ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/mattDisertation/Introduction.tex
Revision: 980
Committed: Sat Jan 24 03:09:15 2004 UTC (20 years, 5 months ago) by mmeineke
Content type: application/x-tex
File size: 25286 byte(s)
Log Message:
added in the last bit about symplectic integrators and the liouville formulation

File Contents

# User Rev Content
1 mmeineke 914
2    
3     \chapter{\label{chapt:intro}Introduction and Theoretical Background}
4    
5    
6    
7     \section{\label{introSec:theory}Theoretical Background}
8    
9 mmeineke 953 The techniques used in the course of this research fall under the two
10     main classes of molecular simulation: Molecular Dynamics and Monte
11     Carlo. Molecular Dynamic simulations integrate the equations of motion
12     for a given system of particles, allowing the researher to gain
13     insight into the time dependent evolution of a system. Diffusion
14     phenomena are readily studied with this simulation technique, making
15     Molecular Dynamics the main simulation technique used in this
16     research. Other aspects of the research fall under the Monte Carlo
17     class of simulations. In Monte Carlo, the configuration space
18     available to the collection of particles is sampled stochastichally,
19     or randomly. Each configuration is chosen with a given probability
20     based on the Maxwell Boltzman distribution. These types of simulations
21     are best used to probe properties of a system that are only dependent
22     only on the state of the system. Structural information about a system
23     is most readily obtained through these types of methods.
24 mmeineke 914
25 mmeineke 953 Although the two techniques employed seem dissimilar, they are both
26     linked by the overarching principles of Statistical
27     Thermodynamics. Statistical Thermodynamics governs the behavior of
28     both classes of simulations and dictates what each method can and
29     cannot do. When investigating a system, one most first analyze what
30     thermodynamic properties of the system are being probed, then chose
31     which method best suits that objective.
32 mmeineke 914
33     \subsection{\label{introSec:statThermo}Statistical Thermodynamics}
34    
35     ergodic hypothesis
36    
37     enesemble averages
38    
39     \subsection{\label{introSec:monteCarlo}Monte Carlo Simulations}
40    
41 mmeineke 953 The Monte Carlo method was developed by Metropolis and Ulam for their
42     work in fissionable material.\cite{metropolis:1949} The method is so
43 mmeineke 955 named, because it heavily uses random numbers in its
44     solution.\cite{allen87:csl} The Monte Carlo method allows for the
45     solution of integrals through the stochastic sampling of the values
46     within the integral. In the simplest case, the evaluation of an
47     integral would follow a brute force method of
48     sampling.\cite{Frenkel1996} Consider the following single dimensional
49     integral:
50     \begin{equation}
51     I = f(x)dx
52     \label{eq:MCex1}
53     \end{equation}
54     The equation can be recast as:
55     \begin{equation}
56 mmeineke 977 I = (b-a)\langle f(x) \rangle
57 mmeineke 955 \label{eq:MCex2}
58     \end{equation}
59 mmeineke 977 Where $\langle f(x) \rangle$ is the unweighted average over the interval
60 mmeineke 955 $[a,b]$. The calculation of the integral could then be solved by
61     randomly choosing points along the interval $[a,b]$ and calculating
62     the value of $f(x)$ at each point. The accumulated average would then
63     approach $I$ in the limit where the number of trials is infintely
64     large.
