1 |
\chapter{\label{chapt:introduction}INTRODUCTION AND THEORETICAL BACKGROUND} |
2 |
|
3 |
\section{\label{introSection:classicalMechanics}Classical |
4 |
Mechanics} |
5 |
|
6 |
Using equations of motion derived from Classical Mechanics, |
7 |
Molecular Dynamics simulations are carried out by integrating the |
8 |
equations of motion for a given system of particles. There are three |
9 |
fundamental ideas behind classical mechanics. Firstly, one can |
10 |
determine the state of a mechanical system at any time of interest; |
11 |
Secondly, all the mechanical properties of the system at that time |
12 |
can be determined by combining the knowledge of the properties of |
13 |
the system with the specification of this state; Finally, the |
14 |
specification of the state when further combined with the laws of |
15 |
mechanics will also be sufficient to predict the future behavior of |
16 |
the system. |
17 |
|
18 |
\subsection{\label{introSection:newtonian}Newtonian Mechanics} |
19 |
The discovery of Newton's three laws of mechanics which govern the |
20 |
motion of particles is the foundation of the classical mechanics. |
21 |
Newton's first law defines a class of inertial frames. Inertial |
22 |
frames are reference frames where a particle not interacting with |
23 |
other bodies will move with constant speed in the same direction. |
24 |
With respect to inertial frames, Newton's second law has the form |
25 |
\begin{equation} |
26 |
F = \frac {dp}{dt} = \frac {mdv}{dt} |
27 |
\label{introEquation:newtonSecondLaw} |
28 |
\end{equation} |
29 |
A point mass interacting with other bodies moves with the |
30 |
acceleration along the direction of the force acting on it. Let |
31 |
$F_{ij}$ be the force that particle $i$ exerts on particle $j$, and |
32 |
$F_{ji}$ be the force that particle $j$ exerts on particle $i$. |
33 |
Newton's third law states that |
34 |
\begin{equation} |
35 |
F_{ij} = -F_{ji}. |
36 |
\label{introEquation:newtonThirdLaw} |
37 |
\end{equation} |
38 |
Conservation laws of Newtonian Mechanics play very important roles |
39 |
in solving mechanics problems. The linear momentum of a particle is |
40 |
conserved if it is free or it experiences no force. The second |
41 |
conservation theorem concerns the angular momentum of a particle. |
42 |
The angular momentum $L$ of a particle with respect to an origin |
43 |
from which $r$ is measured is defined to be |
44 |
\begin{equation} |
45 |
L \equiv r \times p \label{introEquation:angularMomentumDefinition} |
46 |
\end{equation} |
47 |
The torque $\tau$ with respect to the same origin is defined to be |
48 |
\begin{equation} |
49 |
\tau \equiv r \times F \label{introEquation:torqueDefinition} |
50 |
\end{equation} |
51 |
Differentiating Eq.~\ref{introEquation:angularMomentumDefinition}, |
52 |
\[ |
53 |
\dot L = \frac{d}{{dt}}(r \times p) = (\dot r \times p) + (r \times |
54 |
\dot p) |
55 |
\] |
56 |
since |
57 |
\[ |
58 |
\dot r \times p = \dot r \times mv = m\dot r \times \dot r \equiv 0 |
59 |
\] |
60 |
thus, |
61 |
\begin{equation} |
62 |
\dot L = r \times \dot p = \tau |
63 |
\end{equation} |
64 |
If there are no external torques acting on a body, the angular |
65 |
momentum of it is conserved. The last conservation theorem state |
66 |
that if all forces are conservative, energy is conserved, |
67 |
\begin{equation}E = T + V. \label{introEquation:energyConservation} |
68 |
\end{equation} |
69 |
All of these conserved quantities are important factors to determine |
70 |
the quality of numerical integration schemes for rigid |
71 |
bodies.\cite{Dullweber1997} |
72 |
|
73 |
\subsection{\label{introSection:lagrangian}Lagrangian Mechanics} |
74 |
|
75 |
Newtonian Mechanics suffers from an important limitation: motion can |
76 |
only be described in cartesian coordinate systems which make it |
77 |
impossible to predict analytically the properties of the system even |
78 |
if we know all of the details of the interaction. In order to |
79 |
overcome some of the practical difficulties which arise in attempts |
80 |
to apply Newton's equation to complex systems, approximate numerical |
81 |
procedures may be developed. |
82 |
|
83 |
\subsubsection{\label{introSection:halmiltonPrinciple}\textbf{Hamilton's |
84 |
Principle}} |
85 |
|
86 |
Hamilton introduced the dynamical principle upon which it is |
87 |
possible to base all of mechanics and most of classical physics. |
88 |
Hamilton's Principle may be stated as follows: the trajectory, along |
89 |
which a dynamical system may move from one point to another within a |
90 |
specified time, is derived by finding the path which minimizes the |
91 |
time integral of the difference between the kinetic $K$, and |
92 |
potential energies $U$, |
93 |
\begin{equation} |
94 |
\delta \int_{t_1 }^{t_2 } {(K - U)dt = 0}. |
95 |
\label{introEquation:halmitonianPrinciple1} |
96 |
\end{equation} |
97 |
For simple mechanical systems, where the forces acting on the |
98 |
different parts are derivable from a potential, the Lagrangian |
99 |
function $L$ can be defined as the difference between the kinetic |
100 |
energy of the system and its potential energy, |
101 |
\begin{equation} |
102 |
L \equiv K - U = L(q_i ,\dot q_i ). |
103 |
\label{introEquation:lagrangianDef} |
104 |
\end{equation} |
105 |
Thus, Eq.~\ref{introEquation:halmitonianPrinciple1} becomes |
106 |
\begin{equation} |
107 |
\delta \int_{t_1 }^{t_2 } {L dt = 0} . |
108 |
\label{introEquation:halmitonianPrinciple2} |
109 |
\end{equation} |
110 |
|
111 |
\subsubsection{\label{introSection:equationOfMotionLagrangian}\textbf{The |
112 |
Equations of Motion in Lagrangian Mechanics}} |
113 |
|
114 |
For a system of $f$ degrees of freedom, the equations of motion in |
115 |
the Lagrangian form is |
116 |
\begin{equation} |
117 |
\frac{d}{{dt}}\frac{{\partial L}}{{\partial \dot q_i }} - |
118 |
\frac{{\partial L}}{{\partial q_i }} = 0,{\rm{ }}i = 1, \ldots,f |
119 |
\label{introEquation:eqMotionLagrangian} |
120 |
\end{equation} |
121 |
where $q_{i}$ is generalized coordinate and $\dot{q_{i}}$ is |
122 |
generalized velocity. |
123 |
|
124 |
\subsection{\label{introSection:hamiltonian}Hamiltonian Mechanics} |
125 |
|
126 |
Arising from Lagrangian Mechanics, Hamiltonian Mechanics was |
127 |
introduced by William Rowan Hamilton in 1833 as a re-formulation of |
128 |
classical mechanics. If the potential energy of a system is |
129 |
independent of velocities, the momenta can be defined as |
130 |
\begin{equation} |
131 |
p_i = \frac{\partial L}{\partial \dot q_i} |
132 |
\label{introEquation:generalizedMomenta} |
133 |
\end{equation} |
134 |
The Lagrange equations of motion are then expressed by |
135 |
\begin{equation} |
136 |
p_i = \frac{{\partial L}}{{\partial q_i }} |
137 |
\label{introEquation:generalizedMomentaDot} |
138 |
\end{equation} |
139 |
With the help of the generalized momenta, we may now define a new |
140 |
quantity $H$ by the equation |
141 |
\begin{equation} |
142 |
H = \sum\limits_k {p_k \dot q_k } - L , |
143 |
\label{introEquation:hamiltonianDefByLagrangian} |
144 |
\end{equation} |
145 |
where $ \dot q_1 \ldots \dot q_f $ are generalized velocities and |
146 |
$L$ is the Lagrangian function for the system. Differentiating |
147 |
Eq.~\ref{introEquation:hamiltonianDefByLagrangian}, one can obtain |
148 |
\begin{equation} |
149 |
dH = \sum\limits_k {\left( {p_k d\dot q_k + \dot q_k dp_k - |
150 |
\frac{{\partial L}}{{\partial q_k }}dq_k - \frac{{\partial |
151 |
L}}{{\partial \dot q_k }}d\dot q_k } \right)} - \frac{{\partial |
152 |
L}}{{\partial t}}dt . \label{introEquation:diffHamiltonian1} |
153 |
\end{equation} |
154 |
Making use of Eq.~\ref{introEquation:generalizedMomenta}, the second |
155 |
and fourth terms in the parentheses cancel. Therefore, |
156 |
Eq.~\ref{introEquation:diffHamiltonian1} can be rewritten as |
157 |
\begin{equation} |
158 |
dH = \sum\limits_k {\left( {\dot q_k dp_k - \dot p_k dq_k } |
159 |
\right)} - \frac{{\partial L}}{{\partial t}}dt . |
160 |
\label{introEquation:diffHamiltonian2} |
161 |
\end{equation} |
162 |
By identifying the coefficients of $dq_k$, $dp_k$ and dt, we can |
163 |
find |
164 |
\begin{equation} |
165 |
\frac{{\partial H}}{{\partial p_k }} = \dot {q_k} |
166 |
\label{introEquation:motionHamiltonianCoordinate} |
167 |
\end{equation} |
168 |
\begin{equation} |
169 |
\frac{{\partial H}}{{\partial q_k }} = - \dot {p_k} |
170 |
\label{introEquation:motionHamiltonianMomentum} |
171 |
\end{equation} |
172 |
and |
173 |
\begin{equation} |
174 |
\frac{{\partial H}}{{\partial t}} = - \frac{{\partial L}}{{\partial |
175 |
t}} |
176 |
\label{introEquation:motionHamiltonianTime} |
177 |
\end{equation} |
178 |
where Eq.~\ref{introEquation:motionHamiltonianCoordinate} and |
179 |
Eq.~\ref{introEquation:motionHamiltonianMomentum} are Hamilton's |
180 |
equation of motion. Due to their symmetrical formula, they are also |
181 |
known as the canonical equations of motions.\cite{Goldstein2001} |
182 |
|
183 |
An important difference between Lagrangian approach and the |
184 |
Hamiltonian approach is that the Lagrangian is considered to be a |
185 |
function of the generalized velocities $\dot q_i$ and coordinates |
186 |
$q_i$, while the Hamiltonian is considered to be a function of the |
187 |
generalized momenta $p_i$ and the conjugate coordinates $q_i$. |
188 |
Hamiltonian Mechanics is more appropriate for application to |
189 |
statistical mechanics and quantum mechanics, since it treats the |
190 |
coordinate and its time derivative as independent variables and it |
191 |
only works with 1st-order differential equations.\cite{Marion1990} |
192 |
In Newtonian Mechanics, a system described by conservative forces |
193 |
conserves the total energy |
194 |
(Eq.~\ref{introEquation:energyConservation}). It follows that |
195 |
Hamilton's equations of motion conserve the total Hamiltonian |
196 |
\begin{equation} |
197 |
\frac{{dH}}{{dt}} = \sum\limits_i {\left( {\frac{{\partial |
198 |
H}}{{\partial q_i }}\dot q_i + \frac{{\partial H}}{{\partial p_i |
199 |
}}\dot p_i } \right)} = \sum\limits_i {\left( {\frac{{\partial |
200 |
H}}{{\partial q_i }}\frac{{\partial H}}{{\partial p_i }} - |
201 |
\frac{{\partial H}}{{\partial p_i }}\frac{{\partial H}}{{\partial |
202 |
q_i }}} \right) = 0}. \label{introEquation:conserveHalmitonian} |
203 |
\end{equation} |
204 |
|
205 |
\section{\label{introSection:statisticalMechanics}Statistical |
206 |
Mechanics} |
207 |
|
208 |
The thermodynamic behaviors and properties of Molecular Dynamics |
209 |
simulation are governed by the principle of Statistical Mechanics. |
210 |
The following section will give a brief introduction to some of the |
211 |
Statistical Mechanics concepts and theorems presented in this |
212 |
dissertation. |
213 |
|
214 |
\subsection{\label{introSection:ensemble}Phase Space and Ensemble} |
215 |
|
216 |
Mathematically, phase space is the space which represents all |
217 |
possible states of a system. Each possible state of the system |
