ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/tengDissertation/Introduction.tex
Revision: 2948
Committed: Tue Jul 18 14:35:56 2006 UTC (17 years, 11 months ago) by tim
Content type: application/x-tex
File size: 75389 byte(s)
Log Message:
fix an undef reference

File Contents

# Content
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}Manifolds and Bundles}
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.
469 Given vector spaces $V$ and $W$ over same field $F$, $f: V \to W$ is a linear transformation if
470 \begin{eqnarray*}
471 f(x+y) & = & f(x) + f(y) \\
472 f(ax) & = & af(x)
473 \end{eqnarray*}
474 are always satisfied for any two vectors $x$ and $y$ in $V$ and any scalar $a$ in $F$. One can define the dual vector space $V^*$ of $V$ if any two built-in linear transformations $\phi$ and $\psi$ in $V^*$ satisfy the following definition of addition and scalar multiplication:
475 \begin{eqnarray*}
476 (\phi+\psi)(x) & = & \phi(x)+\psi(x) \\
477 (a\phi)(x) & = & a \phi(x)
478 \end{eqnarray*}
479 for all $a$ in $F$ and $x$ in $V$. For a manifold $M$, one can define a tangent vector of a tangent space $TM_q$ at every point $q$
480 \begin{equation}
481 \dot q = \mathop {\lim }\limits_{t \to 0} \frac{{\phi (t) - \phi (0)}}{t}
482 \end{equation}
483 where $\phi(0)=q$ and $\phi(t) \in M$. One may also define a cotangent space $T^*M_q$ as the dual space of the tangent space $TM_q$. The tangent space and the cotangent space are isomorphic to each other, since they are both real vector spaces with same dimension.
484 The union of tangent spaces at every point of $M$ is called the tangent bundle of $M$ and is denoted by $TM$, while cotangent bundle $T^*M$ is defined as the union of the cotangent spaces to $M$.\cite{Jost2002} For a Hamiltonian system with configuration manifold $V$, the $(q,\dot q)$ phase space is the tangent bundle of the configuration manifold $V$, while the cotangent bundle is represented by $(q,p)$.
485
486 \subsection{\label{introSection:ODE}Ordinary Differential Equations}
487
488 For an ordinary differential system defined as
489 \begin{equation}
490 \dot x = f(x)
491 \end{equation}
492 where $x = x(q,p)$, this system is a canonical Hamiltonian, if
493 $f(x) = J\nabla _x H(x)$. Here, $H = H (q, p)$ is Hamiltonian
494 function and $J$ is the skew-symmetric matrix
495 \begin{equation}
496 J = \left( {\begin{array}{*{20}c}
497 0 & I \\
498 { - I} & 0 \\
499 \end{array}} \right)
500 \label{introEquation:canonicalMatrix}
501 \end{equation}
502 where $I$ is an identity matrix. Using this notation, Hamiltonian
503 system can be rewritten as,
504 \begin{equation}
505 \frac{d}{{dt}}x = J\nabla _x H(x).
506 \label{introEquation:compactHamiltonian}
507 \end{equation}In this case, $f$ is
508 called a \emph{Hamiltonian vector field}. Another generalization of
509 Hamiltonian dynamics is Poisson Dynamics,\cite{Olver1986}
510 \begin{equation}
511 \dot x = J(x)\nabla _x H \label{introEquation:poissonHamiltonian}
512 \end{equation}
513 where the most obvious change being that matrix $J$ now depends on
514 $x$.
515
516 \subsection{\label{introSection:exactFlow}Exact Propagator}
517
518 Let $x(t)$ be the exact solution of the ODE
519 system,
520 \begin{equation}
521 \frac{{dx}}{{dt}} = f(x), \label{introEquation:ODE}
522 \end{equation} we can
523 define its exact propagator $\varphi_\tau$:
524 \[ x(t+\tau)
525 =\varphi_\tau(x(t))
526 \]
527 where $\tau$ is a fixed time step and $\varphi$ is a map from phase
528 space to itself. The propagator has the continuous group property,
529 \begin{equation}
530 \varphi _{\tau _1 } \circ \varphi _{\tau _2 } = \varphi _{\tau _1
531 + \tau _2 } .
532 \end{equation}
533 In particular,
534 \begin{equation}
535 \varphi _\tau \circ \varphi _{ - \tau } = I
536 \end{equation}
537 Therefore, the exact propagator is self-adjoint,
538 \begin{equation}
539 \varphi _\tau = \varphi _{ - \tau }^{ - 1}.
540 \end{equation}
541 In most cases, it is not easy to find the exact propagator
542 $\varphi_\tau$. Instead, we use an approximate map, $\psi_\tau$,
543 which is usually called an integrator. The order of an integrator
544 $\psi_\tau$ is $p$, if the Taylor series of $\psi_\tau$ agree to
545 order $p$,
546 \begin{equation}
547 \psi_\tau(x) = x + \tau f(x) + O(\tau^{p+1})
548 \end{equation}
549
550 \subsection{\label{introSection:geometricProperties}Geometric Properties}
551
552 The hidden geometric properties\cite{Budd1999, Marsden1998} of an
553 ODE and its propagator play important roles in numerical studies.
554 Many of them can be found in systems which occur naturally in
555 applications. Let $\varphi$ be the propagator of Hamiltonian vector
556 field, $\varphi$ is a \emph{symplectic} propagator if it satisfies,
557 \begin{equation}
558 {\varphi '}^T J \varphi ' = J.
559 \end{equation}
560 According to Liouville's theorem, the symplectic volume is invariant
561 under a Hamiltonian propagator, which is the basis for classical
562 statistical mechanics. Furthermore, the propagator of a Hamiltonian
563 vector field on a symplectic manifold can be shown to be a
564 symplectomorphism. As to the Poisson system,
565 \begin{equation}
566 {\varphi '}^T J \varphi ' = J \circ \varphi
567 \end{equation}
568 is the property that must be preserved by the integrator. It is
569 possible to construct a \emph{volume-preserving} propagator for a
570 source free ODE ($ \nabla \cdot f = 0 $), if the propagator
571 satisfies $ \det d\varphi = 1$. One can show easily that a
572 symplectic propagator will be volume-preserving. Changing the
573 variables $y = h(x)$ in an ODE (Eq.~\ref{introEquation:ODE}) will
574 result in a new system,
575 \[
576 \dot y = \tilde f(y) = ((dh \cdot f)h^{ - 1} )(y).
577 \]
578 The vector filed $f$ has reversing symmetry $h$ if $f = - \tilde f$.
579 In other words, the propagator of this vector field is reversible if
580 and only if $ h \circ \varphi ^{ - 1} = \varphi \circ h $. A
581 conserved quantity of a general differential function is a function
582 $ G:R^{2d} \to R^d $ which is constant for all solutions of the ODE
583 $\frac{{dx}}{{dt}} = f(x)$ ,
584 \[
585 \frac{{dG(x(t))}}{{dt}} = 0.
586 \]
587 Using the chain rule, one may obtain,
588 \[
589 \sum\limits_i {\frac{{dG}}{{dx_i }}} f_i (x) = f \cdot \nabla G,
590 \]
591 which is the condition for conserved quantities. For a canonical
592 Hamiltonian system, the time evolution of an arbitrary smooth
593 function $G$ is given by,
594 \begin{eqnarray}
595 \frac{{dG(x(t))}}{{dt}} & = & [\nabla _x G(x(t))]^T \dot x(t) \notag\\
596 & = & [\nabla _x G(x(t))]^T J\nabla _x H(x(t)).
597 \label{introEquation:firstIntegral1}
598 \end{eqnarray}
599 Using poisson bracket notion, Eq.~\ref{introEquation:firstIntegral1}
600 can be rewritten as
601 \[
602 \frac{d}{{dt}}G(x(t)) = \left\{ {G,H} \right\}(x(t)).
603 \]
604 Therefore, the sufficient condition for $G$ to be a conserved
605 quantity of a Hamiltonian system is $\left\{ {G,H} \right\} = 0.$ As
606 is well known, the Hamiltonian (or energy) H of a Hamiltonian system
607 is a conserved quantity, which is due to the fact $\{ H,H\} = 0$.
608 When designing any numerical methods, one should always try to
609 preserve the structural properties of the original ODE and its
610 propagator.
611
612 \subsection{\label{introSection:constructionSymplectic}Construction of Symplectic Methods}
613 A lot of well established and very effective numerical methods have
614 been successful precisely because of their symplectic nature even
615 though this fact was not recognized when they were first
616 constructed. The most famous example is the Verlet-leapfrog method
617 in molecular dynamics. In general, symplectic integrators can be
618 constructed using one of four different methods.