65 mmeineke 914
66 mmeineke 955 However, in Statistical Mechanics, one is typically interested in
67     integrals of the form:
68     \begin{equation}
69 mmeineke 977 \langle A \rangle = \frac{\int d^N \mathbf{r}~A(\mathbf{r}^N)%
70     e^{-\beta V(\mathbf{r}^N)}}%
71     {\int d^N \mathbf{r}~e^{-\beta V(\mathbf{r}^N)}}
72 mmeineke 955 \label{eq:mcEnsAvg}
73     \end{equation}
74 mmeineke 977 Where $\mathbf{r}^N$ stands for the coordinates of all $N$ particles
75     and $A$ is some observable that is only dependent on
76     position. $\langle A \rangle$ is the ensemble average of $A$ as
77     presented in Sec.~\ref{introSec:statThermo}. Because $A$ is
78     independent of momentum, the momenta contribution of the integral can
79     be factored out, leaving the configurational integral. Application of
80     the brute force method to this system would yield highly inefficient
81 mmeineke 955 results. Due to the Boltzman weighting of this integral, most random
82     configurations will have a near zero contribution to the ensemble
83     average. This is where a importance sampling comes into
84     play.\cite{allen87:csl}
85 mmeineke 914
86 mmeineke 955 Importance Sampling is a method where one selects a distribution from
87     which the random configurations are chosen in order to more
88     efficiently calculate the integral.\cite{Frenkel1996} Consider again
89     Eq.~\ref{eq:MCex1} rewritten to be:
90 mmeineke 956 \begin{equation}
91 mmeineke 977 I = \int^b_a \frac{f(x)}{\rho(x)} \rho(x) dx
92     \label{introEq:Importance1}
93 mmeineke 956 \end{equation}
94 mmeineke 977 Where $\rho(x)$ is an arbitrary probability distribution in $x$. If
95     one conducts $\tau$ trials selecting a random number, $\zeta_\tau$,
96     from the distribution $\rho(x)$ on the interval $[a,b]$, then
97     Eq.~\ref{introEq:Importance1} becomes
98 mmeineke 956 \begin{equation}
99 mmeineke 977 I= \biggl \langle \frac{f(x)}{\rho(x)} \biggr \rangle_{\text{trials}}
100     \label{introEq:Importance2}
101 mmeineke 956 \end{equation}
102 mmeineke 977 Looking at Eq.~\ref{eq:mcEnsAvg}, and realizing
103 mmeineke 956 \begin {equation}
104 mmeineke 977 \rho_{kT}(\mathbf{r}^N) =
105     \frac{e^{-\beta V(\mathbf{r}^N)}}
106     {\int d^N \mathbf{r}~e^{-\beta V(\mathbf{r}^N)}}
107     \label{introEq:MCboltzman}
108 mmeineke 956 \end{equation}
109 mmeineke 977 Where $\rho_{kT}$ is the boltzman distribution. The ensemble average
110     can be rewritten as
111 mmeineke 956 \begin{equation}
112 mmeineke 977 \langle A \rangle = \int d^N \mathbf{r}~A(\mathbf{r}^N)
113     \rho_{kT}(\mathbf{r}^N)
114     \label{introEq:Importance3}
115 mmeineke 956 \end{equation}
116 mmeineke 977 Applying Eq.~\ref{introEq:Importance1} one obtains
117 mmeineke 956 \begin{equation}
118 mmeineke 977 \langle A \rangle = \biggl \langle
119     \frac{ A \rho_{kT}(\mathbf{r}^N) }
120     {\rho(\mathbf{r}^N)} \biggr \rangle_{\text{trials}}
121     \label{introEq:Importance4}
122 mmeineke 956 \end{equation}
123 mmeineke 977 By selecting $\rho(\mathbf{r}^N)$ to be $\rho_{kT}(\mathbf{r}^N)$
124     Eq.~\ref{introEq:Importance4} becomes
125 mmeineke 956 \begin{equation}
126 mmeineke 977 \langle A \rangle = \langle A(\mathbf{r}^N) \rangle_{\text{trials}}
127     \label{introEq:Importance5}
128 mmeineke 956 \end{equation}
129 mmeineke 977 The difficulty is selecting points $\mathbf{r}^N$ such that they are
130     sampled from the distribution $\rho_{kT}(\mathbf{r}^N)$. A solution
131     was proposed by Metropolis et al.\cite{metropolis:1953} which involved
132     the use of a Markov chain whose limiting distribution was
133     $\rho_{kT}(\mathbf{r}^N)$.
134 mmeineke 955
135 mmeineke 977 \subsubsection{\label{introSec:markovChains}Markov Chains}
136 mmeineke 955
137 mmeineke 956 A Markov chain is a chain of states satisfying the following
138 mmeineke 977 conditions:\cite{leach01:mm}
139     \begin{enumerate}
140 mmeineke 956 \item The outcome of each trial depends only on the outcome of the previous trial.