218 |
corresponds to one unique point in the phase space. For mechanical |
219 |
systems, the phase space usually consists of all possible values of |
220 |
position and momentum variables. Consider a dynamic system of $f$ |
221 |
particles in a cartesian space, where each of the $6f$ coordinates |
222 |
and momenta is assigned to one of $6f$ mutually orthogonal axes, the |
223 |
phase space of this system is a $6f$ dimensional space. A point, $x |
224 |
= |
225 |
(\mathord{\buildrel{\lower3pt\hbox{$\scriptscriptstyle\rightharpoonup$}} |
226 |
\over q} _1 , \ldots |
227 |
,\mathord{\buildrel{\lower3pt\hbox{$\scriptscriptstyle\rightharpoonup$}} |
228 |
\over q} _f |
229 |
,\mathord{\buildrel{\lower3pt\hbox{$\scriptscriptstyle\rightharpoonup$}} |
230 |
\over p} _1 \ldots |
231 |
,\mathord{\buildrel{\lower3pt\hbox{$\scriptscriptstyle\rightharpoonup$}} |
232 |
\over p} _f )$ , with a unique set of values of $6f$ coordinates and |
233 |
momenta is a phase space vector. |
234 |
%%%fix me |
235 |
|
236 |
In statistical mechanics, the condition of an ensemble at any time |
237 |
can be regarded as appropriately specified by the density $\rho$ |
238 |
with which representative points are distributed over the phase |
239 |
space. The density distribution for an ensemble with $f$ degrees of |
240 |
freedom is defined as, |
241 |
\begin{equation} |
242 |
\rho = \rho (q_1 , \ldots ,q_f ,p_1 , \ldots ,p_f ,t). |
243 |
\label{introEquation:densityDistribution} |
244 |
\end{equation} |
245 |
Governed by the principles of mechanics, the phase points change |
246 |
their locations which changes the density at any time at phase |
247 |
space. Hence, the density distribution is also to be taken as a |
248 |
function of the time. The number of systems $\delta N$ at time $t$ |
249 |
can be determined by, |
250 |
\begin{equation} |
251 |
\delta N = \rho (q,p,t)dq_1 \ldots dq_f dp_1 \ldots dp_f. |
252 |
\label{introEquation:deltaN} |
253 |
\end{equation} |
254 |
Assuming enough copies of the systems, we can sufficiently |
255 |
approximate $\delta N$ without introducing discontinuity when we go |
256 |
from one region in the phase space to another. By integrating over |
257 |
the whole phase space, |
258 |
\begin{equation} |
259 |
N = \int { \ldots \int {\rho (q,p,t)dq_1 } ...dq_f dp_1 } ...dp_f |
260 |
\label{introEquation:totalNumberSystem} |
261 |
\end{equation} |
262 |
gives us an expression for the total number of copies. Hence, the |
263 |
probability per unit volume in the phase space can be obtained by, |
264 |
\begin{equation} |
265 |
\frac{{\rho (q,p,t)}}{N} = \frac{{\rho (q,p,t)}}{{\int { \ldots \int |
266 |
{\rho (q,p,t)dq_1 } ...dq_f dp_1 } ...dp_f }}. |
267 |
\label{introEquation:unitProbability} |
268 |
\end{equation} |
269 |
With the help of Eq.~\ref{introEquation:unitProbability} and the |
270 |
knowledge of the system, it is possible to calculate the average |
271 |
value of any desired quantity which depends on the coordinates and |
272 |
momenta of the system. Even when the dynamics of the real system are |
273 |
complex, or stochastic, or even discontinuous, the average |
274 |
properties of the ensemble of possibilities as a whole remain well |
275 |
defined. For a classical system in thermal equilibrium with its |
276 |
environment, the ensemble average of a mechanical quantity, $\langle |
277 |
A(q , p) \rangle_t$, takes the form of an integral over the phase |
278 |
space of the system, |
279 |
\begin{equation} |
280 |
\langle A(q , p) \rangle_t = \frac{{\int { \ldots \int {A(q,p)\rho |
281 |
(q,p,t)dq_1 } ...dq_f dp_1 } ...dp_f }}{{\int { \ldots \int {\rho |
282 |
(q,p,t)dq_1 } ...dq_f dp_1 } ...dp_f }}. |
283 |
\label{introEquation:ensembelAverage} |
284 |
\end{equation} |
285 |
|
286 |
\subsection{\label{introSection:liouville}Liouville's theorem} |
287 |
|
288 |
Liouville's theorem is the foundation on which statistical mechanics |
289 |
rests. It describes the time evolution of the phase space |
290 |
distribution function. In order to calculate the rate of change of |
291 |
$\rho$, we begin from Eq.~\ref{introEquation:deltaN}. If we consider |
292 |
the two faces perpendicular to the $q_1$ axis, which are located at |
293 |
$q_1$ and $q_1 + \delta q_1$, the number of phase points leaving the |
294 |
opposite face is given by the expression, |
295 |
\begin{equation} |
296 |
\left( {\rho + \frac{{\partial \rho }}{{\partial q_1 }}\delta q_1 } |
297 |
\right)\left( {\dot q_1 + \frac{{\partial \dot q_1 }}{{\partial q_1 |
298 |
}}\delta q_1 } \right)\delta q_2 \ldots \delta q_f \delta p_1 |
299 |
\ldots \delta p_f . |
300 |
\end{equation} |
301 |
Summing all over the phase space, we obtain |
302 |
\begin{equation} |
303 |
\frac{{d(\delta N)}}{{dt}} = - \sum\limits_{i = 1}^f {\left[ {\rho |
304 |
\left( {\frac{{\partial \dot q_i }}{{\partial q_i }} + |
305 |
\frac{{\partial \dot p_i }}{{\partial p_i }}} \right) + \left( |
306 |
{\frac{{\partial \rho }}{{\partial q_i }}\dot q_i + \frac{{\partial |
307 |
\rho }}{{\partial p_i }}\dot p_i } \right)} \right]} \delta q_1 |
308 |
\ldots \delta q_f \delta p_1 \ldots \delta p_f . |
309 |
\end{equation} |
310 |
Differentiating the equations of motion in Hamiltonian formalism |
311 |
(\ref{introEquation:motionHamiltonianCoordinate}, |
312 |
\ref{introEquation:motionHamiltonianMomentum}), we can show, |
313 |
\begin{equation} |
314 |
\sum\limits_i {\left( {\frac{{\partial \dot q_i }}{{\partial q_i }} |
315 |
+ \frac{{\partial \dot p_i }}{{\partial p_i }}} \right)} = 0 , |
316 |
\end{equation} |
317 |
which cancels the first terms of the right hand side. Furthermore, |
318 |
dividing $ \delta q_1 \ldots \delta q_f \delta p_1 \ldots \delta |
319 |
p_f $ in both sides, we can write out Liouville's theorem in a |
320 |
simple form, |
321 |
\begin{equation} |
322 |
\frac{{\partial \rho }}{{\partial t}} + \sum\limits_{i = 1}^f |
323 |
{\left( {\frac{{\partial \rho }}{{\partial q_i }}\dot q_i + |
324 |
\frac{{\partial \rho }}{{\partial p_i }}\dot p_i } \right)} = 0 . |
325 |
\label{introEquation:liouvilleTheorem} |
326 |
\end{equation} |
327 |
Liouville's theorem states that the distribution function is |
328 |
constant along any trajectory in phase space. In classical |
329 |
statistical mechanics, since the number of system copies in an |
330 |
ensemble is huge and constant, we can assume the local density has |
331 |
no reason (other than classical mechanics) to change, |
332 |
\begin{equation} |
333 |
\frac{{\partial \rho }}{{\partial t}} = 0. |
334 |
\label{introEquation:stationary} |
335 |
\end{equation} |
336 |
In such stationary system, the density of distribution $\rho$ can be |
337 |
connected to the Hamiltonian $H$ through Maxwell-Boltzmann |
338 |
distribution, |
339 |
\begin{equation} |
340 |
\rho \propto e^{ - \beta H} |
341 |
\label{introEquation:densityAndHamiltonian} |
342 |
\end{equation} |
343 |
|
344 |
\subsubsection{\label{introSection:phaseSpaceConservation}\textbf{Conservation of Phase Space}} |
345 |
Lets consider a region in the phase space, |
346 |
\begin{equation} |
347 |
\delta v = \int { \ldots \int {dq_1 } ...dq_f dp_1 } ..dp_f . |
348 |
\end{equation} |
349 |
If this region is small enough, the density $\rho$ can be regarded |
350 |
as uniform over the whole integral. Thus, the number of phase points |
351 |
inside this region is given by, |
352 |
\begin{equation} |
353 |
\delta N = \rho \delta v = \rho \int { \ldots \int {dq_1 } ...dq_f |
354 |
dp_1 } ..dp_f. |
355 |
\end{equation} |
356 |
|
357 |
\begin{equation} |
358 |
\frac{{d(\delta N)}}{{dt}} = \frac{{d\rho }}{{dt}}\delta v + \rho |
359 |
\frac{d}{{dt}}(\delta v) = 0. |
360 |
\end{equation} |
361 |
With the help of the stationary assumption |
362 |
(Eq.~\ref{introEquation:stationary}), we obtain the principle of |
363 |
\emph{conservation of volume in phase space}, |
364 |
\begin{equation} |
365 |
\frac{d}{{dt}}(\delta v) = \frac{d}{{dt}}\int { \ldots \int {dq_1 } |
366 |
...dq_f dp_1 } ..dp_f = 0. |
367 |
\label{introEquation:volumePreserving} |
368 |
\end{equation} |
369 |
|
370 |
\subsubsection{\label{introSection:liouvilleInOtherForms}\textbf{Liouville's Theorem in Other Forms}} |
371 |
|
372 |
Liouville's theorem can be expressed in a variety of different forms |
373 |
which are convenient within different contexts. For any two function |
374 |
$F$ and $G$ of the coordinates and momenta of a system, the Poisson |
375 |
bracket $\{F,G\}$ is defined as |
376 |
\begin{equation} |
377 |
\left\{ {F,G} \right\} = \sum\limits_i {\left( {\frac{{\partial |
378 |
F}}{{\partial q_i }}\frac{{\partial G}}{{\partial p_i }} - |
379 |
\frac{{\partial F}}{{\partial p_i }}\frac{{\partial G}}{{\partial |
380 |
q_i }}} \right)}. |
381 |
\label{introEquation:poissonBracket} |
382 |
\end{equation} |
383 |
Substituting equations of motion in Hamiltonian formalism |
384 |
(Eq.~\ref{introEquation:motionHamiltonianCoordinate} , |
385 |
Eq.~\ref{introEquation:motionHamiltonianMomentum}) into |
386 |
(Eq.~\ref{introEquation:liouvilleTheorem}), we can rewrite |
387 |
Liouville's theorem using Poisson bracket notion, |
388 |
\begin{equation} |
389 |
\left( {\frac{{\partial \rho }}{{\partial t}}} \right) = - \left\{ |
390 |
{\rho ,H} \right\}. |
391 |
\label{introEquation:liouvilleTheromInPoissin} |
392 |
\end{equation} |
393 |
Moreover, the Liouville operator is defined as |
394 |
\begin{equation} |
395 |
iL = \sum\limits_{i = 1}^f {\left( {\frac{{\partial H}}{{\partial |
396 |
p_i }}\frac{\partial }{{\partial q_i }} - \frac{{\partial |
397 |
H}}{{\partial q_i }}\frac{\partial }{{\partial p_i }}} \right)} |
398 |
\label{introEquation:liouvilleOperator} |
399 |
\end{equation} |
400 |
In terms of Liouville operator, Liouville's equation can also be |
401 |
expressed as |
402 |
\begin{equation} |
403 |
\left( {\frac{{\partial \rho }}{{\partial t}}} \right) = - iL\rho |
404 |
\label{introEquation:liouvilleTheoremInOperator} |
405 |
\end{equation} |
406 |
which can help define a propagator $\rho (t) = e^{-iLt} \rho (0)$. |
407 |
\subsection{\label{introSection:ergodic}The Ergodic Hypothesis} |
408 |
|
409 |
Various thermodynamic properties can be calculated from Molecular |
410 |
Dynamics simulation. By comparing experimental values with the |
411 |
calculated properties, one can determine the accuracy of the |
412 |
simulation and the quality of the underlying model. However, both |
413 |
experiments and computer simulations are usually performed during a |
414 |
certain time interval and the measurements are averaged over a |
415 |
period of time which is different from the average behavior of |
416 |
many-body system in Statistical Mechanics. Fortunately, the Ergodic |
417 |
Hypothesis makes a connection between time average and the ensemble |
418 |
average. It states that the time average and average over the |
419 |
statistical ensemble are identical:\cite{Frenkel1996, Leach2001} |
420 |
\begin{equation} |
421 |