619 \begin{enumerate}
620 \item Generating functions
621 \item Variational methods
622 \item Runge-Kutta methods
623 \item Splitting methods
624 \end{enumerate}
625 Generating functions\cite{Channell1990} tend to lead to methods
626 which are cumbersome and difficult to use. In dissipative systems,
627 variational methods can capture the decay of energy
628 accurately.\cite{Kane2000} Since they are geometrically unstable
629 against non-Hamiltonian perturbations, ordinary implicit Runge-Kutta
630 methods are not suitable for Hamiltonian
631 system.\cite{Cartwright1992} Recently, various high-order explicit
632 Runge-Kutta methods \cite{Owren1992,Chen2003} have been developed to
633 overcome this instability. However, due to computational penalty
634 involved in implementing the Runge-Kutta methods, they have not
635 attracted much attention from the Molecular Dynamics community.
636 Instead, splitting methods have been widely accepted since they
637 exploit natural decompositions of the system.\cite{McLachlan1998,
638 Tuckerman1992}
639
640 \subsubsection{\label{introSection:splittingMethod}\textbf{Splitting Methods}}
641
642 The main idea behind splitting methods is to decompose the discrete
643 $\varphi_h$ as a composition of simpler propagators,
644 \begin{equation}
645 \varphi _h = \varphi _{h_1 } \circ \varphi _{h_2 } \ldots \circ
646 \varphi _{h_n }
647 \label{introEquation:FlowDecomposition}
648 \end{equation}
649 where each of the sub-propagator is chosen such that each represent
650 a simpler integration of the system. Suppose that a Hamiltonian
651 system takes the form,
652 \[
653 H = H_1 + H_2.
654 \]
655 Here, $H_1$ and $H_2$ may represent different physical processes of
656 the system. For instance, they may relate to kinetic and potential
657 energy respectively, which is a natural decomposition of the
658 problem. If $H_1$ and $H_2$ can be integrated using exact
659 propagators $\varphi_1(t)$ and $\varphi_2(t)$, respectively, a
660 simple first order expression is then given by the Lie-Trotter
661 formula\cite{Trotter1959}
662 \begin{equation}
663 \varphi _h = \varphi _{1,h} \circ \varphi _{2,h},
664 \label{introEquation:firstOrderSplitting}
665 \end{equation}
666 where $\varphi _h$ is the result of applying the corresponding
667 continuous $\varphi _i$ over a time $h$. By definition, as
668 $\varphi_i(t)$ is the exact solution of a Hamiltonian system, it
669 must follow that each operator $\varphi_i(t)$ is a symplectic map.
670 It is easy to show that any composition of symplectic propagators
671 yields a symplectic map,
672 \begin{equation}
673 (\varphi '\phi ')^T J\varphi '\phi ' = \phi '^T \varphi '^T J\varphi
674 '\phi ' = \phi '^T J\phi ' = J,
675 \label{introEquation:SymplecticFlowComposition}
676 \end{equation}
677 where $\phi$ and $\psi$ both are symplectic maps. Thus operator
678 splitting in this context automatically generates a symplectic map.
679 The Lie-Trotter
680 splitting(Eq.~\ref{introEquation:firstOrderSplitting}) introduces
681 local errors proportional to $h^2$, while the Strang splitting gives
682 a second-order decomposition,\cite{Strang1968}
683 \begin{equation}
684 \varphi _h = \varphi _{1,h/2} \circ \varphi _{2,h} \circ \varphi
685 _{1,h/2} , \label{introEquation:secondOrderSplitting}
686 \end{equation}
687 which has a local error proportional to $h^3$. The Strang
688 splitting's popularity in molecular simulation community attribute
689 to its symmetric property,
690 \begin{equation}
691 \varphi _h^{ - 1} = \varphi _{ - h}.
692 \label{introEquation:timeReversible}
693 \end{equation}
694
695 \subsubsection{\label{introSection:exampleSplittingMethod}\textbf{Examples of the Splitting Method}}
696 The classical equation for a system consisting of interacting
697 particles can be written in Hamiltonian form,
698 \[
699 H = T + V
700 \]
701 where $T$ is the kinetic energy and $V$ is the potential energy.
702 Setting $H_1 = T, H_2 = V$ and applying the Strang splitting, one
703 obtains the following:
704 \begin{align}
705 q(\Delta t) &= q(0) + \dot{q}(0)\Delta t +
706 \frac{F[q(0)]}{m}\frac{\Delta t^2}{2}, %
707 \label{introEquation:Lp10a} \\%
708 %
709 \dot{q}(\Delta t) &= \dot{q}(0) + \frac{\Delta t}{2m}
710 \biggl [F[q(0)] + F[q(\Delta t)] \biggr]. %
711 \label{introEquation:Lp10b}
712 \end{align}
713 where $F(t)$ is the force at time $t$. This integration scheme is
714 known as \emph{velocity verlet} which is
715 symplectic(Eq.~\ref{introEquation:SymplecticFlowComposition}),
716 time-reversible(Eq.~\ref{introEquation:timeReversible}) and
717 volume-preserving (Eq.~\ref{introEquation:volumePreserving}). These
718 geometric properties attribute to its long-time stability and its
719 popularity in the community. However, the most commonly used
720 velocity verlet integration scheme is written as below,
721 \begin{align}
722 \dot{q}\biggl (\frac{\Delta t}{2}\biggr ) &=
723 \dot{q}(0) + \frac{\Delta t}{2m}\, F[q(0)], \label{introEquation:Lp9a}\\%
724 %
725 q(\Delta t) &= q(0) + \Delta t\, \dot{q}\biggl (\frac{\Delta t}{2}\biggr ),%
726 \label{introEquation:Lp9b}\\%
727 %
728 \dot{q}(\Delta t) &= \dot{q}\biggl (\frac{\Delta t}{2}\biggr ) +
729 \frac{\Delta t}{2m}\, F[q(t)]. \label{introEquation:Lp9c}
730 \end{align}
731 From the preceding splitting, one can see that the integration of
732 the equations of motion would follow:
733 \begin{enumerate}
734 \item calculate the velocities at the half step, $\frac{\Delta t}{2}$, from the forces calculated at the initial position.
735
736 \item Use the half step velocities to move positions one whole step, $\Delta t$.
737
738 \item Evaluate the forces at the new positions, $q(\Delta t)$, and use the new forces to complete the velocity move.
739
740 \item Repeat from step 1 with the new position, velocities, and forces assuming the roles of the initial values.
741 \end{enumerate}
742 By simply switching the order of the propagators in the splitting
743 and composing a new integrator, the \emph{position verlet}
744 integrator, can be generated,
745 \begin{align}
746 \dot q(\Delta t) &= \dot q(0) + \Delta tF(q(0))\left[ {q(0) +
747 \frac{{\Delta t}}{{2m}}\dot q(0)} \right], %
748 \label{introEquation:positionVerlet1} \\%
749 %
750 q(\Delta t) &= q(0) + \frac{{\Delta t}}{2}\left[ {\dot q(0) + \dot
751 q(\Delta t)} \right]. %
752 \label{introEquation:positionVerlet2}
753 \end{align}
754
755 \subsubsection{\label{introSection:errorAnalysis}\textbf{Error Analysis and Higher Order Methods}}
756
757 The Baker-Campbell-Hausdorff formula\cite{Gilmore1974} can be used
758 to determine the local error of a splitting method in terms of the
759 commutator of the
760 operators associated
761 with the sub-propagator. For operators $hX$ and $hY$ which are
762 associated with $\varphi_1(t)$ and $\varphi_2(t)$ respectively , we
763 have
764 \begin{equation}
765 \exp (hX + hY) = \exp (hZ)
766 \end{equation}
767 where
768 \begin{equation}
769 hZ = hX + hY + \frac{{h^2 }}{2}[X,Y] + \frac{{h^3 }}{2}\left(
770 {[X,[X,Y]] + [Y,[Y,X]]} \right) + \ldots .
771 \end{equation}
772 Here, $[X,Y]$ is the commutator of operator $X$ and $Y$ given by
773 \[
774 [X,Y] = XY - YX .
775 \]
776 Applying the Baker-Campbell-Hausdorff formula\cite{Varadarajan1974}
777 to the Strang splitting, we can obtain
778 \begin{eqnarray*}
779 \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 \\
780 & & \mbox{} + h^2 [X,X]/8 + h^2 [Y,Y]/8 \\
781 & & \mbox{} + h^3 [Y,[Y,X]]/12 - h^3[X,[X,Y]]/24 + \ldots
782 ).