141     \item Each trial belongs to a finite set of outcomes called the state space.
142 mmeineke 977 \end{enumerate}
143     If given two configuartions, $\mathbf{r}^N_m$ and $\mathbf{r}^N_n$,
144     $\rho_m$ and $\rho_n$ are the probablilities of being in state
145     $\mathbf{r}^N_m$ and $\mathbf{r}^N_n$ respectively. Further, the two
146     states are linked by a transition probability, $\pi_{mn}$, which is the
147     probability of going from state $m$ to state $n$.
148 mmeineke 955
149 mmeineke 977 \newcommand{\accMe}{\operatorname{acc}}
150    
151 mmeineke 956 The transition probability is given by the following:
152     \begin{equation}
153 mmeineke 977 \pi_{mn} = \alpha_{mn} \times \accMe(m \rightarrow n)
154     \label{introEq:MCpi}
155 mmeineke 956 \end{equation}
156 mmeineke 977 Where $\alpha_{mn}$ is the probability of attempting the move $m
157     \rightarrow n$, and $\accMe$ is the probability of accepting the move
158     $m \rightarrow n$. Defining a probability vector,
159     $\boldsymbol{\rho}$, such that
160 mmeineke 956 \begin{equation}
161 mmeineke 977 \boldsymbol{\rho} = \{\rho_1, \rho_2, \ldots \rho_m, \rho_n,
162     \ldots \rho_N \}
163     \label{introEq:MCrhoVector}
164 mmeineke 956 \end{equation}
165 mmeineke 977 a transition matrix $\boldsymbol{\Pi}$ can be defined,
166     whose elements are $\pi_{mn}$, for each given transition. The
167     limiting distribution of the Markov chain can then be found by
168     applying the transition matrix an infinite number of times to the
169     distribution vector.
170 mmeineke 956 \begin{equation}
171 mmeineke 977 \boldsymbol{\rho}_{\text{limit}} =
172     \lim_{N \rightarrow \infty} \boldsymbol{\rho}_{\text{initial}}
173     \boldsymbol{\Pi}^N
174     \label{introEq:MCmarkovLimit}
175 mmeineke 956 \end{equation}
176     The limiting distribution of the chain is independent of the starting
177     distribution, and successive applications of the transition matrix
178     will only yield the limiting distribution again.
179     \begin{equation}
180 mmeineke 977 \boldsymbol{\rho}_{\text{limit}} = \boldsymbol{\rho}_{\text{initial}}
181     \boldsymbol{\Pi}
182     \label{introEq:MCmarkovEquil}
183 mmeineke 956 \end{equation}
184    
185 mmeineke 977 \subsubsection{\label{introSec:metropolisMethod}The Metropolis Method}
186 mmeineke 956
187 mmeineke 977 In the Metropolis method\cite{metropolis:1953}
188     Eq.~\ref{introEq:MCmarkovEquil} is solved such that
189     $\boldsymbol{\rho}_{\text{limit}}$ matches the Boltzman distribution
190     of states. The method accomplishes this by imposing the strong
191     condition of microscopic reversibility on the equilibrium
192     distribution. Meaning, that at equilibrium the probability of going
193     from $m$ to $n$ is the same as going from $n$ to $m$.
194 mmeineke 956 \begin{equation}
195 mmeineke 977 \rho_m\pi_{mn} = \rho_n\pi_{nm}
196     \label{introEq:MCmicroReverse}
197 mmeineke 956 \end{equation}
198 mmeineke 977 Further, $\boldsymbol{\alpha}$ is chosen to be a symetric matrix in
199     the Metropolis method. Using Eq.~\ref{introEq:MCpi},
200     Eq.~\ref{introEq:MCmicroReverse} becomes
201 mmeineke 956 \begin{equation}
202 mmeineke 977 \frac{\accMe(m \rightarrow n)}{\accMe(n \rightarrow m)} =
203     \frac{\rho_n}{\rho_m}
204     \label{introEq:MCmicro2}
205 mmeineke 956 \end{equation}
206 mmeineke 977 For a Boltxman limiting distribution,
207 mmeineke 956 \begin{equation}
208 mmeineke 977 \frac{\rho_n}{\rho_m} = e^{-\beta[\mathcal{U}(n) - \mathcal{U}(m)]}
209     = e^{-\beta \Delta \mathcal{U}}
210     \label{introEq:MCmicro3}
211 mmeineke 956 \end{equation}
212     This allows for the following set of acceptance rules be defined:
213     \begin{equation}
214     EQ Here
215     \end{equation}
216    
217     Using the acceptance criteria from Eq.~\ref{fix} the Metropolis method
218     proceeds as follows
219     \begin{itemize}
220     \item Generate an initial configuration $fix$ which has some finite probability in $fix$.