\langle A(q , p) \rangle_t = \mathop {\lim }\limits_{t \to \infty } |
422 |
\frac{1}{t}\int\limits_0^t {A(q(t),p(t))dt = \int\limits_\Gamma |
423 |
{A(q(t),p(t))} } \rho (q(t), p(t)) dqdp |
424 |
\end{equation} |
425 |
where $\langle A(q , p) \rangle_t$ is an equilibrium value of a |
426 |
physical quantity and $\rho (p(t), q(t))$ is the equilibrium |
427 |
distribution function. If an observation is averaged over a |
428 |
sufficiently long time (longer than the relaxation time), all |
429 |
accessible microstates in phase space are assumed to be equally |
430 |
probed, giving a properly weighted statistical average. This allows |
431 |
the researcher freedom of choice when deciding how best to measure a |
432 |
given observable. In case an ensemble averaged approach sounds most |
433 |
reasonable, the Monte Carlo methods\cite{Metropolis1949} can be |
434 |
utilized. Or if the system lends itself to a time averaging |
435 |
approach, the Molecular Dynamics techniques in |
436 |
Sec.~\ref{introSection:molecularDynamics} will be the best |
437 |
choice.\cite{Frenkel1996} |
438 |
|
439 |
\section{\label{introSection:geometricIntegratos}Geometric Integrators} |
440 |
A variety of numerical integrators have been proposed to simulate |
441 |
the motions of atoms in MD simulation. They usually begin with |
442 |
initial conditions and move the objects in the direction governed by |
443 |
the differential equations. However, most of them ignore the hidden |
444 |
physical laws contained within the equations. Since 1990, geometric |
445 |
integrators, which preserve various phase-flow invariants such as |
446 |
symplectic structure, volume and time reversal symmetry, were |
447 |
developed to address this issue.\cite{Dullweber1997, McLachlan1998, |
448 |
Leimkuhler1999} The velocity Verlet method, which happens to be a |
449 |
simple example of symplectic integrator, continues to gain |
450 |
popularity in the molecular dynamics community. This fact can be |
451 |
partly explained by its geometric nature. |
452 |
|
453 |
\subsection{\label{introSection:symplecticManifold}Symplectic Manifolds} |
454 |
A \emph{manifold} is an abstract mathematical space. It looks |
455 |
locally like Euclidean space, but when viewed globally, it may have |
456 |
more complicated structure. A good example of manifold is the |
457 |
surface of Earth. It seems to be flat locally, but it is round if |
458 |
viewed as a whole. A \emph{differentiable manifold} (also known as |
459 |
\emph{smooth manifold}) is a manifold on which it is possible to |
460 |
apply calculus.\cite{Hirsch1997} A \emph{symplectic manifold} is |
461 |
defined as a pair $(M, \omega)$ which consists of a |
462 |
\emph{differentiable manifold} $M$ and a close, non-degenerate, |
463 |
bilinear symplectic form, $\omega$. A symplectic form on a vector |
464 |
space $V$ is a function $\omega(x, y)$ which satisfies |
465 |
$\omega(\lambda_1x_1+\lambda_2x_2, y) = \lambda_1\omega(x_1, y)+ |
466 |
\lambda_2\omega(x_2, y)$, $\omega(x, y) = - \omega(y, x)$ and |
467 |
$\omega(x, x) = 0$.\cite{McDuff1998} The cross product operation in |
468 |
vector field is an example of symplectic form. One of the |
469 |
motivations to study \emph{symplectic manifolds} in Hamiltonian |
470 |
Mechanics is that a symplectic manifold can represent all possible |
471 |
configurations of the system and the phase space of the system can |
472 |
be described by it's cotangent bundle.\cite{Jost2002} Every |
473 |
symplectic manifold is even dimensional. For instance, in Hamilton |
474 |
equations, coordinate and momentum always appear in pairs. |
475 |
|
476 |
\subsection{\label{introSection:ODE}Ordinary Differential Equations} |
477 |
|
478 |
For an ordinary differential system defined as |
479 |
\begin{equation} |
480 |
\dot x = f(x) |
481 |
\end{equation} |
482 |
where $x = x(q,p)$, this system is a canonical Hamiltonian, if |
483 |
$f(x) = J\nabla _x H(x)$. Here, $H = H (q, p)$ is Hamiltonian |
484 |
function and $J$ is the skew-symmetric matrix |
485 |
\begin{equation} |
486 |
J = \left( {\begin{array}{*{20}c} |
487 |
0 & I \\ |
488 |
{ - I} & 0 \\ |
489 |
\end{array}} \right) |
490 |
\label{introEquation:canonicalMatrix} |
491 |
\end{equation} |
492 |
where $I$ is an identity matrix. Using this notation, Hamiltonian |
493 |
system can be rewritten as, |
494 |
\begin{equation} |
495 |
\frac{d}{{dt}}x = J\nabla _x H(x). |
496 |
\label{introEquation:compactHamiltonian} |
497 |
\end{equation}In this case, $f$ is |
498 |
called a \emph{Hamiltonian vector field}. Another generalization of |
499 |
Hamiltonian dynamics is Poisson Dynamics,\cite{Olver1986} |
500 |
\begin{equation} |
501 |
\dot x = J(x)\nabla _x H \label{introEquation:poissonHamiltonian} |
502 |
\end{equation} |
503 |
where the most obvious change being that matrix $J$ now depends on |
504 |
$x$. |
505 |
|
506 |
\subsection{\label{introSection:exactFlow}Exact Propagator} |
507 |
|
508 |
Let $x(t)$ be the exact solution of the ODE |
509 |
system, |
510 |
\begin{equation} |
511 |
\frac{{dx}}{{dt}} = f(x), \label{introEquation:ODE} |
512 |
\end{equation} we can |
513 |
define its exact propagator $\varphi_\tau$: |
514 |
\[ x(t+\tau) |
515 |
=\varphi_\tau(x(t)) |
516 |
\] |
517 |
where $\tau$ is a fixed time step and $\varphi$ is a map from phase |
518 |
space to itself. The propagator has the continuous group property, |
519 |
\begin{equation} |
520 |
\varphi _{\tau _1 } \circ \varphi _{\tau _2 } = \varphi _{\tau _1 |
521 |
+ \tau _2 } . |
522 |
\end{equation} |
523 |
In particular, |
524 |
\begin{equation} |
525 |
\varphi _\tau \circ \varphi _{ - \tau } = I |
526 |
\end{equation} |
527 |
Therefore, the exact propagator is self-adjoint, |
528 |
\begin{equation} |
529 |
\varphi _\tau = \varphi _{ - \tau }^{ - 1}. |
530 |
\end{equation} |
531 |
The exact propagator can also be written as an operator, |
532 |
\begin{equation} |
533 |
\varphi _\tau (x) = e^{\tau \sum\limits_i {f_i (x)\frac{\partial |
534 |
}{{\partial x_i }}} } (x) \equiv \exp (\tau f)(x). |
535 |
\label{introEquation:exponentialOperator} |
536 |
\end{equation} |
537 |
In most cases, it is not easy to find the exact propagator |
538 |
$\varphi_\tau$. Instead, we use an approximate map, $\psi_\tau$, |
539 |
which is usually called an integrator. The order of an integrator |
540 |
$\psi_\tau$ is $p$, if the Taylor series of $\psi_\tau$ agree to |
541 |
order $p$, |
542 |
\begin{equation} |
543 |
\psi_\tau(x) = x + \tau f(x) + O(\tau^{p+1}) |
544 |
\end{equation} |
545 |
|
546 |
\subsection{\label{introSection:geometricProperties}Geometric Properties} |
547 |
|
548 |
The hidden geometric properties\cite{Budd1999, Marsden1998} of an |
549 |
ODE and its propagator play important roles in numerical studies. |
550 |
Many of them can be found in systems which occur naturally in |
551 |
applications. Let $\varphi$ be the propagator of Hamiltonian vector |
552 |
field, $\varphi$ is a \emph{symplectic} propagator if it satisfies, |
553 |
\begin{equation} |
554 |
{\varphi '}^T J \varphi ' = J. |
555 |
\end{equation} |
556 |
According to Liouville's theorem, the symplectic volume is invariant |
557 |
under a Hamiltonian propagator, which is the basis for classical |
558 |
statistical mechanics. Furthermore, the propagator of a Hamiltonian |
559 |
vector field on a symplectic manifold can be shown to be a |
560 |
symplectomorphism. As to the Poisson system, |
561 |
\begin{equation} |
562 |
{\varphi '}^T J \varphi ' = J \circ \varphi |
563 |
\end{equation} |
564 |
is the property that must be preserved by the integrator. It is |
565 |
possible to construct a \emph{volume-preserving} propagator for a |
566 |
source free ODE ($ \nabla \cdot f = 0 $), if the propagator |
567 |
satisfies $ \det d\varphi = 1$. One can show easily that a |
568 |
symplectic propagator will be volume-preserving. Changing the |
569 |
variables $y = h(x)$ in an ODE (Eq.~\ref{introEquation:ODE}) will |
570 |
result in a new system, |
571 |
\[ |
572 |
\dot y = \tilde f(y) = ((dh \cdot f)h^{ - 1} )(y). |
573 |
\] |
574 |
The vector filed $f$ has reversing symmetry $h$ if $f = - \tilde f$. |
575 |
In other words, the propagator of this vector field is reversible if |
576 |
and only if $ h \circ \varphi ^{ - 1} = \varphi \circ h $. A |
577 |
conserved quantity of a general differential function is a function |
578 |
$ G:R^{2d} \to R^d $ which is constant for all solutions of the ODE |
579 |
$\frac{{dx}}{{dt}} = f(x)$ , |
580 |
\[ |
581 |
\frac{{dG(x(t))}}{{dt}} = 0. |
582 |
\] |
583 |
Using the chain rule, one may obtain, |
584 |
\[ |
585 |
\sum\limits_i {\frac{{dG}}{{dx_i }}} f_i (x) = f \cdot \nabla G, |
586 |
\] |
587 |
which is the condition for conserved quantities. For a canonical |
588 |
Hamiltonian system, the time evolution of an arbitrary smooth |
589 |
function $G$ is given by, |
590 |
\begin{eqnarray} |
591 |
\frac{{dG(x(t))}}{{dt}} & = & [\nabla _x G(x(t))]^T \dot x(t) \notag\\ |
592 |
& = & [\nabla _x G(x(t))]^T J\nabla _x H(x(t)). |
593 |
\label{introEquation:firstIntegral1} |
594 |
\end{eqnarray} |
595 |
Using poisson bracket notion, Eq.~\ref{introEquation:firstIntegral1} |
596 |
can be rewritten as |
597 |
\[ |
598 |
\frac{d}{{dt}}G(x(t)) = \left\{ {G,H} \right\}(x(t)). |
599 |
\] |
600 |
Therefore, the sufficient condition for $G$ to be a conserved |
601 |
quantity of a Hamiltonian system is $\left\{ {G,H} \right\} = 0.$ As |
602 |
is well known, the Hamiltonian (or energy) H of a Hamiltonian system |
603 |
is a conserved quantity, which is due to the fact $\{ H,H\} = 0$. |
604 |
When designing any numerical methods, one should always try to |
605 |
preserve the structural properties of the original ODE and its |
606 |
propagator. |
607 |
|
608 |
\subsection{\label{introSection:constructionSymplectic}Construction of Symplectic Methods} |
609 |
A lot of well established and very effective numerical methods have |
610 |
been successful precisely because of their symplectic nature even |
611 |
though this fact was not recognized when they were first |
612 |
constructed. The most famous example is the Verlet-leapfrog method |
613 |
in molecular dynamics. In general, symplectic integrators can be |
614 |
constructed using one of four different methods. |
615 |
\begin{enumerate} |
616 |
\item Generating functions |
617 |
\item Variational methods |
618 |
\item Runge-Kutta methods |
619 |
\item Splitting methods |
620 |
\end{enumerate} |
621 |
Generating functions\cite{Channell1990} tend to lead to methods |
622 |
which are cumbersome and difficult to use. In dissipative systems, |
623 |
variational methods can capture the decay of energy |
624 |
accurately.\cite{Kane2000} Since they are geometrically unstable |
625 |
against non-Hamiltonian perturbations, ordinary implicit Runge-Kutta |
626 |
methods are not suitable for Hamiltonian |
627 |
system.\cite{Cartwright1992} Recently, various high-order explicit |
628 |
Runge-Kutta methods \cite{Owren1992,Chen2003} have been developed to |
