 > \[
494 > \xi (t) = \sum\limits_{\alpha  = 1}^N {\left( { - \frac{{g_\alpha ^2
495 > }}{{m_\alpha  \omega _\alpha ^2 }}} \right)\cos (\omega _\alpha  t)}
496 > \]
497 > For an infinite harmonic bath, we can use the spectral density and
498 > an integral over frequencies.
499 >
500 > \[
501 > R(t) = \sum\limits_{\alpha  = 1}^N {\left( {g_\alpha  x_\alpha  (0)
502 > - \frac{{g_\alpha ^2 }}{{m_\alpha  \omega _\alpha ^2 }}x(0)}
503 > \right)\cos (\omega _\alpha  t)}  + \frac{{\dot x_\alpha
504 > (0)}}{{\omega _\alpha  }}\sin (\omega _\alpha  t)
505 > \]
506 > The random forces depend only on initial conditions.
507 >
508 > \subsubsection{\label{introSection:secondFluctuationDissipation}The Second Fluctuation Dissipation Theorem}
509 > So we can define a new set of coordinates,
510 > \[
511 > q_\alpha  (t) = x_\alpha  (t) - \frac{1}{{m_\alpha  \omega _\alpha
512 > ^2 }}x(0)
513 > \]
514 > This makes
515 > \[
516 > R(t) = \sum\limits_{\alpha  = 1}^N {g_\alpha  q_\alpha  (t)}
517 > \]
518 > And since the $q$ coordinates are harmonic oscillators,
519 > \[
520 > \begin{array}{l}
521 > \left\langle {q_\alpha  (t)q_\alpha  (0)} \right\rangle  = \left\langle {q_\alpha ^2 (0)} \right\rangle \cos (\omega _\alpha  t) \\
522 > \left\langle {q_\alpha  (t)q_\beta  (0)} \right\rangle  = \delta _{\alpha \beta } \left\langle {q_\alpha  (t)q_\alpha  (0)} \right\rangle  \\
523 > \end{array}
524 > \]
525 >
526 > \begin{align}
527 > \left\langle {R(t)R(0)} \right\rangle  &= \sum\limits_\alpha
528 > {\sum\limits_\beta  {g_\alpha  g_\beta  \left\langle {q_\alpha
529 > (t)q_\beta  (0)} \right\rangle } }
530 > %
531 > &= \sum\limits_\alpha  {g_\alpha ^2 \left\langle {q_\alpha ^2 (0)}
532 > \right\rangle \cos (\omega _\alpha  t)}
533 > %
534 > &= kT\xi (t)
535 > \end{align}
536 >
537 > \begin{equation}
538 > \xi (t) = \left\langle {R(t)R(0)} \right\rangle
539 > \label{introEquation:secondFluctuationDissipation}
540 > \end{equation}
541 >
542 > \section{\label{introSection:hydroynamics}Hydrodynamics}
543 >
544 > \subsection{\label{introSection:frictionTensor} Friction Tensor}
545 > \subsection{\label{introSection:analyticalApproach}Analytical
546 > Approach}
547 >
548 > \subsection{\label{introSection:approximationApproach}Approximation
549 > Approach}
550 >
551 > \subsection{\label{introSection:centersRigidBody}Centers of Rigid
552 > Body}

Diff Legend

Removed lines
+ Added lines
< Changed lines
> Changed lines