221     \item Modify $fix$, to generate configuratioon $fix$.
222     \item If configuration $n$ lowers the energy of the system, accept the move with unity ($fix$ becomes $fix$). Otherwise accept with probability $fix$.
223     \item Accumulate the average for the configurational observable of intereest.
224     \item Repeat from step 2 until average converges.
225     \end{itemize}
226     One important note is that the average is accumulated whether the move
227     is accepted or not, this ensures proper weighting of the average.
228     Using Eq.~\ref{fix} it becomes clear that the accumulated averages are
229     the ensemble averages, as this method ensures that the limiting
230     distribution is the Boltzman distribution.
231    
232 mmeineke 977 \subsection{\label{introSec:MD}Molecular Dynamics Simulations}
233 mmeineke 914
234 mmeineke 956 The main simulation tool used in this research is Molecular Dynamics.
235     Molecular Dynamics is when the equations of motion for a system are
236     integrated in order to obtain information about both the positions and
237     momentum of a system, allowing the calculation of not only
238     configurational observables, but momenta dependent ones as well:
239     diffusion constants, velocity auto correlations, folding/unfolding
240     events, etc. Due to the principle of ergodicity, Eq.~\ref{fix}, the
241     average of these observables over the time period of the simulation
242     are taken to be the ensemble averages for the system.
243 mmeineke 914
244 mmeineke 956 The choice of when to use molecular dynamics over Monte Carlo
245     techniques, is normally decided by the observables in which the
246     researcher is interested. If the observabvles depend on momenta in
247     any fashion, then the only choice is molecular dynamics in some form.
248     However, when the observable is dependent only on the configuration,
249     then most of the time Monte Carlo techniques will be more efficent.
250 mmeineke 914
251 mmeineke 956 The focus of research in the second half of this dissertation is
252     centered around the dynamic properties of phospholipid bilayers,
253     making molecular dynamics key in the simulation of those properties.
254 mmeineke 914
255 mmeineke 977 \subsubsection{Molecular dynamics Algorithm}
256 mmeineke 914
257 mmeineke 956 To illustrate how the molecular dynamics technique is applied, the
258     following sections will describe the sequence involved in a
259     simulation. Sec.~\ref{fix} deals with the initialization of a
260     simulation. Sec.~\ref{fix} discusses issues involved with the
261     calculation of the forces. Sec.~\ref{fix} concludes the algorithm
262     discussion with the integration of the equations of motion. \cite{fix}
263 mmeineke 914
264 mmeineke 977 \subsubsection{initialization}
265 mmeineke 914
266 mmeineke 956 When selecting the initial configuration for the simulation it is
267     important to consider what dynamics one is hoping to observe.
268     Ch.~\ref{fix} deals with the formation and equilibrium dynamics of
269     phospholipid membranes. Therefore in these simulations initial
270     positions were selected that in some cases dispersed the lipids in
271     water, and in other cases structured the lipids into preformed
272     bilayers. Important considerations at this stage of the simulation are:
273     \begin{itemize}
274     \item There are no major overlaps of molecular or atomic orbitals
275     \item Velocities are chosen in such a way as to not gie the system a non=zero total momentum or angular momentum.
276     \item It is also sometimes desireable to select the velocities to correctly sample the target temperature.