629 |
overcome this instability. However, due to computational penalty |
630 |
involved in implementing the Runge-Kutta methods, they have not |
631 |
attracted much attention from the Molecular Dynamics community. |
632 |
Instead, splitting methods have been widely accepted since they |
633 |
exploit natural decompositions of the system.\cite{McLachlan1998, |
634 |
Tuckerman1992} |
635 |
|
636 |
\subsubsection{\label{introSection:splittingMethod}\textbf{Splitting Methods}} |
637 |
|
638 |
The main idea behind splitting methods is to decompose the discrete |
639 |
$\varphi_h$ as a composition of simpler propagators, |
640 |
\begin{equation} |
641 |
\varphi _h = \varphi _{h_1 } \circ \varphi _{h_2 } \ldots \circ |
642 |
\varphi _{h_n } |
643 |
\label{introEquation:FlowDecomposition} |
644 |
\end{equation} |
645 |
where each of the sub-propagator is chosen such that each represent |
646 |
a simpler integration of the system. Suppose that a Hamiltonian |
647 |
system takes the form, |
648 |
\[ |
649 |
H = H_1 + H_2. |
650 |
\] |
651 |
Here, $H_1$ and $H_2$ may represent different physical processes of |
652 |
the system. For instance, they may relate to kinetic and potential |
653 |
energy respectively, which is a natural decomposition of the |
654 |
problem. If $H_1$ and $H_2$ can be integrated using exact |
655 |
propagators $\varphi_1(t)$ and $\varphi_2(t)$, respectively, a |
656 |
simple first order expression is then given by the Lie-Trotter |
657 |
formula\cite{Trotter1959} |
658 |
\begin{equation} |
659 |
\varphi _h = \varphi _{1,h} \circ \varphi _{2,h}, |
660 |
\label{introEquation:firstOrderSplitting} |
661 |
\end{equation} |
662 |
where $\varphi _h$ is the result of applying the corresponding |
663 |
continuous $\varphi _i$ over a time $h$. By definition, as |
664 |
$\varphi_i(t)$ is the exact solution of a Hamiltonian system, it |
665 |
must follow that each operator $\varphi_i(t)$ is a symplectic map. |
666 |
It is easy to show that any composition of symplectic propagators |
667 |
yields a symplectic map, |
668 |
\begin{equation} |
669 |
(\varphi '\phi ')^T J\varphi '\phi ' = \phi '^T \varphi '^T J\varphi |
670 |
'\phi ' = \phi '^T J\phi ' = J, |
671 |
\label{introEquation:SymplecticFlowComposition} |
672 |
\end{equation} |
673 |
where $\phi$ and $\psi$ both are symplectic maps. Thus operator |
674 |
splitting in this context automatically generates a symplectic map. |
675 |
The Lie-Trotter |
676 |
splitting(Eq.~\ref{introEquation:firstOrderSplitting}) introduces |
677 |
local errors proportional to $h^2$, while the Strang splitting gives |
678 |
a second-order decomposition,\cite{Strang1968} |
679 |
\begin{equation} |
680 |
\varphi _h = \varphi _{1,h/2} \circ \varphi _{2,h} \circ \varphi |
681 |
_{1,h/2} , \label{introEquation:secondOrderSplitting} |
682 |
\end{equation} |
683 |
which has a local error proportional to $h^3$. The Strang |
684 |
splitting's popularity in molecular simulation community attribute |
685 |
to its symmetric property, |
686 |
\begin{equation} |
687 |
\varphi _h^{ - 1} = \varphi _{ - h}. |
688 |
\label{introEquation:timeReversible} |
689 |
\end{equation} |
690 |
|
691 |
\subsubsection{\label{introSection:exampleSplittingMethod}\textbf{Examples of the Splitting Method}} |
692 |
The classical equation for a system consisting of interacting |
693 |
particles can be written in Hamiltonian form, |
694 |
\[ |
695 |
H = T + V |
696 |
\] |
697 |
where $T$ is the kinetic energy and $V$ is the potential energy. |
698 |
Setting $H_1 = T, H_2 = V$ and applying the Strang splitting, one |
699 |
obtains the following: |
700 |
\begin{align} |
701 |
q(\Delta t) &= q(0) + \dot{q}(0)\Delta t + |
702 |
\frac{F[q(0)]}{m}\frac{\Delta t^2}{2}, % |
703 |
\label{introEquation:Lp10a} \\% |
704 |
% |
705 |
\dot{q}(\Delta t) &= \dot{q}(0) + \frac{\Delta t}{2m} |
706 |
\biggl [F[q(0)] + F[q(\Delta t)] \biggr]. % |
707 |
\label{introEquation:Lp10b} |
708 |
\end{align} |
709 |
where $F(t)$ is the force at time $t$. This integration scheme is |
710 |
known as \emph{velocity verlet} which is |
711 |
symplectic(Eq.~\ref{introEquation:SymplecticFlowComposition}), |
712 |
time-reversible(Eq.~\ref{introEquation:timeReversible}) and |
713 |
volume-preserving (Eq.~\ref{introEquation:volumePreserving}). These |
714 |
geometric properties attribute to its long-time stability and its |
715 |
popularity in the community. However, the most commonly used |
716 |
velocity verlet integration scheme is written as below, |
717 |
\begin{align} |
718 |
\dot{q}\biggl (\frac{\Delta t}{2}\biggr ) &= |
719 |
\dot{q}(0) + \frac{\Delta t}{2m}\, F[q(0)], \label{introEquation:Lp9a}\\% |
720 |
% |
721 |
q(\Delta t) &= q(0) + \Delta t\, \dot{q}\biggl (\frac{\Delta t}{2}\biggr ),% |
722 |
\label{introEquation:Lp9b}\\% |
723 |
% |
724 |
\dot{q}(\Delta t) &= \dot{q}\biggl (\frac{\Delta t}{2}\biggr ) + |
725 |
\frac{\Delta t}{2m}\, F[q(t)]. \label{introEquation:Lp9c} |
726 |
\end{align} |
727 |
From the preceding splitting, one can see that the integration of |
728 |
the equations of motion would follow: |
729 |
\begin{enumerate} |
730 |
\item calculate the velocities at the half step, $\frac{\Delta t}{2}$, from the forces calculated at the initial position. |
731 |
|
732 |
\item Use the half step velocities to move positions one whole step, $\Delta t$. |
733 |
|
734 |
\item Evaluate the forces at the new positions, $q(\Delta t)$, and use the new forces to complete the velocity move. |
735 |
|
736 |
\item Repeat from step 1 with the new position, velocities, and forces assuming the roles of the initial values. |
737 |
\end{enumerate} |
738 |
By simply switching the order of the propagators in the splitting |
739 |
and composing a new integrator, the \emph{position verlet} |
740 |
integrator, can be generated, |
741 |
\begin{align} |
742 |
\dot q(\Delta t) &= \dot q(0) + \Delta tF(q(0))\left[ {q(0) + |
743 |
\frac{{\Delta t}}{{2m}}\dot q(0)} \right], % |
744 |
\label{introEquation:positionVerlet1} \\% |
745 |
% |
746 |
q(\Delta t) &= q(0) + \frac{{\Delta t}}{2}\left[ {\dot q(0) + \dot |
747 |
q(\Delta t)} \right]. % |
748 |
\label{introEquation:positionVerlet2} |
749 |
\end{align} |
750 |
|
751 |
\subsubsection{\label{introSection:errorAnalysis}\textbf{Error Analysis and Higher Order Methods}} |
752 |
|
753 |
The Baker-Campbell-Hausdorff formula\cite{Gilmore1974} can be used |
754 |
to determine the local error of a splitting method in terms of the |
755 |
commutator of the operators associated with the sub-propagator. For |
756 |
operators $hX$ and $hY$ which are associated with $\varphi_1(t)$ and |
757 |
$\varphi_2(t)$ respectively , we have |
758 |
\begin{equation} |
759 |
\exp (hX + hY) = \exp (hZ) |
760 |
\end{equation} |
761 |
where |
762 |
\begin{equation} |
763 |
hZ = hX + hY + \frac{{h^2 }}{2}[X,Y] + \frac{{h^3 }}{2}\left( |
764 |
{[X,[X,Y]] + [Y,[Y,X]]} \right) + \ldots . |
765 |
\end{equation} |
766 |
Here, $[X,Y]$ is the commutator of operator $X$ and $Y$ given by |
767 |
\[ |
768 |
[X,Y] = XY - YX . |
769 |
\] |
770 |
Applying the Baker-Campbell-Hausdorff formula\cite{Varadarajan1974} |
771 |
to the Strang splitting, we can obtain |
772 |
\begin{eqnarray*} |
773 |
\exp (h X/2)\exp (h Y)\exp (h X/2) & = & \exp (h X + h Y + h^2 [X,Y]/4 + h^2 [Y,X]/4 \\ |
774 |
& & \mbox{} + h^2 [X,X]/8 + h^2 [Y,Y]/8 \\ |
775 |
& & \mbox{} + h^3 [Y,[Y,X]]/12 - h^3[X,[X,Y]]/24 + \ldots |
776 |
). |
777 |
\end{eqnarray*} |
778 |
Since $ [X,Y] + [Y,X] = 0$ and $ [X,X] = 0$, the dominant local |
779 |
error of Strang splitting is proportional to $h^3$. The same |
780 |
procedure can be applied to a general splitting of the form |
781 |
\begin{equation} |
782 |
\varphi _{b_m h}^2 \circ \varphi _{a_m h}^1 \circ \varphi _{b_{m - |
783 |
1} h}^2 \circ \ldots \circ \varphi _{a_1 h}^1 . |
784 |
\end{equation} |
785 |
A careful choice of coefficient $a_1 \ldots b_m$ will lead to higher |
786 |
order methods. Yoshida proposed an elegant way to compose higher |
787 |
order methods based on symmetric splitting.\cite{Yoshida1990} Given |
788 |
a symmetric second order base method $ \varphi _h^{(2)} $, a |
789 |
fourth-order symmetric method can be constructed by composing, |
790 |
\[ |
791 |
\varphi _h^{(4)} = \varphi _{\alpha h}^{(2)} \circ \varphi _{\beta |
792 |
h}^{(2)} \circ \varphi _{\alpha h}^{(2)} |
793 |
\] |
794 |
where $ \alpha = - \frac{{2^{1/3} }}{{2 - 2^{1/3} }}$ and $ \beta |
795 |
= \frac{{2^{1/3} }}{{2 - 2^{1/3} }}$. Moreover, a symmetric |
796 |
integrator $ \varphi _h^{(2n + 2)}$ can be composed by |
797 |
\begin{equation} |
798 |
\varphi _h^{(2n + 2)} = \varphi _{\alpha h}^{(2n)} \circ \varphi |
799 |
_{\beta h}^{(2n)} \circ \varphi _{\alpha h}^{(2n)}, |
800 |
\end{equation} |
801 |
if the weights are chosen as |
802 |
\[ |
803 |
\alpha = - \frac{{2^{1/(2n + 1)} }}{{2 - 2^{1/(2n + 1)} }},\beta = |
804 |
\frac{{2^{1/(2n + 1)} }}{{2 - 2^{1/(2n + 1)} }} . |
805 |
\] |
806 |
|
807 |
\section{\label{introSection:molecularDynamics}Molecular Dynamics} |
808 |
|
809 |
As one of the principal tools of molecular modeling, Molecular |
810 |
dynamics has proven to be a powerful tool for studying the functions |
811 |
of biological systems, providing structural, thermodynamic and |
812 |
dynamical information. The basic idea of molecular dynamics is that |
813 |
macroscopic properties are related to microscopic behavior and |
814 |
microscopic behavior can be calculated from the trajectories in |
815 |
simulations. For instance, instantaneous temperature of a |
816 |
Hamiltonian system of $N$ particles can be measured by |
817 |
\[ |
818 |
T = \sum\limits_{i = 1}^N {\frac{{m_i v_i^2 }}{{fk_B }}} |
819 |
\] |
820 |
where $m_i$ and $v_i$ are the mass and velocity of $i$th particle |
821 |
respectively, $f$ is the number of degrees of freedom, and $k_B$ is |
822 |
the Boltzman constant. |
823 |
|
824 |
A typical molecular dynamics run consists of three essential steps: |
825 |
\begin{enumerate} |
826 |
\item Initialization |
827 |
\begin{enumerate} |
828 |
\item Preliminary preparation |
829 |
\item Minimization |
830 |
\item Heating |
831 |
\item Equilibration |
832 |
\end{enumerate} |
833 |
\item Production |
834 |
\item Analysis |
835 |
\end{enumerate} |
836 |
These three individual steps will be covered in the following |
837 |
sections. Sec.~\ref{introSec:initialSystemSettings} deals with the |
838 |
initialization of a simulation. Sec.~\ref{introSection:production} |
839 |
discusses issues of production runs. |
840 |
Sec.~\ref{introSection:Analysis} provides the theoretical tools for |
841 |
analysis of trajectories. |
842 |
|
843 |
\subsection{\label{introSec:initialSystemSettings}Initialization} |
844 |
|
845 |
\subsubsection{\textbf{Preliminary preparation}} |
846 |
|
847 |
When selecting the starting structure of a molecule for molecular |
848 |
simulation, one may retrieve its Cartesian coordinates from public |
849 |
databases, such as RCSB Protein Data Bank \textit{etc}. Although |
850 |