783 \end{eqnarray*}
784 Since $ [X,Y] + [Y,X] = 0$ and $ [X,X] = 0$, the dominant local
785 error of Strang splitting is proportional to $h^3$. The same
786 procedure can be applied to a general splitting of the form
787 \begin{equation}
788 \varphi _{b_m h}^2 \circ \varphi _{a_m h}^1 \circ \varphi _{b_{m -
789 1} h}^2 \circ \ldots \circ \varphi _{a_1 h}^1 .
790 \end{equation}
791 A careful choice of coefficient $a_1 \ldots b_m$ will lead to higher
792 order methods. Yoshida proposed an elegant way to compose higher
793 order methods based on symmetric splitting.\cite{Yoshida1990} Given
794 a symmetric second order base method $ \varphi _h^{(2)} $, a
795 fourth-order symmetric method can be constructed by composing,
796 \[
797 \varphi _h^{(4)} = \varphi _{\alpha h}^{(2)} \circ \varphi _{\beta
798 h}^{(2)} \circ \varphi _{\alpha h}^{(2)}
799 \]
800 where $ \alpha = - \frac{{2^{1/3} }}{{2 - 2^{1/3} }}$ and $ \beta
801 = \frac{{2^{1/3} }}{{2 - 2^{1/3} }}$. Moreover, a symmetric
802 integrator $ \varphi _h^{(2n + 2)}$ can be composed by
803 \begin{equation}
804 \varphi _h^{(2n + 2)} = \varphi _{\alpha h}^{(2n)} \circ \varphi
805 _{\beta h}^{(2n)} \circ \varphi _{\alpha h}^{(2n)},
806 \end{equation}
807 if the weights are chosen as
808 \[
809 \alpha = - \frac{{2^{1/(2n + 1)} }}{{2 - 2^{1/(2n + 1)} }},\beta =
810 \frac{{2^{1/(2n + 1)} }}{{2 - 2^{1/(2n + 1)} }} .
811 \]
812
813 \section{\label{introSection:molecularDynamics}Molecular Dynamics}
814
815 As one of the principal tools of molecular modeling, Molecular
816 dynamics has proven to be a powerful tool for studying the functions
817 of biological systems, providing structural, thermodynamic and
818 dynamical information. The basic idea of molecular dynamics is that
819 macroscopic properties are related to microscopic behavior and
820 microscopic behavior can be calculated from the trajectories in
821 simulations. For instance, instantaneous temperature of a
822 Hamiltonian system of $N$ particles can be measured by
823 \[
824 T = \sum\limits_{i = 1}^N {\frac{{m_i v_i^2 }}{{fk_B }}}
825 \]
826 where $m_i$ and $v_i$ are the mass and velocity of $i$th particle
827 respectively, $f$ is the number of degrees of freedom, and $k_B$ is
828 the Boltzman constant.
829
830 A typical molecular dynamics run consists of three essential steps:
831 \begin{enumerate}
832 \item Initialization
833 \begin{enumerate}
834 \item Preliminary preparation
835 \item Minimization
836 \item Heating
837 \item Equilibration
838 \end{enumerate}
839 \item Production
840 \item Analysis
841 \end{enumerate}
842 These three individual steps will be covered in the following
843 sections. Sec.~\ref{introSec:initialSystemSettings} deals with the
844 initialization of a simulation. Sec.~\ref{introSection:production}
845 discusses issues of production runs.
846 Sec.~\ref{introSection:Analysis} provides the theoretical tools for
847 analysis of trajectories.
848
849 \subsection{\label{introSec:initialSystemSettings}Initialization}
850
851 \subsubsection{\textbf{Preliminary preparation}}
852
853 When selecting the starting structure of a molecule for molecular
854 simulation, one may retrieve its Cartesian coordinates from public
855 databases, such as RCSB Protein Data Bank \textit{etc}. Although
856 thousands of crystal structures of molecules are discovered every
857 year, many more remain unknown due to the difficulties of
858 purification and crystallization. Even for molecules with known
859 structures, some important information is missing. For example, a
860 missing hydrogen atom which acts as donor in hydrogen bonding must
861 be added. Moreover, in order to include electrostatic interactions,
862 one may need to specify the partial charges for individual atoms.
863 Under some circumstances, we may even need to prepare the system in
864 a special configuration. For instance, when studying transport
865 phenomenon in membrane systems, we may prepare the lipids in a
866 bilayer structure instead of placing lipids randomly in solvent,
867 since we are not interested in the slow self-aggregation process.
868
869 \subsubsection{\textbf{Minimization}}
870
871 It is quite possible that some of molecules in the system from
872 preliminary preparation may be overlapping with each other. This
873 close proximity leads to high initial potential energy which
874 consequently jeopardizes any molecular dynamics simulations. To
875 remove these steric overlaps, one typically performs energy
876 minimization to find a more reasonable conformation. Several energy
877 minimization methods have been developed to exploit the energy
878 surface and to locate the local minimum. While converging slowly
879 near the minimum, the steepest descent method is extremely robust when
880 systems are strongly anharmonic. Thus, it is often used to refine
881 structures from crystallographic data. Relying on the Hessian,
882 advanced methods like Newton-Raphson converge rapidly to a local
883 minimum, but become unstable if the energy surface is far from
884 quadratic. Another factor that must be taken into account, when
885 choosing energy minimization method, is the size of the system.
886 Steepest descent and conjugate gradient can deal with models of any
887 size. Because of the limits on computer memory to store the hessian
888 matrix and the computing power needed to diagonalize these matrices,
889 most Newton-Raphson methods can not be used with very large systems.
890
891 \subsubsection{\textbf{Heating}}
892
893 Typically, heating is performed by assigning random velocities
894 according to a Maxwell-Boltzman distribution for a desired
895 temperature. Beginning at a lower temperature and gradually
896 increasing the temperature by assigning larger random velocities, we
897 end up setting the temperature of the system to a final temperature
898 at which the simulation will be conducted. In the heating phase, we
899 should also keep the system from drifting or rotating as a whole. To
900 do this, the net linear momentum and angular momentum of the system
901 is shifted to zero after each resampling from the Maxwell -Boltzman
902 distribution.
903
904 \subsubsection{\textbf{Equilibration}}
905
906 The purpose of equilibration is to allow the system to evolve
907 spontaneously for a period of time and reach equilibrium. The
908 procedure is continued until various statistical properties, such as
909 temperature, pressure, energy, volume and other structural
910 properties \textit{etc}, become independent of time. Strictly
911 speaking, minimization and heating are not necessary, provided the
912 equilibration process is long enough. However, these steps can serve
913 as a mean to arrive at an equilibrated structure in an effective
914 way.
915
916 \subsection{\label{introSection:production}Production}
917
918 The production run is the most important step of the simulation, in
919 which the equilibrated structure is used as a starting point and the
920 motions of the molecules are collected for later analysis. In order
921 to capture the macroscopic properties of the system, the molecular
922 dynamics simulation must be performed by sampling correctly and
923 efficiently from the relevant thermodynamic ensemble.
924
925 The most expensive part of a molecular dynamics simulation is the
926 calculation of non-bonded forces, such as van der Waals force and
927 Coulombic forces \textit{etc}. For a system of $N$ particles, the
928 complexity of the algorithm for pair-wise interactions is $O(N^2 )$,
929 which makes large simulations prohibitive in the absence of any
930 algorithmic tricks. A natural approach to avoid system size issues
931 is to represent the bulk behavior by a finite number of the
932 particles. However, this approach will suffer from surface effects
933 at the edges of the simulation. To offset this, \textit{Periodic
934 boundary conditions} (see Fig.~\ref{introFig:pbc}) were developed to
935 simulate bulk properties with a relatively small number of
936 particles. In this method, the simulation box is replicated
937 throughout space to form an infinite lattice. During the simulation,
938 when a particle moves in the primary cell, its image in other cells
939 move in exactly the same direction with exactly the same
940 orientation. Thus, as a particle leaves the primary cell, one of its
941 images will enter through the opposite face.