277     \end{itemize}
278    
279     The first point is important due to the amount of potential energy
280     generated by having two particles too close together. If overlap
281     occurs, the first evaluation of forces will return numbers so large as
282     to render the numerical integration of teh motion meaningless. The
283     second consideration keeps the system from drifting or rotating as a
284     whole. This arises from the fact that most simulations are of systems
285     in equilibrium in the absence of outside forces. Therefore any net
286     movement would be unphysical and an artifact of the simulation method
287     used. The final point addresses teh selection of the magnitude of the
288     initial velocities. For many simulations it is convienient to use
289     this opportunity to scale the amount of kinetic energy to reflect the
290     desired thermal distribution of the system. However, it must be noted
291     that most systems will require further velocity rescaling after the
292     first few initial simulation steps due to either loss or gain of
293     kinetic energy from energy stored in potential degrees of freedom.
294    
295 mmeineke 977 \subsubsection{Force Evaluation}
296 mmeineke 956
297     The evaluation of forces is the most computationally expensive portion
298     of a given molecular dynamics simulation. This is due entirely to the
299     evaluation of long range forces in a simulation, typically pair-wise.
300     These forces are most commonly the Van der Waals force, and sometimes
301     Coulombic forces as well. For a pair-wise force, there are $fix$
302     pairs to be evaluated, where $n$ is the number of particles in the
303     system. This leads to the calculations scaling as $fix$, making large
304     simulations prohibitive in the absence of any computation saving
305     techniques.
306    
307     Another consideration one must resolve, is that in a given simulation
308     a disproportionate number of the particles will feel the effects of
309     the surface. \cite{fix} For a cubic system of 1000 particles arranged
310     in a $10x10x10$ cube, 488 particles will be exposed to the surface.
311     Unless one is simulating an isolated particle group in a vacuum, the
312     behavior of the system will be far from the desired bulk
313     charecteristics. To offset this, simulations employ the use of
314     periodic boundary images. \cite{fix}
315    
316     The technique involves the use of an algorithm that replicates the
317     simulation box on an infinite lattice in cartesian space. Any given
318     particle leaving the simulation box on one side will have an image of
319     itself enter on the opposite side (see Fig.~\ref{fix}).
320     \begin{equation}
321     EQ Here
322     \end{equation}
323     In addition, this sets that any given particle pair has an image, real
324     or periodic, within $fix$ of each other. A discussion of the method
325     used to calculate the periodic image can be found in Sec.\ref{fix}.
326    
327     Returning to the topic of the computational scale of the force
328     evaluation, the use of periodic boundary conditions requires that a
329     cutoff radius be employed. Using a cutoff radius improves the
330     efficiency of the force evaluation, as particles farther than a
331     predetermined distance, $fix$, are not included in the
332     calculation. \cite{fix} In a simultation with periodic images, $fix$
333     has a maximum value of $fix$. Fig.~\ref{fix} illustrates how using an
334     $fix$ larger than this value, or in the extreme limit of no $fix$ at
335     all, the corners of the simulation box are unequally weighted due to
336     the lack of particle images in the $x$, $y$, or $z$ directions past a
337     disance of $fix$.
338    
339 mmeineke 978 With the use of an $fix$, however, comes a discontinuity in the
340     potential energy curve (Fig.~\ref{fix}). To fix this discontinuity,
341     one calculates the potential energy at the $r_{\text{cut}}$, and add
342     that value to the potential. This causes the function to go smoothly
343     to zero at the cutoff radius. This ensures conservation of energy
344     when integrating the Newtonian equations of motion.
345 mmeineke 956
346 mmeineke 978 The second main simplification used in this research is the Verlet
347     neighbor list. \cite{allen87:csl} In the Verlet method, one generates
348     a list of all neighbor atoms, $j$, surrounding atom $i$ within some
349     cutoff $r_{\text{list}}$, where $r_{\text{list}}>r_{\text{cut}}$.
350     This list is created the first time forces are evaluated, then on
351     subsequent force evaluations, pair calculations are only calculated
352     from the neighbor lists. The lists are updated if any given particle
353     in the system moves farther than $r_{\text{list}}-r_{\text{cut}}$,
354     giving rise to the possibility that a particle has left or joined a
355     neighbor list.