thousands of crystal structures of molecules are discovered every |
851 |
year, many more remain unknown due to the difficulties of |
852 |
purification and crystallization. Even for molecules with known |
853 |
structures, some important information is missing. For example, a |
854 |
missing hydrogen atom which acts as donor in hydrogen bonding must |
855 |
be added. Moreover, in order to include electrostatic interactions, |
856 |
one may need to specify the partial charges for individual atoms. |
857 |
Under some circumstances, we may even need to prepare the system in |
858 |
a special configuration. For instance, when studying transport |
859 |
phenomenon in membrane systems, we may prepare the lipids in a |
860 |
bilayer structure instead of placing lipids randomly in solvent, |
861 |
since we are not interested in the slow self-aggregation process. |
862 |
|
863 |
\subsubsection{\textbf{Minimization}} |
864 |
|
865 |
It is quite possible that some of molecules in the system from |
866 |
preliminary preparation may be overlapping with each other. This |
867 |
close proximity leads to high initial potential energy which |
868 |
consequently jeopardizes any molecular dynamics simulations. To |
869 |
remove these steric overlaps, one typically performs energy |
870 |
minimization to find a more reasonable conformation. Several energy |
871 |
minimization methods have been developed to exploit the energy |
872 |
surface and to locate the local minimum. While converging slowly |
873 |
near the minimum, the steepest descent method is extremely robust when |
874 |
systems are strongly anharmonic. Thus, it is often used to refine |
875 |
structures from crystallographic data. Relying on the Hessian, |
876 |
advanced methods like Newton-Raphson converge rapidly to a local |
877 |
minimum, but become unstable if the energy surface is far from |
878 |
quadratic. Another factor that must be taken into account, when |
879 |
choosing energy minimization method, is the size of the system. |
880 |
Steepest descent and conjugate gradient can deal with models of any |
881 |
size. Because of the limits on computer memory to store the hessian |
882 |
matrix and the computing power needed to diagonalize these matrices, |
883 |
most Newton-Raphson methods can not be used with very large systems. |
884 |
|
885 |
\subsubsection{\textbf{Heating}} |
886 |
|
887 |
Typically, heating is performed by assigning random velocities |
888 |
according to a Maxwell-Boltzman distribution for a desired |
889 |
temperature. Beginning at a lower temperature and gradually |
890 |
increasing the temperature by assigning larger random velocities, we |
891 |
end up setting the temperature of the system to a final temperature |
892 |
at which the simulation will be conducted. In the heating phase, we |
893 |
should also keep the system from drifting or rotating as a whole. To |
894 |
do this, the net linear momentum and angular momentum of the system |
895 |
is shifted to zero after each resampling from the Maxwell -Boltzman |
896 |
distribution. |
897 |
|
898 |
\subsubsection{\textbf{Equilibration}} |
899 |
|
900 |
The purpose of equilibration is to allow the system to evolve |
901 |
spontaneously for a period of time and reach equilibrium. The |
902 |
procedure is continued until various statistical properties, such as |
903 |
temperature, pressure, energy, volume and other structural |
904 |
properties \textit{etc}, become independent of time. Strictly |
905 |
speaking, minimization and heating are not necessary, provided the |
906 |
equilibration process is long enough. However, these steps can serve |
907 |
as a mean to arrive at an equilibrated structure in an effective |
908 |
way. |
909 |
|
910 |
\subsection{\label{introSection:production}Production} |
911 |
|
912 |
The production run is the most important step of the simulation, in |
913 |
which the equilibrated structure is used as a starting point and the |
914 |
motions of the molecules are collected for later analysis. In order |
915 |
to capture the macroscopic properties of the system, the molecular |
916 |
dynamics simulation must be performed by sampling correctly and |
917 |
efficiently from the relevant thermodynamic ensemble. |
918 |
|
919 |
The most expensive part of a molecular dynamics simulation is the |
920 |
calculation of non-bonded forces, such as van der Waals force and |
921 |
Coulombic forces \textit{etc}. For a system of $N$ particles, the |
922 |
complexity of the algorithm for pair-wise interactions is $O(N^2 )$, |
923 |
which makes large simulations prohibitive in the absence of any |
924 |
algorithmic tricks. A natural approach to avoid system size issues |
925 |
is to represent the bulk behavior by a finite number of the |
926 |
particles. However, this approach will suffer from surface effects |
927 |
at the edges of the simulation. To offset this, \textit{Periodic |
928 |
boundary conditions} (see Fig.~\ref{introFig:pbc}) were developed to |
929 |
simulate bulk properties with a relatively small number of |
930 |
particles. In this method, the simulation box is replicated |
931 |
throughout space to form an infinite lattice. During the simulation, |
932 |
when a particle moves in the primary cell, its image in other cells |
933 |
move in exactly the same direction with exactly the same |
934 |
orientation. Thus, as a particle leaves the primary cell, one of its |
935 |
images will enter through the opposite face. |
936 |
\begin{figure} |
937 |
\centering |
938 |
\includegraphics[width=\linewidth]{pbc.eps} |
939 |
\caption[An illustration of periodic boundary conditions]{A 2-D |
940 |
illustration of periodic boundary conditions. As one particle leaves |
941 |
the left of the simulation box, an image of it enters the right.} |
942 |
\label{introFig:pbc} |
943 |
\end{figure} |
944 |
|
945 |
%cutoff and minimum image convention |
946 |
Another important technique to improve the efficiency of force |
947 |
evaluation is to apply spherical cutoffs where particles farther |
948 |
than a predetermined distance are not included in the |
949 |
calculation.\cite{Frenkel1996} The use of a cutoff radius will cause |
950 |
a discontinuity in the potential energy curve. Fortunately, one can |
951 |
shift a simple radial potential to ensure the potential curve go |
952 |
smoothly to zero at the cutoff radius. The cutoff strategy works |
953 |
well for Lennard-Jones interaction because of its short range |
954 |
nature. However, simply truncating the electrostatic interaction |
955 |
with the use of cutoffs has been shown to lead to severe artifacts |
956 |
in simulations. The Ewald summation, in which the slowly decaying |
957 |
Coulomb potential is transformed into direct and reciprocal sums |
958 |
with rapid and absolute convergence, has proved to minimize the |
959 |
periodicity artifacts in liquid simulations. Taking advantage of |
960 |
fast Fourier transform (FFT) techniques for calculating discrete |
961 |
Fourier transforms, the particle mesh-based |
962 |
methods\cite{Hockney1981,Shimada1993, Luty1994} are accelerated from |
963 |
$O(N^{3/2})$ to $O(N logN)$. An alternative approach is the |
964 |
\emph{fast multipole method}\cite{Greengard1987, Greengard1994}, |
965 |
which treats Coulombic interactions exactly at short range, and |
966 |
approximate the potential at long range through multipolar |
967 |
expansion. In spite of their wide acceptance at the molecular |
968 |
simulation community, these two methods are difficult to implement |
969 |
correctly and efficiently. Instead, we use a damped and |
970 |
charge-neutralized Coulomb potential method developed by Wolf and |
971 |
his coworkers.\cite{Wolf1999} The shifted Coulomb potential for |
972 |
particle $i$ and particle $j$ at distance $r_{rj}$ is given by: |
973 |
\begin{equation} |
974 |
V(r_{ij})= \frac{q_i q_j \textrm{erfc}(\alpha |
975 |
r_{ij})}{r_{ij}}-\lim_{r_{ij}\rightarrow |
976 |
R_\textrm{c}}\left\{\frac{q_iq_j \textrm{erfc}(\alpha |
977 |
r_{ij})}{r_{ij}}\right\}, \label{introEquation:shiftedCoulomb} |
978 |
\end{equation} |
979 |
where $\alpha$ is the convergence parameter. Due to the lack of |
980 |
inherent periodicity and rapid convergence,this method is extremely |
981 |
efficient and easy to implement. |
982 |
\begin{figure} |
983 |
\centering |
984 |
\includegraphics[width=\linewidth]{shifted_coulomb.eps} |
985 |
\caption[An illustration of shifted Coulomb potential]{An |
986 |
illustration of shifted Coulomb potential.} |
987 |
\label{introFigure:shiftedCoulomb} |
988 |
\end{figure} |
989 |
|
990 |
%multiple time step |
991 |
|
992 |
\subsection{\label{introSection:Analysis} Analysis} |
993 |
|
994 |
Recently, advanced visualization techniques have been applied to |
995 |
monitor the motions of molecules. Although the dynamics of the |
996 |
system can be described qualitatively from animation, quantitative |
997 |
trajectory analysis is more useful. According to the principles of |
998 |
Statistical Mechanics in |
999 |
Sec.~\ref{introSection:statisticalMechanics}, one can compute |
1000 |
thermodynamic properties, analyze fluctuations of structural |
1001 |
parameters, and investigate time-dependent processes of the molecule |
1002 |
from the trajectories. |
1003 |
|
1004 |
\subsubsection{\label{introSection:thermodynamicsProperties}\textbf{Thermodynamic Properties}} |
1005 |
|
1006 |
Thermodynamic properties, which can be expressed in terms of some |
1007 |
function of the coordinates and momenta of all particles in the |
1008 |
system, can be directly computed from molecular dynamics. The usual |
1009 |
way to measure the pressure is based on virial theorem of Clausius |
1010 |
which states that the virial is equal to $-3Nk_BT$. For a system |
1011 |
with forces between particles, the total virial, $W$, contains the |
1012 |
contribution from external pressure and interaction between the |
1013 |
particles: |
1014 |
\[ |
1015 |
W = - 3PV + \left\langle {\sum\limits_{i < j} {r{}_{ij} \cdot |
1016 |
f_{ij} } } \right\rangle |
1017 |
\] |
1018 |
where $f_{ij}$ is the force between particle $i$ and $j$ at a |
1019 |
distance $r_{ij}$. Thus, the expression for the pressure is given |
1020 |
by: |
1021 |
\begin{equation} |
1022 |
P = \frac{{Nk_B T}}{V} - \frac{1}{{3V}}\left\langle {\sum\limits_{i |
1023 |
< j} {r{}_{ij} \cdot f_{ij} } } \right\rangle |
1024 |
\end{equation} |
1025 |
|
1026 |
\subsubsection{\label{introSection:structuralProperties}\textbf{Structural Properties}} |
1027 |
|
1028 |
Structural Properties of a simple fluid can be described by a set of |
1029 |
distribution functions. Among these functions,the \emph{pair |
1030 |
distribution function}, also known as \emph{radial distribution |
1031 |
function}, is of most fundamental importance to liquid theory. |
1032 |
Experimentally, pair distribution functions can be gathered by |
1033 |
Fourier transforming raw data from a series of neutron diffraction |
1034 |