942 \begin{figure}
943 \centering
944 \includegraphics[width=\linewidth]{pbc.eps}
945 \caption[An illustration of periodic boundary conditions]{A 2-D
946 illustration of periodic boundary conditions. As one particle leaves
947 the left of the simulation box, an image of it enters the right.}
948 \label{introFig:pbc}
949 \end{figure}
950
951 %cutoff and minimum image convention
952 Another important technique to improve the efficiency of force
953 evaluation is to apply spherical cutoffs where particles farther
954 than a predetermined distance are not included in the
955 calculation.\cite{Frenkel1996} The use of a cutoff radius will cause
956 a discontinuity in the potential energy curve. Fortunately, one can
957 shift a simple radial potential to ensure the potential curve go
958 smoothly to zero at the cutoff radius. The cutoff strategy works
959 well for Lennard-Jones interaction because of its short range
960 nature. However, simply truncating the electrostatic interaction
961 with the use of cutoffs has been shown to lead to severe artifacts
962 in simulations. The Ewald summation, in which the slowly decaying
963 Coulomb potential is transformed into direct and reciprocal sums
964 with rapid and absolute convergence, has proved to minimize the
965 periodicity artifacts in liquid simulations. Taking advantage of
966 fast Fourier transform (FFT) techniques for calculating discrete
967 Fourier transforms, the particle mesh-based
968 methods\cite{Hockney1981,Shimada1993, Luty1994} are accelerated from
969 $O(N^{3/2})$ to $O(N logN)$. An alternative approach is the
970 \emph{fast multipole method}\cite{Greengard1987, Greengard1994},
971 which treats Coulombic interactions exactly at short range, and
972 approximate the potential at long range through multipolar
973 expansion. In spite of their wide acceptance at the molecular
974 simulation community, these two methods are difficult to implement
975 correctly and efficiently. Instead, we use a damped and
976 charge-neutralized Coulomb potential method developed by Wolf and
977 his coworkers.\cite{Wolf1999} The shifted Coulomb potential for
978 particle $i$ and particle $j$ at distance $r_{rj}$ is given by:
979 \begin{equation}
980 V(r_{ij})= \frac{q_i q_j \textrm{erfc}(\alpha
981 r_{ij})}{r_{ij}}-\lim_{r_{ij}\rightarrow
982 R_\textrm{c}}\left\{\frac{q_iq_j \textrm{erfc}(\alpha
983 r_{ij})}{r_{ij}}\right\}, \label{introEquation:shiftedCoulomb}
984 \end{equation}
985 where $\alpha$ is the convergence parameter. Due to the lack of
986 inherent periodicity and rapid convergence,this method is extremely
987 efficient and easy to implement.
988 \begin{figure}
989 \centering
990 \includegraphics[width=\linewidth]{shifted_coulomb.eps}
991 \caption[An illustration of shifted Coulomb potential]{An
992 illustration of shifted Coulomb potential.}
993 \label{introFigure:shiftedCoulomb}
994 \end{figure}
995
996 %multiple time step
997
998 \subsection{\label{introSection:Analysis} Analysis}
999
1000 Recently, advanced visualization techniques have been applied to
1001 monitor the motions of molecules. Although the dynamics of the
1002 system can be described qualitatively from animation, quantitative
1003 trajectory analysis is more useful. According to the principles of
1004 Statistical Mechanics in
1005 Sec.~\ref{introSection:statisticalMechanics}, one can compute
1006 thermodynamic properties, analyze fluctuations of structural
1007 parameters, and investigate time-dependent processes of the molecule
1008 from the trajectories.
1009
1010 \subsubsection{\label{introSection:thermodynamicsProperties}\textbf{Thermodynamic Properties}}
1011
1012 Thermodynamic properties, which can be expressed in terms of some
1013 function of the coordinates and momenta of all particles in the
1014 system, can be directly computed from molecular dynamics. The usual
1015 way to measure the pressure is based on virial theorem of Clausius
1016 which states that the virial is equal to $-3Nk_BT$. For a system
1017 with forces between particles, the total virial, $W$, contains the
1018 contribution from external pressure and interaction between the
1019 particles:
1020 \[
1021 W = - 3PV + \left\langle {\sum\limits_{i < j} {r{}_{ij} \cdot
1022 f_{ij} } } \right\rangle
1023 \]
1024 where $f_{ij}$ is the force between particle $i$ and $j$ at a
1025 distance $r_{ij}$. Thus, the expression for the pressure is given
1026 by:
1027 \begin{equation}
1028 P = \frac{{Nk_B T}}{V} - \frac{1}{{3V}}\left\langle {\sum\limits_{i
1029 < j} {r{}_{ij} \cdot f_{ij} } } \right\rangle
1030 \end{equation}
1031
1032 \subsubsection{\label{introSection:structuralProperties}\textbf{Structural Properties}}
1033
1034 Structural Properties of a simple fluid can be described by a set of
1035 distribution functions. Among these functions,the \emph{pair
1036 distribution function}, also known as \emph{radial distribution
1037 function}, is of most fundamental importance to liquid theory.
1038 Experimentally, pair distribution functions can be gathered by
1039 Fourier transforming raw data from a series of neutron diffraction
1040 experiments and integrating over the surface
1041 factor.\cite{Powles1973} The experimental results can serve as a
1042 criterion to justify the correctness of a liquid model. Moreover,
1043 various equilibrium thermodynamic and structural properties can also
1044 be expressed in terms of the radial distribution
1045 function.\cite{Allen1987} The pair distribution functions $g(r)$
1046 gives the probability that a particle $i$ will be located at a
1047 distance $r$ from a another particle $j$ in the system
1048 \begin{equation}
1049 g(r) = \frac{V}{{N^2 }}\left\langle {\sum\limits_i {\sum\limits_{j
1050 \ne i} {\delta (r - r_{ij} )} } } \right\rangle = \frac{\rho
1051 (r)}{\rho}.
1052 \end{equation}
1053 Note that the delta function can be replaced by a histogram in
1054 computer simulation. Peaks in $g(r)$ represent solvent shells, and
1055 the height of these peaks gradually decreases to 1 as the liquid of
1056 large distance approaches the bulk density.
1057
1058
1059 \subsubsection{\label{introSection:timeDependentProperties}\textbf{Time-dependent
1060 Properties}}
1061
1062 Time-dependent properties are usually calculated using \emph{time
1063 correlation functions}, which correlate random variables $A$ and $B$
1064 at two different times,
1065 \begin{equation}
1066 C_{AB} (t) = \left\langle {A(t)B(0)} \right\rangle.
1067 \label{introEquation:timeCorrelationFunction}
1068 \end{equation}
1069 If $A$ and $B$ refer to same variable, this kind of correlation
1070 functions are called \emph{autocorrelation functions}. One typical example is the velocity autocorrelation
1071 function which is directly related to transport properties of
1072 molecular liquids:
1073 \begin{equation}
1074 D = \frac{1}{3}\int\limits_0^\infty {\left\langle {v(t) \cdot v(0)}
1075 \right\rangle } dt
1076 \end{equation}
1077 where $D$ is diffusion constant. Unlike the velocity autocorrelation
1078 function, which is averaged over time origins and over all the
1079 atoms, the dipole autocorrelation functions is calculated for the
1080 entire system. The dipole autocorrelation function is given by:
1081 \begin{equation}
1082 c_{dipole} = \left\langle {u_{tot} (t) \cdot u_{tot} (t)}
1083 \right\rangle
1084 \end{equation}
1085 Here $u_{tot}$ is the net dipole of the entire system and is given
1086 by
1087 \begin{equation}
1088 u_{tot} (t) = \sum\limits_i {u_i (t)}.
1089 \end{equation}
1090 In principle, many time correlation functions can be related to
1091 Fourier transforms of the infrared, Raman, and inelastic neutron
1092 scattering spectra of molecular liquids. In practice, one can
1093 extract the IR spectrum from the intensity of the molecular dipole
1094 fluctuation at each frequency using the following relationship:
1095 \begin{equation}
1096 \hat c_{dipole} (v) = \int_{ - \infty }^\infty {c_{dipole} (t)e^{ -
1097 i2\pi vt} dt}.
1098 \end{equation}
1099
1100 \section{\label{introSection:rigidBody}Dynamics of Rigid Bodies}
1101
1102 Rigid bodies are frequently involved in the modeling of different
1103 areas, including engineering, physics and chemistry. For example,
1104 missiles and vehicles are usually modeled by rigid bodies. The
1105 movement of the objects in 3D gaming engines or other physics
1106 simulators is governed by rigid body dynamics. In molecular
1107 simulations, rigid bodies are used to simplify protein-protein
1108 docking studies.\cite{Gray2003}
1109
1110 It is very important to develop stable and efficient methods to
1111 integrate the equations of motion for orientational degrees of
1112 freedom. Euler angles are the natural choice to describe the
1113 rotational degrees of freedom. However, due to $\frac {1}{sin
1114 \theta}$ singularities, the numerical integration of corresponding
1115 equations of these motion is very inefficient and inaccurate.