356 mmeineke 956
357 mmeineke 978 \subsection{\label{introSec:MDintegrate} Integration of the equations of motion}
358    
359     A starting point for the discussion of molecular dynamics integrators
360     is the Verlet algorithm. \cite{Frenkel1996} It begins with a Taylor
361     expansion of position in time:
362     \begin{equation}
363     eq here
364     \label{introEq:verletForward}
365     \end{equation}
366     As well as,
367     \begin{equation}
368     eq here
369     \label{introEq:verletBack}
370     \end{equation}
371     Adding together Eq.~\ref{introEq:verletForward} and
372     Eq.~\ref{introEq:verletBack} results in,
373     \begin{equation}
374     eq here
375     \label{introEq:verletSum}
376     \end{equation}
377     Or equivalently,
378     \begin{equation}
379     eq here
380     \label{introEq:verletFinal}
381     \end{equation}
382     Which contains an error in the estimate of the new positions on the
383     order of $\Delta t^4$.
384    
385     In practice, however, the simulations in this research were integrated
386     with a velocity reformulation of teh Verlet method. \cite{allen87:csl}
387     \begin{equation}
388     eq here
389     \label{introEq:MDvelVerletPos}
390     \end{equation}
391     \begin{equation}
392     eq here
393     \label{introEq:MDvelVerletVel}
394     \end{equation}
395     The original Verlet algorithm can be regained by substituting the
396     velocity back into Eq.~\ref{introEq:MDvelVerletPos}. The Verlet
397     formulations are chosen in this research because the algorithms have
398     very little long term drift in energy conservation. Energy
399     conservation in a molecular dynamics simulation is of extreme
400     importance, as it is a measure of how closely one is following the
401     ``true'' trajectory wtih the finite integration scheme. An exact
402     solution to the integration will conserve area in phase space, as well
403     as be reversible in time, that is, the trajectory integrated forward
404     or backwards will exactly match itself. Having a finite algorithm
405     that both conserves area in phase space and is time reversible,
406     therefore increases, but does not guarantee the ``correctness'' or the
407     integrated trajectory.
408    
409     It can be shown, \cite{Frenkel1996} that although the Verlet algorithm
410     does not rigorously preserve the actual Hamiltonian, it does preserve
411     a pseudo-Hamiltonian which shadows the real one in phase space. This
412     pseudo-Hamiltonian is proveably area-conserving as well as time
413     reversible. The fact that it shadows the true Hamiltonian in phase
414     space is acceptable in actual simulations as one is interested in the
415     ensemble average of the observable being measured. From the ergodic
416     hypothesis (Sec.~\ref{introSec:StatThermo}), it is known that the time
417     average will match the ensemble average, therefore two similar
418     trajectories in phase space should give matching statistical averages.
419    
420 mmeineke 979 \subsection{\label{introSec:MDfurther}Further Considerations}
421 mmeineke 978 In the simulations presented in this research, a few additional
422     parameters are needed to describe the motions. The simulations
423     involving water and phospholipids in Chapt.~\ref{chaptLipids} are
424     required to integrate the equations of motions for dipoles on atoms.
425     This involves an additional three parameters be specified for each
426     dipole atom: $\phi$, $\theta$, and $\psi$. These three angles are
427     taken to be the Euler angles, where $\phi$ is a rotation about the
428     $z$-axis, and $\theta$ is a rotation about the new $x$-axis, and
429     $\psi$ is a final rotation about the new $z$-axis (see
430     Fig.~\ref{introFig:euleerAngles}). This sequence of rotations can be
431 mmeineke 979 accumulated into a single $3 \times 3$ matrix $\mathbf{A}$
432 mmeineke 978 defined as follows:
433     \begin{equation}
434     eq here
435     \label{introEq:EulerRotMat}
436     \end{equation}
437    
438     The equations of motion for Euler angles can be written down as
439     \cite{allen87:csl}
440     \begin{equation}
441     eq here
442     \label{introEq:MDeuleeerPsi}
443     \end{equation}
444     Where $\omega^s_i$ is the angular velocity in the lab space frame
445     along cartesian coordinate $i$. However, a difficulty arises when
446 mmeineke 979 attempting to integrate Eq.~\ref{introEq:MDeulerPhi} and
447 mmeineke 978 Eq.~\ref{introEq:MDeulerPsi}. The $\frac{1}{\sin \theta}$ present in
448     both equations means there is a non-physical instability present when
449     $\theta$ is 0 or $\pi$.