experiments and integrating over the surface |
1035 |
factor.\cite{Powles1973} The experimental results can serve as a |
1036 |
criterion to justify the correctness of a liquid model. Moreover, |
1037 |
various equilibrium thermodynamic and structural properties can also |
1038 |
be expressed in terms of the radial distribution |
1039 |
function.\cite{Allen1987} The pair distribution functions $g(r)$ |
1040 |
gives the probability that a particle $i$ will be located at a |
1041 |
distance $r$ from a another particle $j$ in the system |
1042 |
\begin{equation} |
1043 |
g(r) = \frac{V}{{N^2 }}\left\langle {\sum\limits_i {\sum\limits_{j |
1044 |
\ne i} {\delta (r - r_{ij} )} } } \right\rangle = \frac{\rho |
1045 |
(r)}{\rho}. |
1046 |
\end{equation} |
1047 |
Note that the delta function can be replaced by a histogram in |
1048 |
computer simulation. Peaks in $g(r)$ represent solvent shells, and |
1049 |
the height of these peaks gradually decreases to 1 as the liquid of |
1050 |
large distance approaches the bulk density. |
1051 |
|
1052 |
|
1053 |
\subsubsection{\label{introSection:timeDependentProperties}\textbf{Time-dependent |
1054 |
Properties}} |
1055 |
|
1056 |
Time-dependent properties are usually calculated using \emph{time |
1057 |
correlation functions}, which correlate random variables $A$ and $B$ |
1058 |
at two different times, |
1059 |
\begin{equation} |
1060 |
C_{AB} (t) = \left\langle {A(t)B(0)} \right\rangle. |
1061 |
\label{introEquation:timeCorrelationFunction} |
1062 |
\end{equation} |
1063 |
If $A$ and $B$ refer to same variable, this kind of correlation |
1064 |
functions are called \emph{autocorrelation functions}. One typical example is the velocity autocorrelation |
1065 |
function which is directly related to transport properties of |
1066 |
molecular liquids: |
1067 |
\begin{equation} |
1068 |
D = \frac{1}{3}\int\limits_0^\infty {\left\langle {v(t) \cdot v(0)} |
1069 |
\right\rangle } dt |
1070 |
\end{equation} |
1071 |
where $D$ is diffusion constant. Unlike the velocity autocorrelation |
1072 |
function, which is averaged over time origins and over all the |
1073 |
atoms, the dipole autocorrelation functions is calculated for the |
1074 |
entire system. The dipole autocorrelation function is given by: |
1075 |
\begin{equation} |
1076 |
c_{dipole} = \left\langle {u_{tot} (t) \cdot u_{tot} (t)} |
1077 |
\right\rangle |
1078 |
\end{equation} |
1079 |
Here $u_{tot}$ is the net dipole of the entire system and is given |
1080 |
by |
1081 |
\begin{equation} |
1082 |
u_{tot} (t) = \sum\limits_i {u_i (t)}. |
1083 |
\end{equation} |
1084 |
In principle, many time correlation functions can be related to |
1085 |
Fourier transforms of the infrared, Raman, and inelastic neutron |
1086 |
scattering spectra of molecular liquids. In practice, one can |
1087 |
extract the IR spectrum from the intensity of the molecular dipole |
1088 |
fluctuation at each frequency using the following relationship: |
1089 |
\begin{equation} |
1090 |
\hat c_{dipole} (v) = \int_{ - \infty }^\infty {c_{dipole} (t)e^{ - |
1091 |
i2\pi vt} dt}. |
1092 |
\end{equation} |
1093 |
|
1094 |
\section{\label{introSection:rigidBody}Dynamics of Rigid Bodies} |
1095 |
|
1096 |
Rigid bodies are frequently involved in the modeling of different |
1097 |
areas, including engineering, physics and chemistry. For example, |
1098 |
missiles and vehicles are usually modeled by rigid bodies. The |
1099 |
movement of the objects in 3D gaming engines or other physics |
1100 |
simulators is governed by rigid body dynamics. In molecular |
1101 |
simulations, rigid bodies are used to simplify protein-protein |
1102 |
docking studies.\cite{Gray2003} |
1103 |
|
1104 |
It is very important to develop stable and efficient methods to |
1105 |
integrate the equations of motion for orientational degrees of |
1106 |
freedom. Euler angles are the natural choice to describe the |
1107 |
rotational degrees of freedom. However, due to $\frac {1}{sin |
1108 |
\theta}$ singularities, the numerical integration of corresponding |
1109 |
equations of these motion is very inefficient and inaccurate. |
1110 |
Although an alternative integrator using multiple sets of Euler |
1111 |
angles can overcome this difficulty\cite{Barojas1973}, the |
1112 |
computational penalty and the loss of angular momentum conservation |
1113 |
still remain. A singularity-free representation utilizing |
1114 |
quaternions was developed by Evans in 1977.\cite{Evans1977} |
1115 |
Unfortunately, this approach used a nonseparable Hamiltonian |
1116 |
resulting from the quaternion representation, which prevented the |
1117 |
symplectic algorithm from being utilized. Another different approach |
1118 |
is to apply holonomic constraints to the atoms belonging to the |
1119 |
rigid body. Each atom moves independently under the normal forces |
1120 |
deriving from potential energy and constraint forces which are used |
1121 |
to guarantee the rigidness. However, due to their iterative nature, |
1122 |
the SHAKE and Rattle algorithms also converge very slowly when the |
1123 |
number of constraints increases.\cite{Ryckaert1977, Andersen1983} |
1124 |
|
1125 |
A break-through in geometric literature suggests that, in order to |
1126 |
develop a long-term integration scheme, one should preserve the |
1127 |
symplectic structure of the propagator. By introducing a conjugate |
1128 |
momentum to the rotation matrix $Q$ and re-formulating Hamiltonian's |
1129 |
equation, a symplectic integrator, RSHAKE\cite{Kol1997}, was |
1130 |
proposed to evolve the Hamiltonian system in a constraint manifold |
1131 |
by iteratively satisfying the orthogonality constraint $Q^T Q = 1$. |
1132 |
An alternative method using the quaternion representation was |
1133 |
developed by Omelyan.\cite{Omelyan1998} However, both of these |
1134 |
methods are iterative and inefficient. In this section, we descibe a |
1135 |
symplectic Lie-Poisson integrator for rigid bodies developed by |
1136 |
Dullweber and his coworkers\cite{Dullweber1997} in depth. |
1137 |
|
1138 |
\subsection{\label{introSection:constrainedHamiltonianRB}Constrained Hamiltonian for Rigid Bodies} |
1139 |
The Hamiltonian of a rigid body is given by |
1140 |
\begin{equation} |
1141 |
H = \frac{1}{2}(p^T m^{ - 1} p) + \frac{1}{2}tr(PJ^{ - 1} P) + |
1142 |
V(q,Q) + \frac{1}{2}tr[(QQ^T - 1)\Lambda ]. |
1143 |
\label{introEquation:RBHamiltonian} |
1144 |
\end{equation} |
1145 |
Here, $q$ and $Q$ are the position vector and rotation matrix for |
1146 |
the rigid-body, $p$ and $P$ are conjugate momenta to $q$ and $Q$ , |
1147 |
and $J$, a diagonal matrix, is defined by |
1148 |
\[ |
1149 |
I_{ii}^{ - 1} = \frac{1}{2}\sum\limits_{i \ne j} {J_{jj}^{ - 1} } |
1150 |
\] |
1151 |
where $I_{ii}$ is the diagonal element of the inertia tensor. This |
1152 |
constrained Hamiltonian equation is subjected to a holonomic |
1153 |
constraint, |
1154 |
\begin{equation} |
1155 |
Q^T Q = 1, \label{introEquation:orthogonalConstraint} |
1156 |
\end{equation} |
1157 |
which is used to ensure the rotation matrix's unitarity. Using |
1158 |
Eq.~\ref{introEquation:motionHamiltonianCoordinate} and Eq.~ |
1159 |
\ref{introEquation:motionHamiltonianMomentum}, one can write down |
1160 |
the equations of motion, |
1161 |
\begin{eqnarray} |
1162 |
\frac{{dq}}{{dt}} & = & \frac{p}{m}, \label{introEquation:RBMotionPosition}\\ |
1163 |
\frac{{dp}}{{dt}} & = & - \nabla _q V(q,Q), \label{introEquation:RBMotionMomentum}\\ |
1164 |
\frac{{dQ}}{{dt}} & = & PJ^{ - 1}, \label{introEquation:RBMotionRotation}\\ |
1165 |
\frac{{dP}}{{dt}} & = & - \nabla _Q V(q,Q) - 2Q\Lambda . \label{introEquation:RBMotionP} |
1166 |
\end{eqnarray} |
1167 |
Differentiating Eq.~\ref{introEquation:orthogonalConstraint} and |
1168 |
using Eq.~\ref{introEquation:RBMotionMomentum}, one may obtain, |
1169 |
\begin{equation} |
1170 |
Q^T PJ^{ - 1} + J^{ - 1} P^T Q = 0 . \\ |
1171 |
\label{introEquation:RBFirstOrderConstraint} |
1172 |
\end{equation} |
1173 |
In general, there are two ways to satisfy the holonomic constraints. |
1174 |
We can use a constraint force provided by a Lagrange multiplier on |
1175 |
the normal manifold to keep the motion on the constraint space. Or |
1176 |
we can simply evolve the system on the constraint manifold. These |
1177 |
two methods have been proved to be equivalent. The holonomic |
1178 |
constraint and equations of motions define a constraint manifold for |
1179 |
rigid bodies |
1180 |
\[ |
1181 |
M = \left\{ {(Q,P):Q^T Q = 1,Q^T PJ^{ - 1} + J^{ - 1} P^T Q = 0} |
1182 |
\right\}. |
1183 |
\] |
1184 |
Unfortunately, this constraint manifold is not $T^* SO(3)$ which is |
1185 |
a symplectic manifold on Lie rotation group $SO(3)$. However, it |
1186 |
turns out that under symplectic transformation, the cotangent space |
1187 |
and the phase space are diffeomorphic. By introducing |
1188 |
\[ |
1189 |
\tilde Q = Q,\tilde P = \frac{1}{2}\left( {P - QP^T Q} \right), |
1190 |
\] |
1191 |
the mechanical system subjected to a holonomic constraint manifold $M$ |
1192 |
can be re-formulated as a Hamiltonian system on the cotangent space |
1193 |
\[ |
1194 |
T^* SO(3) = \left\{ {(\tilde Q,\tilde P):\tilde Q^T \tilde Q = |
1195 |
1,\tilde Q^T \tilde PJ^{ - 1} + J^{ - 1} P^T \tilde Q = 0} \right\} |
1196 |
\] |
1197 |
For a body fixed vector $X_i$ with respect to the center of mass of |
1198 |
the rigid body, its corresponding lab fixed vector $X_0^{lab}$ is |
1199 |
given as |
1200 |
\begin{equation} |
1201 |
X_i^{lab} = Q X_i + q. |
1202 |
\end{equation} |
1203 |
Therefore, potential energy $V(q,Q)$ is defined by |
1204 |
\[ |
1205 |
V(q,Q) = V(Q X_0 + q). |
1206 |
\] |
1207 |
Hence, the force and torque are given by |
1208 |
\[ |
1209 |
\nabla _q V(q,Q) = F(q,Q) = \sum\limits_i {F_i (q,Q)}, |
1210 |
\] |
1211 |
and |
1212 |
\[ |
1213 |
\nabla _Q V(q,Q) = F(q,Q)X_i^t |
1214 |
\] |
1215 |
respectively. As a common choice to describe the rotation dynamics |
1216 |
of the rigid body, the angular momentum on the body fixed frame $\Pi |
1217 |
= Q^t P$ is introduced to rewrite the equations of motion, |
1218 |
\begin{equation} |
1219 |
\begin{array}{l} |
1220 |
\dot \Pi = J^{ - 1} \Pi ^T \Pi + Q^T \sum\limits_i {F_i (q,Q)X_i^T } - \Lambda, \\ |
1221 |
\dot Q = Q\Pi {\rm{ }}J^{ - 1}, \\ |
1222 |
\end{array} |
1223 |
\label{introEqaution:RBMotionPI} |
1224 |
\end{equation} |
1225 |
as well as holonomic constraints $\Pi J^{ - 1} + J^{ - 1} \Pi ^t = |
1226 |
0$ and $Q^T Q = 1$. For a vector $v(v_1 ,v_2 ,v_3 ) \in R^3$ and a |
1227 |
matrix $\hat v \in so(3)^ \star$, the hat-map isomorphism, |
1228 |
\begin{equation} |
1229 |
v(v_1 ,v_2 ,v_3 ) \Leftrightarrow \hat v = \left( |
1230 |
{\begin{array}{*{20}c} |
1231 |
0 & { - v_3 } & {v_2 } \\ |
1232 |
{v_3 } & 0 & { - v_1 } \\ |
1233 |
{ - v_2 } & {v_1 } & 0 \\ |
1234 |
\end{array}} \right), |
1235 |
\label{introEquation:hatmapIsomorphism} |
1236 |
\end{equation} |
1237 |
will let us associate the matrix products with traditional vector |