1116 Although an alternative integrator using multiple sets of Euler
1117 angles can overcome this difficulty\cite{Barojas1973}, the
1118 computational penalty and the loss of angular momentum conservation
1119 still remain. A singularity-free representation utilizing
1120 quaternions was developed by Evans in 1977.\cite{Evans1977}
1121 Unfortunately, this approach used a nonseparable Hamiltonian
1122 resulting from the quaternion representation, which prevented the
1123 symplectic algorithm from being utilized. Another different approach
1124 is to apply holonomic constraints to the atoms belonging to the
1125 rigid body. Each atom moves independently under the normal forces
1126 deriving from potential energy and constraint forces which are used
1127 to guarantee the rigidness. However, due to their iterative nature,
1128 the SHAKE and Rattle algorithms also converge very slowly when the
1129 number of constraints increases.\cite{Ryckaert1977, Andersen1983}
1130
1131 A break-through in geometric literature suggests that, in order to
1132 develop a long-term integration scheme, one should preserve the
1133 symplectic structure of the propagator. By introducing a conjugate
1134 momentum to the rotation matrix $Q$ and re-formulating Hamiltonian's
1135 equation, a symplectic integrator, RSHAKE\cite{Kol1997}, was
1136 proposed to evolve the Hamiltonian system in a constraint manifold
1137 by iteratively satisfying the orthogonality constraint $Q^T Q = 1$.
1138 An alternative method using the quaternion representation was
1139 developed by Omelyan.\cite{Omelyan1998} However, both of these
1140 methods are iterative and inefficient. In this section, we descibe a
1141 symplectic Lie-Poisson integrator for rigid bodies developed by
1142 Dullweber and his coworkers\cite{Dullweber1997} in depth.
1143
1144 \subsection{\label{introSection:constrainedHamiltonianRB}Constrained Hamiltonian for Rigid Bodies}
1145 The Hamiltonian of a rigid body is given by
1146 \begin{equation}
1147 H = \frac{1}{2}(p^T m^{ - 1} p) + \frac{1}{2}tr(PJ^{ - 1} P) +
1148 V(q,Q) + \frac{1}{2}tr[(QQ^T - 1)\Lambda ].
1149 \label{introEquation:RBHamiltonian}
1150 \end{equation}
1151 Here, $q$ and $Q$ are the position vector and rotation matrix for
1152 the rigid-body, $p$ and $P$ are conjugate momenta to $q$ and $Q$ ,
1153 and $J$, a diagonal matrix, is defined by
1154 \[
1155 I_{ii}^{ - 1} = \frac{1}{2}\sum\limits_{i \ne j} {J_{jj}^{ - 1} }
1156 \]
1157 where $I_{ii}$ is the diagonal element of the inertia tensor. This
1158 constrained Hamiltonian equation is subjected to a holonomic
1159 constraint,
1160 \begin{equation}
1161 Q^T Q = 1, \label{introEquation:orthogonalConstraint}
1162 \end{equation}
1163 which is used to ensure the rotation matrix's unitarity. Using
1164 Eq.~\ref{introEquation:motionHamiltonianCoordinate} and Eq.~
1165 \ref{introEquation:motionHamiltonianMomentum}, one can write down
1166 the equations of motion,
1167 \begin{eqnarray}
1168 \frac{{dq}}{{dt}} & = & \frac{p}{m}, \label{introEquation:RBMotionPosition}\\
1169 \frac{{dp}}{{dt}} & = & - \nabla _q V(q,Q), \label{introEquation:RBMotionMomentum}\\
1170 \frac{{dQ}}{{dt}} & = & PJ^{ - 1}, \label{introEquation:RBMotionRotation}\\
1171 \frac{{dP}}{{dt}} & = & - \nabla _Q V(q,Q) - 2Q\Lambda . \label{introEquation:RBMotionP}
1172 \end{eqnarray}
1173 Differentiating Eq.~\ref{introEquation:orthogonalConstraint} and
1174 using Eq.~\ref{introEquation:RBMotionMomentum}, one may obtain,
1175 \begin{equation}
1176 Q^T PJ^{ - 1} + J^{ - 1} P^T Q = 0 . \\
1177 \label{introEquation:RBFirstOrderConstraint}
1178 \end{equation}
1179 In general, there are two ways to satisfy the holonomic constraints.
1180 We can use a constraint force provided by a Lagrange multiplier on
1181 the normal manifold to keep the motion on the constraint space. Or
1182 we can simply evolve the system on the constraint manifold. These
1183 two methods have been proved to be equivalent. The holonomic
1184 constraint and equations of motions define a constraint manifold for
1185 rigid bodies
1186 \[
1187 M = \left\{ {(Q,P):Q^T Q = 1,Q^T PJ^{ - 1} + J^{ - 1} P^T Q = 0}
1188 \right\}.
1189 \]
1190 Unfortunately, this constraint manifold is not $T^* SO(3)$ which is
1191 a symplectic manifold on Lie rotation group $SO(3)$. However, it
1192 turns out that under symplectic transformation, the cotangent space
1193 and the phase space are diffeomorphic. By introducing
1194 \[
1195 \tilde Q = Q,\tilde P = \frac{1}{2}\left( {P - QP^T Q} \right),
1196 \]
1197 the mechanical system subjected to a holonomic constraint manifold $M$
1198 can be re-formulated as a Hamiltonian system on the cotangent space
1199 \[
1200 T^* SO(3) = \left\{ {(\tilde Q,\tilde P):\tilde Q^T \tilde Q =
1201 1,\tilde Q^T \tilde PJ^{ - 1} + J^{ - 1} P^T \tilde Q = 0} \right\}
1202 \]
1203 For a body fixed vector $X_i$ with respect to the center of mass of
1204 the rigid body, its corresponding lab fixed vector $X_i^{lab}$ is
1205 given as
1206 \begin{equation}
1207 X_i^{lab} = Q X_i + q.
1208 \end{equation}
1209 Therefore, potential energy $V(q,Q)$ is defined by
1210 \[
1211 V(q,Q) = V(Q X_0 + q).
1212 \]
1213 Hence, the force and torque are given by
1214 \[
1215 \nabla _q V(q,Q) = F(q,Q) = \sum\limits_i {F_i (q,Q)},
1216 \]
1217 and
1218 \[
1219 \nabla _Q V(q,Q) = F(q,Q)X_i^t
1220 \]
1221 respectively. As a common choice to describe the rotation dynamics
1222 of the rigid body, the angular momentum on the body fixed frame $\Pi
1223 = Q^t P$ is introduced to rewrite the equations of motion,
1224 \begin{equation}
1225 \begin{array}{l}
1226 \dot \Pi = J^{ - 1} \Pi ^T \Pi + Q^T \sum\limits_i {F_i (q,Q)X_i^T } - \Lambda, \\
1227 \dot Q = Q\Pi {\rm{ }}J^{ - 1}, \\
1228 \end{array}
1229 \label{introEqaution:RBMotionPI}
1230 \end{equation}
1231 as well as holonomic constraints $\Pi J^{ - 1} + J^{ - 1} \Pi ^t =
1232 0$ and $Q^T Q = 1$. For a vector $v(v_1 ,v_2 ,v_3 ) \in R^3$ and a
1233 matrix $\hat v \in so(3)^ \star$, the hat-map isomorphism,
1234 \begin{equation}
1235 v(v_1 ,v_2 ,v_3 ) \Leftrightarrow \hat v = \left(
1236 {\begin{array}{*{20}c}
1237 0 & { - v_3 } & {v_2 } \\
1238 {v_3 } & 0 & { - v_1 } \\
1239 { - v_2 } & {v_1 } & 0 \\
1240 \end{array}} \right),
1241 \label{introEquation:hatmapIsomorphism}
1242 \end{equation}
1243 will let us associate the matrix products with traditional vector
1244 operations
1245 \[
1246 \hat vu = v \times u.