1238 |
operations |
1239 |
\[ |
1240 |
\hat vu = v \times u. |
1241 |
\] |
1242 |
Using Eq.~\ref{introEqaution:RBMotionPI}, one can construct a skew |
1243 |
matrix, |
1244 |
\begin{eqnarray} |
1245 |
(\dot \Pi - \dot \Pi ^T )&= &(\Pi - \Pi ^T )(J^{ - 1} \Pi + \Pi J^{ - 1} ) \notag \\ |
1246 |
& & + \sum\limits_i {[Q^T F_i (r,Q)X_i^T - X_i F_i (r,Q)^T Q]} - |
1247 |
(\Lambda - \Lambda ^T ). \label{introEquation:skewMatrixPI} |
1248 |
\end{eqnarray} |
1249 |
Since $\Lambda$ is symmetric, the last term of |
1250 |
Eq.~\ref{introEquation:skewMatrixPI} is zero, which implies the |
1251 |
Lagrange multiplier $\Lambda$ is absent from the equations of |
1252 |
motion. This unique property eliminates the requirement of |
1253 |
iterations which can not be avoided in other methods.\cite{Kol1997, |
1254 |
Omelyan1998} Applying the hat-map isomorphism, we obtain the |
1255 |
equation of motion for angular momentum in the body frame |
1256 |
\begin{equation} |
1257 |
\dot \pi = \pi \times I^{ - 1} \pi + \sum\limits_i {\left( {Q^T |
1258 |
F_i (r,Q)} \right) \times X_i }. |
1259 |
\label{introEquation:bodyAngularMotion} |
1260 |
\end{equation} |
1261 |
In the same manner, the equation of motion for rotation matrix is |
1262 |
given by |
1263 |
\[ |
1264 |
\dot Q = Qskew(I^{ - 1} \pi ). |
1265 |
\] |
1266 |
|
1267 |
\subsection{\label{introSection:SymplecticFreeRB}Symplectic |
1268 |
Lie-Poisson Integrator for Free Rigid Bodies} |
1269 |
|
1270 |
If there are no external forces exerted on the rigid body, the only |
1271 |
contribution to the rotational motion is from the kinetic energy |
1272 |
(the first term of \ref{introEquation:bodyAngularMotion}). The free |
1273 |
rigid body is an example of a Lie-Poisson system with Hamiltonian |
1274 |
function |
1275 |
\begin{equation} |
1276 |
T^r (\pi ) = T_1 ^r (\pi _1 ) + T_2^r (\pi _2 ) + T_3^r (\pi _3 ) |
1277 |
\label{introEquation:rotationalKineticRB} |
1278 |
\end{equation} |
1279 |
where $T_i^r (\pi _i ) = \frac{{\pi _i ^2 }}{{2I_i }}$ and |
1280 |
Lie-Poisson structure matrix, |
1281 |
\begin{equation} |
1282 |
J(\pi ) = \left( {\begin{array}{*{20}c} |
1283 |
0 & {\pi _3 } & { - \pi _2 } \\ |
1284 |
{ - \pi _3 } & 0 & {\pi _1 } \\ |
1285 |
{\pi _2 } & { - \pi _1 } & 0 \\ |
1286 |
\end{array}} \right). |
1287 |
\end{equation} |
1288 |
Thus, the dynamics of free rigid body is governed by |
1289 |
\begin{equation} |
1290 |
\frac{d}{{dt}}\pi = J(\pi )\nabla _\pi T^r (\pi ). |
1291 |
\end{equation} |
1292 |
One may notice that each $T_i^r$ in |
1293 |
Eq.~\ref{introEquation:rotationalKineticRB} can be solved exactly. |
1294 |
For instance, the equations of motion due to $T_1^r$ are given by |
1295 |
\begin{equation} |
1296 |
\frac{d}{{dt}}\pi = R_1 \pi ,\frac{d}{{dt}}Q = QR_1 |
1297 |
\label{introEqaution:RBMotionSingleTerm} |
1298 |
\end{equation} |
1299 |
with |
1300 |
\[ R_1 = \left( {\begin{array}{*{20}c} |
1301 |
0 & 0 & 0 \\ |
1302 |
0 & 0 & {\pi _1 } \\ |
1303 |
0 & { - \pi _1 } & 0 \\ |
1304 |
\end{array}} \right). |
1305 |
\] |
1306 |
The solutions of Eq.~\ref{introEqaution:RBMotionSingleTerm} is |
1307 |
\[ |
1308 |
\pi (\Delta t) = e^{\Delta tR_1 } \pi (0),Q(\Delta t) = |
1309 |
Q(0)e^{\Delta tR_1 } |
1310 |
\] |
1311 |
with |
1312 |
\[ |
1313 |
e^{\Delta tR_1 } = \left( {\begin{array}{*{20}c} |
1314 |
0 & 0 & 0 \\ |
1315 |
0 & {\cos \theta _1 } & {\sin \theta _1 } \\ |
1316 |
0 & { - \sin \theta _1 } & {\cos \theta _1 } \\ |
1317 |
\end{array}} \right),\theta _1 = \frac{{\pi _1 }}{{I_1 }}\Delta t. |
1318 |
\] |
1319 |
To reduce the cost of computing expensive functions in $e^{\Delta |
1320 |
tR_1 }$, we can use the Cayley transformation to obtain a |
1321 |
single-aixs propagator, |
1322 |
\begin{eqnarray*} |
1323 |
e^{\Delta tR_1 } & \approx & (1 - \Delta tR_1 )^{ - 1} (1 + \Delta |
1324 |
tR_1 ) \\ |
1325 |
% |
1326 |
& \approx & \left( \begin{array}{ccc} |
1327 |
1 & 0 & 0 \\ |
1328 |
0 & \frac{1-\theta^2 / 4}{1 + \theta^2 / 4} & -\frac{\theta}{1+ |
1329 |
\theta^2 / 4} \\ |
1330 |
0 & \frac{\theta}{1+ \theta^2 / 4} & \frac{1-\theta^2 / 4}{1 + |
1331 |
\theta^2 / 4} |
1332 |
\end{array} |
1333 |
\right). |
1334 |
\end{eqnarray*} |
1335 |
The propagators for $T_2^r$ and $T_3^r$ can be found in the same |
1336 |
manner. In order to construct a second-order symplectic method, we |
1337 |
split the angular kinetic Hamiltonian function into five terms |
1338 |
\[ |
1339 |
T^r (\pi ) = \frac{1}{2}T_1 ^r (\pi _1 ) + \frac{1}{2}T_2^r (\pi _2 |
1340 |
) + T_3^r (\pi _3 ) + \frac{1}{2}T_2^r (\pi _2 ) + \frac{1}{2}T_1 ^r |
1341 |
(\pi _1 ). |
1342 |
\] |
1343 |
By concatenating the propagators corresponding to these five terms, |
1344 |
we can obtain an symplectic integrator, |
1345 |
\[ |
1346 |
\varphi _{\Delta t,T^r } = \varphi _{\Delta t/2,\pi _1 } \circ |
1347 |
\varphi _{\Delta t/2,\pi _2 } \circ \varphi _{\Delta t,\pi _3 } |
1348 |
\circ \varphi _{\Delta t/2,\pi _2 } \circ \varphi _{\Delta t/2,\pi |
1349 |
_1 }. |
1350 |
\] |
1351 |
The non-canonical Lie-Poisson bracket $\{F, G\}$ of two functions $F(\pi )$ and $G(\pi )$ is defined by |
1352 |
\[ |
1353 |
\{ F,G\} (\pi ) = [\nabla _\pi F(\pi )]^T J(\pi )\nabla _\pi G(\pi |
1354 |
). |
1355 |
\] |
1356 |
If the Poisson bracket of a function $F$ with an arbitrary smooth |
1357 |
function $G$ is zero, $F$ is a \emph{Casimir}, which is the |
1358 |
conserved quantity in Poisson system. We can easily verify that the |
1359 |
norm of the angular momentum, $\parallel \pi |
1360 |
\parallel$, is a \emph{Casimir}.\cite{McLachlan1993} Let $F(\pi ) = S(\frac{{\parallel |
1361 |
\pi \parallel ^2 }}{2})$ for an arbitrary function $ S:R \to R$ , |
1362 |
then by the chain rule |
1363 |
\[ |
1364 |
\nabla _\pi F(\pi ) = S'(\frac{{\parallel \pi \parallel ^2 |
1365 |
}}{2})\pi. |
1366 |
\] |
1367 |
Thus, $ [\nabla _\pi F(\pi )]^T J(\pi ) = - S'(\frac{{\parallel |
1368 |
\pi |
1369 |
\parallel ^2 }}{2})\pi \times \pi = 0 $. This explicit |
1370 |
Lie-Poisson integrator is found to be both extremely efficient and |
1371 |
stable. These properties can be explained by the fact the small |
1372 |
angle approximation is used and the norm of the angular momentum is |
1373 |
conserved. |
1374 |
|
1375 |
\subsection{\label{introSection:RBHamiltonianSplitting} Hamiltonian |
1376 |
Splitting for Rigid Body} |
1377 |
|
1378 |
The Hamiltonian of rigid body can be separated in terms of kinetic |
1379 |
energy and potential energy, $H = T(p,\pi ) + V(q,Q)$. The equations |
1380 |
of motion corresponding to potential energy and kinetic energy are |
1381 |
listed in Table~\ref{introTable:rbEquations}. |
1382 |
\begin{table} |
1383 |
\caption{EQUATIONS OF MOTION DUE TO POTENTIAL AND KINETIC ENERGIES} |
1384 |
\label{introTable:rbEquations} |
1385 |
\begin{center} |
1386 |
\begin{tabular}{|l|l|} |
1387 |
\hline |
1388 |
% after \\: \hline or \cline{col1-col2} \cline{col3-col4} ... |
1389 |
Potential & Kinetic \\ |
1390 |
$\frac{{dq}}{{dt}} = \frac{p}{m}$ & $\frac{d}{{dt}}q = p$ \\ |
1391 |
$\frac{d}{{dt}}p = - \frac{{\partial V}}{{\partial q}}$ & $ \frac{d}{{dt}}p = 0$ \\ |
1392 |
$\frac{d}{{dt}}Q = 0$ & $ \frac{d}{{dt}}Q = Qskew(I^{ - 1} j)$ \\ |
1393 |
$ \frac{d}{{dt}}\pi = \sum\limits_i {\left( {Q^T F_i (r,Q)} \right) \times X_i }$ & $\frac{d}{{dt}}\pi = \pi \times I^{ - 1} \pi$\\ |
1394 |
\hline |
1395 |
\end{tabular} |
1396 |
\end{center} |
1397 |
\end{table} |
1398 |
A second-order symplectic method is now obtained by the composition |
1399 |
of the position and velocity propagators, |
1400 |
\[ |
1401 |
\varphi _{\Delta t} = \varphi _{\Delta t/2,V} \circ \varphi |
1402 |
_{\Delta t,T} \circ \varphi _{\Delta t/2,V}. |
1403 |
\] |
1404 |
Moreover, $\varphi _{\Delta t/2,V}$ can be divided into two |
1405 |
sub-propagators which corresponding to force and torque |
1406 |
respectively, |
1407 |
\[ |
1408 |
\varphi _{\Delta t/2,V} = \varphi _{\Delta t/2,F} \circ \varphi |
1409 |
_{\Delta t/2,\tau }. |
1410 |
\] |
1411 |
Since the associated operators of $\varphi _{\Delta t/2,F} $ and |
1412 |
$\circ \varphi _{\Delta t/2,\tau }$ commute, the composition order |
1413 |
inside $\varphi _{\Delta t/2,V}$ does not matter. Furthermore, the |
1414 |
kinetic energy can be separated to translational kinetic term, $T^t |
1415 |
(p)$, and rotational kinetic term, $T^r (\pi )$, |
1416 |
\begin{equation} |
1417 |
T(p,\pi ) =T^t (p) + T^r (\pi ). |
1418 |
\end{equation} |
1419 |
where $ T^t (p) = \frac{1}{2}p^T m^{ - 1} p $ and $T^r (\pi )$ is |
1420 |
defined by Eq.~\ref{introEquation:rotationalKineticRB}. Therefore, |
1421 |
the corresponding propagators are given by |
1422 |
\[ |
1423 |
\varphi _{\Delta t,T} = \varphi _{\Delta t,T^t } \circ \varphi |
1424 |
_{\Delta t,T^r }. |
1425 |
\] |
1426 |
Finally, we obtain the overall symplectic propagators for freely |
1427 |
moving rigid bodies |
1428 |
\begin{eqnarray} |
1429 |
\varphi _{\Delta t} &=& \varphi _{\Delta t/2,F} \circ \varphi _{\Delta t/2,\tau } \notag\\ |
1430 |
& & \circ \varphi _{\Delta t,T^t } \circ \varphi _{\Delta t/2,\pi _1 } \circ \varphi _{\Delta t/2,\pi _2 } \circ \varphi _{\Delta t,\pi _3 } \circ \varphi _{\Delta t/2,\pi _2 } \circ \varphi _{\Delta t/2,\pi _1 } \notag\\ |
1431 |
& & \circ \varphi _{\Delta t/2,\tau } \circ \varphi _{\Delta t/2,F} . |
1432 |
\label{introEquation:overallRBFlowMaps} |
1433 |
\end{eqnarray} |
1434 |
|
1435 |
\section{\label{introSection:langevinDynamics}Langevin Dynamics} |
1436 |
As an alternative to newtonian dynamics, Langevin dynamics, which |
1437 |
mimics a simple heat bath with stochastic and dissipative forces, |
1438 |
has been applied in a variety of studies. This section will review |
1439 |
the theory of Langevin dynamics. A brief derivation of the generalized |
1440 |
Langevin equation will be given first. Following that, we will |
1441 |
discuss the physical meaning of the terms appearing in the equation. |
1442 |
|
1443 |
\subsection{\label{introSection:generalizedLangevinDynamics}Derivation of Generalized Langevin Equation} |
1444 |
|
1445 |
A harmonic bath model, in which an effective set of harmonic |
1446 |
oscillators are used to mimic the effect of a linearly responding |
1447 |
environment, has been widely used in quantum chemistry and |
1448 |
statistical mechanics. One of the successful applications of |
1449 |
Harmonic bath model is the derivation of the Generalized Langevin |
1450 |
Dynamics (GLE). Consider a system, in which the degree of |
1451 |
freedom $x$ is assumed to couple to the bath linearly, giving a |
1452 |
Hamiltonian of the form |
1453 |
\begin{equation} |
1454 |
H = \frac{{p^2 }}{{2m}} + U(x) + H_B + \Delta U(x,x_1 , \ldots x_N) |
1455 |
\label{introEquation:bathGLE}. |
1456 |
\end{equation} |
1457 |
Here $p$ is a momentum conjugate to $x$, $m$ is the mass associated |
1458 |
with this degree of freedom, $H_B$ is a harmonic bath Hamiltonian, |
1459 |
\[ |
1460 |
H_B = \sum\limits_{\alpha = 1}^N {\left\{ {\frac{{p_\alpha ^2 |
1461 |
}}{{2m_\alpha }} + \frac{1}{2}m_\alpha x_\alpha ^2 } |