1247 \]
1248 Using Eq.~\ref{introEqaution:RBMotionPI}, one can construct a skew
1249 matrix,
1250 \begin{eqnarray}
1251 (\dot \Pi - \dot \Pi ^T )&= &(\Pi - \Pi ^T )(J^{ - 1} \Pi + \Pi J^{ - 1} ) \notag \\
1252 & & + \sum\limits_i {[Q^T F_i (r,Q)X_i^T - X_i F_i (r,Q)^T Q]} -
1253 (\Lambda - \Lambda ^T ). \label{introEquation:skewMatrixPI}
1254 \end{eqnarray}
1255 Since $\Lambda$ is symmetric, the last term of
1256 Eq.~\ref{introEquation:skewMatrixPI} is zero, which implies the
1257 Lagrange multiplier $\Lambda$ is absent from the equations of
1258 motion. This unique property eliminates the requirement of
1259 iterations which can not be avoided in other methods.\cite{Kol1997,
1260 Omelyan1998} Applying the hat-map isomorphism, we obtain the
1261 equation of motion for angular momentum in the body frame
1262 \begin{equation}
1263 \dot \pi = \pi \times I^{ - 1} \pi + \sum\limits_i {\left( {Q^T
1264 F_i (r,Q)} \right) \times X_i }.
1265 \label{introEquation:bodyAngularMotion}
1266 \end{equation}
1267 In the same manner, the equation of motion for rotation matrix is
1268 given by
1269 \[
1270 \dot Q = Qskew(I^{ - 1} \pi ).
1271 \]
1272
1273 \subsection{\label{introSection:SymplecticFreeRB}Symplectic
1274 Lie-Poisson Integrator for Free Rigid Bodies}
1275
1276 If there are no external forces exerted on the rigid body, the only
1277 contribution to the rotational motion is from the kinetic energy
1278 (the first term of \ref{introEquation:bodyAngularMotion}). The free
1279 rigid body is an example of a Lie-Poisson system with Hamiltonian
1280 function
1281 \begin{equation}
1282 T^r (\pi ) = T_1 ^r (\pi _1 ) + T_2^r (\pi _2 ) + T_3^r (\pi _3 )
1283 \label{introEquation:rotationalKineticRB}
1284 \end{equation}
1285 where $T_i^r (\pi _i ) = \frac{{\pi _i ^2 }}{{2I_i }}$ and
1286 Lie-Poisson structure matrix,
1287 \begin{equation}
1288 J(\pi ) = \left( {\begin{array}{*{20}c}
1289 0 & {\pi _3 } & { - \pi _2 } \\
1290 { - \pi _3 } & 0 & {\pi _1 } \\
1291 {\pi _2 } & { - \pi _1 } & 0 \\
1292 \end{array}} \right).
1293 \end{equation}
1294 Thus, the dynamics of free rigid body is governed by
1295 \begin{equation}
1296 \frac{d}{{dt}}\pi = J(\pi )\nabla _\pi T^r (\pi ).
1297 \end{equation}
1298 One may notice that each $T_i^r$ in
1299 Eq.~\ref{introEquation:rotationalKineticRB} can be solved exactly.
1300 For instance, the equations of motion due to $T_1^r$ are given by
1301 \begin{equation}
1302 \frac{d}{{dt}}\pi = R_1 \pi ,\frac{d}{{dt}}Q = QR_1
1303 \label{introEqaution:RBMotionSingleTerm}
1304 \end{equation}
1305 with
1306 \[ R_1 = \left( {\begin{array}{*{20}c}
1307 0 & 0 & 0 \\
1308 0 & 0 & {\pi _1 } \\
1309 0 & { - \pi _1 } & 0 \\
1310 \end{array}} \right).
1311 \]
1312 The solutions of Eq.~\ref{introEqaution:RBMotionSingleTerm} is
1313 \[
1314 \pi (\Delta t) = e^{\Delta tR_1 } \pi (0),Q(\Delta t) =
1315 Q(0)e^{\Delta tR_1 }
1316 \]
1317 with
1318 \[
1319 e^{\Delta tR_1 } = \left( {\begin{array}{*{20}c}
1320 0 & 0 & 0 \\
1321 0 & {\cos \theta _1 } & {\sin \theta _1 } \\
1322 0 & { - \sin \theta _1 } & {\cos \theta _1 } \\
1323 \end{array}} \right),\theta _1 = \frac{{\pi _1 }}{{I_1 }}\Delta t.
1324 \]
1325 To reduce the cost of computing expensive functions in $e^{\Delta
1326 tR_1 }$, we can use the Cayley transformation to obtain a
1327 single-aixs propagator,
1328 \begin{eqnarray*}
1329 e^{\Delta tR_1 } & \approx & (1 - \Delta tR_1 )^{ - 1} (1 + \Delta
1330 tR_1 ) \\
1331 %
1332 & \approx & \left( \begin{array}{ccc}
1333 1 & 0 & 0 \\
1334 0 & \frac{1-\theta^2 / 4}{1 + \theta^2 / 4} & -\frac{\theta}{1+
1335 \theta^2 / 4} \\
1336 0 & \frac{\theta}{1+ \theta^2 / 4} & \frac{1-\theta^2 / 4}{1 +
1337 \theta^2 / 4}
1338 \end{array}
1339 \right).
1340 \end{eqnarray*}
1341 The propagators for $T_2^r$ and $T_3^r$ can be found in the same
1342 manner. In order to construct a second-order symplectic method, we
1343 split the angular kinetic Hamiltonian function into five terms
1344 \[
1345 T^r (\pi ) = \frac{1}{2}T_1 ^r (\pi _1 ) + \frac{1}{2}T_2^r (\pi _2
1346 ) + T_3^r (\pi _3 ) + \frac{1}{2}T_2^r (\pi _2 ) + \frac{1}{2}T_1 ^r
1347 (\pi _1 ).
1348 \]
1349 By concatenating the propagators corresponding to these five terms,
1350 we can obtain an symplectic integrator,
1351 \[
1352 \varphi _{\Delta t,T^r } = \varphi _{\Delta t/2,\pi _1 } \circ
1353 \varphi _{\Delta t/2,\pi _2 } \circ \varphi _{\Delta t,\pi _3 }
1354 \circ \varphi _{\Delta t/2,\pi _2 } \circ \varphi _{\Delta t/2,\pi
1355 _1 }.
1356 \]
1357 The non-canonical Lie-Poisson bracket $\{F, G\}$ of two functions $F(\pi )$ and $G(\pi )$ is defined by
1358 \[
1359 \{ F,G\} (\pi ) = [\nabla _\pi F(\pi )]^T J(\pi )\nabla _\pi G(\pi
1360 ).
1361 \]
1362 If the Poisson bracket of a function $F$ with an arbitrary smooth
1363 function $G$ is zero, $F$ is a \emph{Casimir}, which is the
1364 conserved quantity in Poisson system. We can easily verify that the
1365 norm of the angular momentum, $\parallel \pi
1366 \parallel$, is a \emph{Casimir}.\cite{McLachlan1993} Let $F(\pi ) = S(\frac{{\parallel
1367 \pi \parallel ^2 }}{2})$ for an arbitrary function $ S:R \to R$ ,
1368 then by the chain rule
1369 \[
1370 \nabla _\pi F(\pi ) = S'(\frac{{\parallel \pi \parallel ^2
1371 }}{2})\pi.
1372 \]
1373 Thus, $ [\nabla _\pi F(\pi )]^T J(\pi ) = - S'(\frac{{\parallel
1374 \pi
1375 \parallel ^2 }}{2})\pi \times \pi = 0 $. This explicit
1376 Lie-Poisson integrator is found to be both extremely efficient and
1377 stable. These properties can be explained by the fact the small
1378 angle approximation is used and the norm of the angular momentum is
1379 conserved.
1380
1381 \subsection{\label{introSection:RBHamiltonianSplitting} Hamiltonian
1382 Splitting for Rigid Body}
1383
1384 The Hamiltonian of rigid body can be separated in terms of kinetic
1385 energy and potential energy, $H = T(p,\pi ) + V(q,Q)$. The equations
1386 of motion corresponding to potential energy and kinetic energy are
1387 listed in Table~\ref{introTable:rbEquations}.
1388 \begin{table}
1389 \caption{EQUATIONS OF MOTION DUE TO POTENTIAL AND KINETIC ENERGIES}
1390 \label{introTable:rbEquations}
1391 \begin{center}
1392 \begin{tabular}{|l|l|}
1393 \hline
1394 % after \\: \hline or \cline{col1-col2} \cline{col3-col4} ...
1395 Potential & Kinetic \\
1396 $\frac{{dq}}{{dt}} = \frac{p}{m}$ & $\frac{d}{{dt}}q = p$ \\
1397 $\frac{d}{{dt}}p = - \frac{{\partial V}}{{\partial q}}$ & $ \frac{d}{{dt}}p = 0$ \\
1398 $\frac{d}{{dt}}Q = 0$ & $ \frac{d}{{dt}}Q = Qskew(I^{ - 1} j)$ \\
1399 $ \frac{d}{{dt}}\pi = \sum\limits_i {\left( {Q^T F_i (r,Q)} \right) \times X_i }$ & $\frac{d}{{dt}}\pi = \pi \times I^{ - 1} \pi$\\
1400 \hline
1401 \end{tabular}
1402 \end{center}
1403 \end{table}
1404 A second-order symplectic method is now obtained by the composition
1405 of the position and velocity propagators,
1406 \[
1407 \varphi _{\Delta t} = \varphi _{\Delta t/2,V} \circ \varphi
1408 _{\Delta t,T} \circ \varphi _{\Delta t/2,V}.
1409 \]
1410 Moreover, $\varphi _{\Delta t/2,V}$ can be divided into two
1411 sub-propagators which corresponding to force and torque
1412 respectively,
1413 \[
1414 \varphi _{\Delta t/2,V} = \varphi _{\Delta t/2,F} \circ \varphi
1415 _{\Delta t/2,\tau }.
1416 \]
1417 Since the associated operators of $\varphi _{\Delta t/2,F} $ and
1418 $\circ \varphi _{\Delta t/2,\tau }$ commute, the composition order
1419 inside $\varphi _{\Delta t/2,V}$ does not matter. Furthermore, the
1420 kinetic energy can be separated to translational kinetic term, $T^t
1421 (p)$, and rotational kinetic term, $T^r (\pi )$,
1422 \begin{equation}
1423 T(p,\pi ) =T^t (p) + T^r (\pi ).
1424 \end{equation}
1425 where $ T^t (p) = \frac{1}{2}p^T m^{ - 1} p $ and $T^r (\pi )$ is
1426 defined by Eq.~\ref{introEquation:rotationalKineticRB}. Therefore,
1427 the corresponding propagators are given by
1428 \[
1429 \varphi _{\Delta t,T} = \varphi _{\Delta t,T^t } \circ \varphi
1430 _{\Delta t,T^r }.
1431 \]
1432 Finally, we obtain the overall symplectic propagators for freely
1433 moving rigid bodies
1434 \begin{eqnarray}
1435 \varphi _{\Delta t} &=& \varphi _{\Delta t/2,F} \circ \varphi _{\Delta t/2,\tau } \notag\\
1436 & & \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\\
1437 & & \circ \varphi _{\Delta t/2,\tau } \circ \varphi _{\Delta t/2,F} .
1438 \label{introEquation:overallRBFlowMaps}
1439 \end{eqnarray}
1440
1441 \section{\label{introSection:langevinDynamics}Langevin Dynamics}
1442 As an alternative to newtonian dynamics, Langevin dynamics, which
1443 mimics a simple heat bath with stochastic and dissipative forces,
1444 has been applied in a variety of studies. This section will review
1445 the theory of Langevin dynamics. A brief derivation of the generalized
1446 Langevin equation will be given first. Following that, we will
1447 discuss the physical meaning of the terms appearing in the equation.
1448
1449 \subsection{\label{introSection:generalizedLangevinDynamics}Derivation of Generalized Langevin Equation}
1450
1451 A harmonic bath model, in which an effective set of harmonic
1452 oscillators are used to mimic the effect of a linearly responding
1453 environment, has been widely used in quantum chemistry and
1454 statistical mechanics. One of the successful applications of
1455 Harmonic bath model is the derivation of the Generalized Langevin
1456 Dynamics (GLE). Consider a system, in which the degree of
1457 freedom $x$ is assumed to couple to the bath linearly, giving a
1458 Hamiltonian of the form
1459 \begin{equation}
1460 H = \frac{{p^2 }}{{2m}} + U(x) + H_B + \Delta U(x,x_1 , \ldots x_N)
1461 \label{introEquation:bathGLE}.
1462 \end{equation}
1463 Here $p$ is a momentum conjugate to $x$, $m$ is the mass associated
1464 with this degree of freedom, $H_B$ is a harmonic bath Hamiltonian,
1465 \[
1466 H_B = \sum\limits_{\alpha = 1}^N {\left\{ {\frac{{p_\alpha ^2
1467 }}{{2m_\alpha }} + \frac{1}{2}m_\alpha x_\alpha ^2 }
1468 \right\}}
1469 \]
1470 where the index $\alpha$ runs over all the bath degrees of freedom,
1471 $\omega _\alpha$ are the harmonic bath frequencies, $m_\alpha$ are
1472 the harmonic bath masses, and $\Delta U$ is a bilinear system-bath
1473 coupling,
1474 \[
1475 \Delta U = - \sum\limits_{\alpha = 1}^N {g_\alpha x_\alpha x}
1476 \]
1477 where $g_\alpha$ are the coupling constants between the bath
1478 coordinates ($x_ \alpha$) and the system coordinate ($x$).
1479 Introducing
1480 \[
1481 W(x) = U(x) - \sum\limits_{\alpha = 1}^N {\frac{{g_\alpha ^2
1482 }}{{2m_\alpha w_\alpha ^2 }}} x^2
1483 \]
1484 and combining the last two terms in Eq.~\ref{introEquation:bathGLE}, we may rewrite the Harmonic bath Hamiltonian as
1485 \[
1486 H = \frac{{p^2 }}{{2m}} + W(x) + \sum\limits_{\alpha = 1}^N
1487 {\left\{ {\frac{{p_\alpha ^2 }}{{2m_\alpha }} + \frac{1}{2}m_\alpha
1488 w_\alpha ^2 \left( {x_\alpha - \frac{{g_\alpha }}{{m_\alpha
1489 w_\alpha ^2 }}x} \right)^2 } \right\}}.
1490 \]
1491 Since the first two terms of the new Hamiltonian depend only on the
1492 system coordinates, we can get the equations of motion for
1493 Generalized Langevin Dynamics by Hamilton's equations,
1494 \begin{equation}
1495 m\ddot x = - \frac{{\partial W(x)}}{{\partial x}} -
1496 \sum\limits_{\alpha = 1}^N {g_\alpha \left( {x_\alpha -
1497 \frac{{g_\alpha }}{{m_\alpha w_\alpha ^2 }}x} \right)},
1498 \label{introEquation:coorMotionGLE}
1499 \end{equation}
1500 and
1501 \begin{equation}
1502 m\ddot x_\alpha = - m_\alpha w_\alpha ^2 \left( {x_\alpha -
1503 \frac{{g_\alpha }}{{m_\alpha w_\alpha ^2 }}x} \right).
1504 \label{introEquation:bathMotionGLE}
1505 \end{equation}
1506 In order to derive an equation for $x$, the dynamics of the bath
1507 variables $x_\alpha$ must be solved exactly first. As an integral
1508 transform which is particularly useful in solving linear ordinary
1509 differential equations,the Laplace transform is the appropriate tool
1510 to solve this problem. The basic idea is to transform the difficult
1511 differential equations into simple algebra problems which can be
1512 solved easily. Then, by applying the inverse Laplace transform, we
1513 can retrieve the solutions of the original problems. Let $f(t)$ be a
1514 function defined on $ [0,\infty ) $, the Laplace transform of $f(t)$
1515 is a new function defined as
1516 \[
1517 L(f(t)) \equiv F(p) = \int_0^\infty {f(t)e^{ - pt} dt}
1518 \]
1519 where $p$ is real and $L$ is called the Laplace Transform
1520 Operator. Below are some important properties of the Laplace transform
1521 \begin{eqnarray*}
1522 L(x + y) & = & L(x) + L(y) \\
1523 L(ax) & = & aL(x) \\
1524 L(\dot x) & = & pL(x) - px(0) \\
1525 L(\ddot x)& = & p^2 L(x) - px(0) - \dot x(0) \\
1526 L\left( {\int_0^t {g(t - \tau )h(\tau )d\tau } } \right)& = & G(p)H(p) \\
1527 \end{eqnarray*}
1528 Applying the Laplace transform to the bath coordinates, we obtain
1529 \begin{eqnarray*}
1530 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), \\
1531 L(x_\alpha ) & = & \frac{{\frac{{g_\alpha }}{{\omega _\alpha }}L(x) + px_\alpha (0) + \dot x_\alpha (0)}}{{p^2 + \omega _\alpha ^2 }}. \\
1532 \end{eqnarray*}
1533 In the same way, the system coordinates become
1534 \begin{eqnarray*}
1535 mL(\ddot x) & = &
1536 - \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\}} \\
1537 & & - \frac{1}{p}\frac{{\partial W(x)}}{{\partial x}}.