1462 |
\right\}} |
1463 |
\] |
1464 |
where the index $\alpha$ runs over all the bath degrees of freedom, |
1465 |
$\omega _\alpha$ are the harmonic bath frequencies, $m_\alpha$ are |
1466 |
the harmonic bath masses, and $\Delta U$ is a bilinear system-bath |
1467 |
coupling, |
1468 |
\[ |
1469 |
\Delta U = - \sum\limits_{\alpha = 1}^N {g_\alpha x_\alpha x} |
1470 |
\] |
1471 |
where $g_\alpha$ are the coupling constants between the bath |
1472 |
coordinates ($x_ \alpha$) and the system coordinate ($x$). |
1473 |
Introducing |
1474 |
\[ |
1475 |
W(x) = U(x) - \sum\limits_{\alpha = 1}^N {\frac{{g_\alpha ^2 |
1476 |
}}{{2m_\alpha w_\alpha ^2 }}} x^2 |
1477 |
\] |
1478 |
and combining the last two terms in Eq.~\ref{introEquation:bathGLE}, we may rewrite the Harmonic bath Hamiltonian as |
1479 |
\[ |
1480 |
H = \frac{{p^2 }}{{2m}} + W(x) + \sum\limits_{\alpha = 1}^N |
1481 |
{\left\{ {\frac{{p_\alpha ^2 }}{{2m_\alpha }} + \frac{1}{2}m_\alpha |
1482 |
w_\alpha ^2 \left( {x_\alpha - \frac{{g_\alpha }}{{m_\alpha |
1483 |
w_\alpha ^2 }}x} \right)^2 } \right\}}. |
1484 |
\] |
1485 |
Since the first two terms of the new Hamiltonian depend only on the |
1486 |
system coordinates, we can get the equations of motion for |
1487 |
Generalized Langevin Dynamics by Hamilton's equations, |
1488 |
\begin{equation} |
1489 |
m\ddot x = - \frac{{\partial W(x)}}{{\partial x}} - |
1490 |
\sum\limits_{\alpha = 1}^N {g_\alpha \left( {x_\alpha - |
1491 |
\frac{{g_\alpha }}{{m_\alpha w_\alpha ^2 }}x} \right)}, |
1492 |
\label{introEquation:coorMotionGLE} |
1493 |
\end{equation} |
1494 |
and |
1495 |
\begin{equation} |
1496 |
m\ddot x_\alpha = - m_\alpha w_\alpha ^2 \left( {x_\alpha - |
1497 |
\frac{{g_\alpha }}{{m_\alpha w_\alpha ^2 }}x} \right). |
1498 |
\label{introEquation:bathMotionGLE} |
1499 |
\end{equation} |
1500 |
In order to derive an equation for $x$, the dynamics of the bath |
1501 |
variables $x_\alpha$ must be solved exactly first. As an integral |
1502 |
transform which is particularly useful in solving linear ordinary |
1503 |
differential equations,the Laplace transform is the appropriate tool |
1504 |
to solve this problem. The basic idea is to transform the difficult |
1505 |
differential equations into simple algebra problems which can be |
1506 |
solved easily. Then, by applying the inverse Laplace transform, we |
1507 |
can retrieve the solutions of the original problems. Let $f(t)$ be a |
1508 |
function defined on $ [0,\infty ) $, the Laplace transform of $f(t)$ |
1509 |
is a new function defined as |
1510 |
\[ |
1511 |
L(f(t)) \equiv F(p) = \int_0^\infty {f(t)e^{ - pt} dt} |
1512 |
\] |
1513 |
where $p$ is real and $L$ is called the Laplace Transform |
1514 |
Operator. Below are some important properties of the Laplace transform |
1515 |
\begin{eqnarray*} |
1516 |
L(x + y) & = & L(x) + L(y) \\ |
1517 |
L(ax) & = & aL(x) \\ |
1518 |
L(\dot x) & = & pL(x) - px(0) \\ |
1519 |
L(\ddot x)& = & p^2 L(x) - px(0) - \dot x(0) \\ |
1520 |
L\left( {\int_0^t {g(t - \tau )h(\tau )d\tau } } \right)& = & G(p)H(p) \\ |
1521 |
\end{eqnarray*} |
1522 |
Applying the Laplace transform to the bath coordinates, we obtain |
1523 |
\begin{eqnarray*} |
1524 |
p^2 L(x_\alpha ) - px_\alpha (0) - \dot x_\alpha (0) & = & - \omega _\alpha ^2 L(x_\alpha ) + \frac{{g_\alpha }}{{\omega _\alpha }}L(x), \\ |
1525 |
L(x_\alpha ) & = & \frac{{\frac{{g_\alpha }}{{\omega _\alpha }}L(x) + px_\alpha (0) + \dot x_\alpha (0)}}{{p^2 + \omega _\alpha ^2 }}. \\ |
1526 |
\end{eqnarray*} |
1527 |
In the same way, the system coordinates become |
1528 |
\begin{eqnarray*} |
1529 |
mL(\ddot x) & = & |
1530 |
- \sum\limits_{\alpha = 1}^N {\left\{ { - \frac{{g_\alpha ^2 }}{{m_\alpha \omega _\alpha ^2 }}\frac{p}{{p^2 + \omega _\alpha ^2 }}pL(x) - \frac{p}{{p^2 + \omega _\alpha ^2 }}g_\alpha x_\alpha (0) - \frac{1}{{p^2 + \omega _\alpha ^2 }}g_\alpha \dot x_\alpha (0)} \right\}} \\ |
1531 |
& & - \frac{1}{p}\frac{{\partial W(x)}}{{\partial x}}. |
1532 |
\end{eqnarray*} |
1533 |
With the help of some relatively important inverse Laplace |
1534 |
transformations: |
1535 |
\[ |
1536 |
\begin{array}{c} |
1537 |
L(\cos at) = \frac{p}{{p^2 + a^2 }} \\ |
1538 |
L(\sin at) = \frac{a}{{p^2 + a^2 }} \\ |
1539 |
L(1) = \frac{1}{p} \\ |
1540 |
\end{array} |
1541 |
\] |
1542 |
we obtain |
1543 |
\begin{eqnarray*} |
1544 |
m\ddot x & = & - \frac{{\partial W(x)}}{{\partial x}} - |
1545 |
\sum\limits_{\alpha = 1}^N {\left\{ {\left( { - \frac{{g_\alpha ^2 |
1546 |
}}{{m_\alpha \omega _\alpha ^2 }}} \right)\int_0^t {\cos (\omega |
1547 |
_\alpha t)\dot x(t - \tau )d\tau } } \right\}} \\ |
1548 |
& & + \sum\limits_{\alpha = 1}^N {\left\{ {\left[ {g_\alpha |
1549 |
x_\alpha (0) - \frac{{g_\alpha }}{{m_\alpha \omega _\alpha }}} |
1550 |
\right]\cos (\omega _\alpha t) + \frac{{g_\alpha \dot x_\alpha |
1551 |
(0)}}{{\omega _\alpha }}\sin (\omega _\alpha t)} \right\}}\\ |
1552 |
% |
1553 |
& = & - |
1554 |
\frac{{\partial W(x)}}{{\partial x}} - \int_0^t {\sum\limits_{\alpha |
1555 |
= 1}^N {\left( { - \frac{{g_\alpha ^2 }}{{m_\alpha \omega _\alpha |
1556 |
^2 }}} \right)\cos (\omega _\alpha |
1557 |
t)\dot x(t - \tau )d} \tau } \\ |
1558 |
& & + \sum\limits_{\alpha = 1}^N {\left\{ {\left[ {g_\alpha |
1559 |
x_\alpha (0) - \frac{{g_\alpha }}{{m_\alpha \omega _\alpha }}} |
1560 |
\right]\cos (\omega _\alpha t) + \frac{{g_\alpha \dot x_\alpha |
1561 |
(0)}}{{\omega _\alpha }}\sin (\omega _\alpha t)} \right\}} |
1562 |
\end{eqnarray*} |
1563 |
Introducing a \emph{dynamic friction kernel} |
1564 |
\begin{equation} |
1565 |
\xi (t) = \sum\limits_{\alpha = 1}^N {\left( { - \frac{{g_\alpha ^2 |
1566 |
}}{{m_\alpha \omega _\alpha ^2 }}} \right)\cos (\omega _\alpha t)} |
1567 |
\label{introEquation:dynamicFrictionKernelDefinition} |
1568 |
\end{equation} |
1569 |
and \emph{a random force} |
1570 |
\begin{equation} |
1571 |
R(t) = \sum\limits_{\alpha = 1}^N {\left( {g_\alpha x_\alpha (0) |
1572 |
- \frac{{g_\alpha ^2 }}{{m_\alpha \omega _\alpha ^2 }}x(0)} |
1573 |
\right)\cos (\omega _\alpha t)} + \frac{{\dot x_\alpha |
1574 |
(0)}}{{\omega _\alpha }}\sin (\omega _\alpha t), |
1575 |
\label{introEquation:randomForceDefinition} |
1576 |
\end{equation} |
1577 |
the equation of motion can be rewritten as |
1578 |
\begin{equation} |
1579 |
m\ddot x = - \frac{{\partial W}}{{\partial x}} - \int_0^t {\xi |
1580 |
(t)\dot x(t - \tau )d\tau } + R(t) |
1581 |
\label{introEuqation:GeneralizedLangevinDynamics} |
1582 |
\end{equation} |
1583 |
which is known as the \emph{generalized Langevin equation} (GLE). |
1584 |
|
1585 |
\subsubsection{\label{introSection:randomForceDynamicFrictionKernel}\textbf{Random Force and Dynamic Friction Kernel}} |
1586 |
|
1587 |
One may notice that $R(t)$ depends only on initial conditions, which |
1588 |
implies it is completely deterministic within the context of a |
1589 |
harmonic bath. However, it is easy to verify that $R(t)$ is totally |
1590 |
uncorrelated to $x$ and $\dot x$, $\left\langle {x(t)R(t)} |
1591 |
\right\rangle = 0, \left\langle {\dot x(t)R(t)} \right\rangle = |
1592 |
0.$ This property is what we expect from a truly random process. As |
1593 |
long as the model chosen for $R(t)$ was a gaussian distribution in |
1594 |
general, the stochastic nature of the GLE still remains. |
1595 |
%dynamic friction kernel |
1596 |
The convolution integral |
1597 |
\[ |
1598 |
\int_0^t {\xi (t)\dot x(t - \tau )d\tau } |
1599 |
\] |
1600 |
depends on the entire history of the evolution of $x$, which implies |
1601 |
that the bath retains memory of previous motions. In other words, |
1602 |
the bath requires a finite time to respond to change in the motion |
1603 |
of the system. For a sluggish bath which responds slowly to changes |
1604 |
in the system coordinate, we may regard $\xi(t)$ as a constant |
1605 |
$\xi(t) = \Xi_0$. Hence, the convolution integral becomes |
1606 |
\[ |
1607 |
\int_0^t {\xi (t)\dot x(t - \tau )d\tau } = \xi _0 (x(t) - x(0)) |
1608 |
\] |
1609 |
and Eq.~\ref{introEuqation:GeneralizedLangevinDynamics} becomes |
1610 |
\[ |
1611 |
m\ddot x = - \frac{\partial }{{\partial x}}\left( {W(x) + |
1612 |
\frac{1}{2}\xi _0 (x - x_0 )^2 } \right) + R(t), |
1613 |
\] |
1614 |
which can be used to describe the effect of dynamic caging in |
1615 |
viscous solvents. The other extreme is the bath that responds |
1616 |
infinitely quickly to motions in the system. Thus, $\xi (t)$ can be |
1617 |
taken as a $delta$ function in time: |
1618 |
\[ |
1619 |
\xi (t) = 2\xi _0 \delta (t). |
1620 |
\] |
1621 |
Hence, the convolution integral becomes |
1622 |
\[ |
1623 |
\int_0^t {\xi (t)\dot x(t - \tau )d\tau } = 2\xi _0 \int_0^t |
1624 |
{\delta (t)\dot x(t - \tau )d\tau } = \xi _0 \dot x(t), |
1625 |
\] |
1626 |
and Eq.~\ref{introEuqation:GeneralizedLangevinDynamics} becomes |
1627 |
\begin{equation} |
1628 |
m\ddot x = - \frac{{\partial W(x)}}{{\partial x}} - \xi _0 \dot |
1629 |
x(t) + R(t) \label{introEquation:LangevinEquation} |
1630 |
\end{equation} |
1631 |
which is known as the Langevin equation. The static friction |
1632 |
coefficient $\xi _0$ can either be calculated from spectral density |
1633 |
or be determined by Stokes' law for regular shaped particles. A |
1634 |
brief review on calculating friction tensors for arbitrary shaped |
1635 |
particles is given in Sec.~\ref{introSection:frictionTensor}. |
1636 |
|
1637 |
\subsubsection{\label{introSection:secondFluctuationDissipation}\textbf{The Second Fluctuation Dissipation Theorem}} |
1638 |
|
1639 |
Defining a new set of coordinates |
1640 |
\[ |
1641 |
q_\alpha (t) = x_\alpha (t) - \frac{1}{{m_\alpha \omega _\alpha |
1642 |
^2 }}x(0), |
1643 |
\] |
1644 |
we can rewrite $R(t)$ as |
1645 |
\[ |
1646 |
R(t) = \sum\limits_{\alpha = 1}^N {g_\alpha q_\alpha (t)}. |
1647 |
\] |
1648 |
And since the $q$ coordinates are harmonic oscillators, |
1649 |
\begin{eqnarray*} |
1650 |
\left\langle {q_\alpha ^2 } \right\rangle & = & \frac{{kT}}{{m_\alpha \omega _\alpha ^2 }} \\ |
1651 |
\left\langle {q_\alpha (t)q_\alpha (0)} \right\rangle & = & \left\langle {q_\alpha ^2 (0)} \right\rangle \cos (\omega _\alpha t) \\ |
1652 |
\left\langle {q_\alpha (t)q_\beta (0)} \right\rangle & = &\delta _{\alpha \beta } \left\langle {q_\alpha (t)q_\alpha (0)} \right\rangle \\ |
1653 |
\left\langle {R(t)R(0)} \right\rangle & = & \sum\limits_\alpha {\sum\limits_\beta {g_\alpha g_\beta \left\langle {q_\alpha (t)q_\beta (0)} \right\rangle } } \\ |
1654 |
& = &\sum\limits_\alpha {g_\alpha ^2 \left\langle {q_\alpha ^2 (0)} \right\rangle \cos (\omega _\alpha t)} \\ |
1655 |
& = &kT\xi (t) |
1656 |
\end{eqnarray*} |
1657 |
Thus, we recover the \emph{second fluctuation dissipation theorem} |
1658 |
\begin{equation} |
1659 |
\xi (t) = \left\langle {R(t)R(0)} \right\rangle |
1660 |
\label{introEquation:secondFluctuationDissipation}, |
1661 |
\end{equation} |
1662 |
which acts as a constraint on the possible ways in which one can |
1663 |
model the random force and friction kernel. |