1538 \end{eqnarray*}
1539 With the help of some relatively important inverse Laplace
1540 transformations:
1541 \[
1542 \begin{array}{c}
1543 L(\cos at) = \frac{p}{{p^2 + a^2 }} \\
1544 L(\sin at) = \frac{a}{{p^2 + a^2 }} \\
1545 L(1) = \frac{1}{p} \\
1546 \end{array}
1547 \]
1548 we obtain
1549 \begin{eqnarray*}
1550 m\ddot x & = & - \frac{{\partial W(x)}}{{\partial x}} -
1551 \sum\limits_{\alpha = 1}^N {\left\{ {\left( { - \frac{{g_\alpha ^2
1552 }}{{m_\alpha \omega _\alpha ^2 }}} \right)\int_0^t {\cos (\omega
1553 _\alpha t)\dot x(t - \tau )d\tau } } \right\}} \\
1554 & & + \sum\limits_{\alpha = 1}^N {\left\{ {\left[ {g_\alpha
1555 x_\alpha (0) - \frac{{g_\alpha }}{{m_\alpha \omega _\alpha }}}
1556 \right]\cos (\omega _\alpha t) + \frac{{g_\alpha \dot x_\alpha
1557 (0)}}{{\omega _\alpha }}\sin (\omega _\alpha t)} \right\}}\\
1558 %
1559 & = & -
1560 \frac{{\partial W(x)}}{{\partial x}} - \int_0^t {\sum\limits_{\alpha
1561 = 1}^N {\left( { - \frac{{g_\alpha ^2 }}{{m_\alpha \omega _\alpha
1562 ^2 }}} \right)\cos (\omega _\alpha
1563 t)\dot x(t - \tau )d} \tau } \\
1564 & & + \sum\limits_{\alpha = 1}^N {\left\{ {\left[ {g_\alpha
1565 x_\alpha (0) - \frac{{g_\alpha }}{{m_\alpha \omega _\alpha }}}
1566 \right]\cos (\omega _\alpha t) + \frac{{g_\alpha \dot x_\alpha
1567 (0)}}{{\omega _\alpha }}\sin (\omega _\alpha t)} \right\}}
1568 \end{eqnarray*}
1569 Introducing a \emph{dynamic friction kernel}
1570 \begin{equation}
1571 \xi (t) = \sum\limits_{\alpha = 1}^N {\left( { - \frac{{g_\alpha ^2
1572 }}{{m_\alpha \omega _\alpha ^2 }}} \right)\cos (\omega _\alpha t)}
1573 \label{introEquation:dynamicFrictionKernelDefinition}
1574 \end{equation}
1575 and \emph{a random force}
1576 \begin{equation}
1577 R(t) = \sum\limits_{\alpha = 1}^N {\left( {g_\alpha x_\alpha (0)
1578 - \frac{{g_\alpha ^2 }}{{m_\alpha \omega _\alpha ^2 }}x(0)}
1579 \right)\cos (\omega _\alpha t)} + \frac{{\dot x_\alpha
1580 (0)}}{{\omega _\alpha }}\sin (\omega _\alpha t),
1581 \label{introEquation:randomForceDefinition}
1582 \end{equation}
1583 the equation of motion can be rewritten as
1584 \begin{equation}
1585 m\ddot x = - \frac{{\partial W}}{{\partial x}} - \int_0^t {\xi
1586 (t)\dot x(t - \tau )d\tau } + R(t)
1587 \label{introEuqation:GeneralizedLangevinDynamics}
1588 \end{equation}
1589 which is known as the \emph{generalized Langevin equation} (GLE).
1590
1591 \subsubsection{\label{introSection:randomForceDynamicFrictionKernel}\textbf{Random Force and Dynamic Friction Kernel}}
1592
1593 One may notice that $R(t)$ depends only on initial conditions, which
1594 implies it is completely deterministic within the context of a
1595 harmonic bath. However, it is easy to verify that $R(t)$ is totally
1596 uncorrelated to $x$ and $\dot x$, $\left\langle {x(t)R(t)}
1597 \right\rangle = 0, \left\langle {\dot x(t)R(t)} \right\rangle =
1598 0.$ This property is what we expect from a truly random process. As
1599 long as the model chosen for $R(t)$ was a gaussian distribution in
1600 general, the stochastic nature of the GLE still remains.
1601 %dynamic friction kernel
1602 The convolution integral
1603 \[
1604 \int_0^t {\xi (t)\dot x(t - \tau )d\tau }
1605 \]
1606 depends on the entire history of the evolution of $x$, which implies
1607 that the bath retains memory of previous motions. In other words,
1608 the bath requires a finite time to respond to change in the motion
1609 of the system. For a sluggish bath which responds slowly to changes
1610 in the system coordinate, we may regard $\xi(t)$ as a constant
1611 $\xi(t) = \Xi_0$. Hence, the convolution integral becomes
1612 \[
1613 \int_0^t {\xi (t)\dot x(t - \tau )d\tau } = \xi _0 (x(t) - x(0))
1614 \]
1615 and Eq.~\ref{introEuqation:GeneralizedLangevinDynamics} becomes
1616 \[
1617 m\ddot x = - \frac{\partial }{{\partial x}}\left( {W(x) +
1618 \frac{1}{2}\xi _0 (x - x_0 )^2 } \right) + R(t),
1619 \]
1620 which can be used to describe the effect of dynamic caging in
1621 viscous solvents. The other extreme is the bath that responds
1622 infinitely quickly to motions in the system. Thus, $\xi (t)$ can be
1623 taken as a $delta$ function in time:
1624 \[
1625 \xi (t) = 2\xi _0 \delta (t).
1626 \]
1627 Hence, the convolution integral becomes
1628 \[
1629 \int_0^t {\xi (t)\dot x(t - \tau )d\tau } = 2\xi _0 \int_0^t
1630 {\delta (t)\dot x(t - \tau )d\tau } = \xi _0 \dot x(t),
1631 \]
1632 and Eq.~\ref{introEuqation:GeneralizedLangevinDynamics} becomes
1633 \begin{equation}
1634 m\ddot x = - \frac{{\partial W(x)}}{{\partial x}} - \xi _0 \dot
1635 x(t) + R(t) \label{introEquation:LangevinEquation}
1636 \end{equation}
1637 which is known as the Langevin equation. The static friction
1638 coefficient $\xi _0$ can either be calculated from spectral density
1639 or be determined by Stokes' law for regular shaped particles. A
1640 brief review on calculating friction tensors for arbitrary shaped
1641 particles is given in Sec.~\ref{introSection:frictionTensor}.
1642
1643 \subsubsection{\label{introSection:secondFluctuationDissipation}\textbf{The Second Fluctuation Dissipation Theorem}}
1644
1645 Defining a new set of coordinates
1646 \[
1647 q_\alpha (t) = x_\alpha (t) - \frac{1}{{m_\alpha \omega _\alpha
1648 ^2 }}x(0),
1649 \]
1650 we can rewrite $R(t)$ as
1651 \[
1652 R(t) = \sum\limits_{\alpha = 1}^N {g_\alpha q_\alpha (t)}.
1653 \]
1654 And since the $q$ coordinates are harmonic oscillators,
1655 \begin{eqnarray*}
1656 \left\langle {q_\alpha ^2 } \right\rangle & = & \frac{{kT}}{{m_\alpha \omega _\alpha ^2 }} \\
1657 \left\langle {q_\alpha (t)q_\alpha (0)} \right\rangle & = & \left\langle {q_\alpha ^2 (0)} \right\rangle \cos (\omega _\alpha t) \\
1658 \left\langle {q_\alpha (t)q_\beta (0)} \right\rangle & = &\delta _{\alpha \beta } \left\langle {q_\alpha (t)q_\alpha (0)} \right\rangle \\
1659 \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 } } \\
1660 & = &\sum\limits_\alpha {g_\alpha ^2 \left\langle {q_\alpha ^2 (0)} \right\rangle \cos (\omega _\alpha t)} \\
1661 & = &kT\xi (t)
1662 \end{eqnarray*}
1663 Thus, we recover the \emph{second fluctuation dissipation theorem}
1664 \begin{equation}
1665 \xi (t) = \left\langle {R(t)R(0)} \right\rangle
1666 \label{introEquation:secondFluctuationDissipation},
1667 \end{equation}
1668 which acts as a constraint on the possible ways in which one can
1669 model the random force and friction kernel.