450    
451     To correct for this, the simulations integrate the rotation matrix,
452 mmeineke 979 $\mathbf{A}$, directly, thus avoiding the instability.
453 mmeineke 978 This method was proposed by Dullwebber
454     \emph{et. al.}\cite{Dullwebber:1997}, and is presented in
455     Sec.~\ref{introSec:MDsymplecticRot}.
456    
457     \subsubsection{\label{introSec:MDliouville}Liouville Propagator}
458    
459 mmeineke 980 Before discussing the integration of the rotation matrix, it is
460     necessary to understand the construction of a ``good'' integration
461     scheme. It has been previously
462     discussed(Sec.~\ref{introSec:MDintegrate}) how it is desirable for an
463     integrator to be symplectic, or time reversible. The following is an
464     outline of the Trotter factorization of the Liouville Propagator as a
465     scheme for generating symplectic integrators. \cite{Tuckerman:1992}
466 mmeineke 978
467 mmeineke 980 For a system with $f$ degrees of freedom the Liouville operator can be
468     defined as,
469     \begin{equation}
470     eq here
471     \label{introEq:LiouvilleOperator}
472     \end{equation}
473     Here, $r_j$ and $p_j$ are the position and conjugate momenta of a
474     degree of freedom, and $f_j$ is the force on that degree of freedom.
475     $\Gamma$ is defined as the set of all positions nad conjugate momenta,
476     $\{r_j,p_j\}$, and the propagator, $U(t)$, is defined
477     \begin {equation}
478     eq here
479     \label{introEq:Lpropagator}
480     \end{equation}
481     This allows the specification of $\Gamma$ at any time $t$ as
482     \begin{equation}
483     eq here
484     \label{introEq:Lp2}
485     \end{equation}
486     It is important to note, $U(t)$ is a unitary operator meaning
487     \begin{equation}
488     U(-t)=U^{-1}(t)
489     \label{introEq:Lp3}
490     \end{equation}
491    
492     Decomposing $L$ into two parts, $iL_1$ and $iL_2$, one can use the
493     Trotter theorem to yield
494     \begin{equation}
495     eq here
496     \label{introEq:Lp4}
497     \end{equation}
498     Where $\Delta t = \frac{t}{P}$.
499     With this, a discrete time operator $G(\Delta t)$ can be defined:
500     \begin{equation}
501     eq here
502     \label{introEq:Lp5}
503     \end{equation}
504     Because $U_1(t)$ and $U_2(t)$ are unitary, $G|\Delta t)$ is also
505     unitary. Meaning an integrator based on this factorization will be
506     reversible in time.
507    
508     As an example, consider the following decomposition of $L$:
509     \begin{equation}
510     eq here
511     \label{introEq:Lp6}
512     \end{equation}
513     Operating $G(\Delta t)$ on $\Gamma)0)$, and utilizing the operator property
514     \begin{equation}
515     eq here
516     \label{introEq:Lp8}
517     \end{equation}
518     Where $c$ is independent of $q$. One obtains the following:
519     \begin{equation}
520     eq here
521     \label{introEq:Lp8}
522     \end{equation}
523     Or written another way,
524     \begin{equation}
525     eq here
526     \label{intorEq:Lp9}
527     \end{equation}
528     This is the velocity Verlet formulation presented in
529     Sec.~\ref{introSec:MDintegrate}. Because this integration scheme is
530     comprised of unitary propagators, it is symplectic, and therefore area
531     preserving in phase space. From the preceeding fatorization, one can
532     see that the integration of the equations of motion would follow:
533     \begin{enumerate}
534     \item calculate the velocities at the half step, $\frac{\Delta t}{2}$, from the forces calculated at the initial position.
535    
536     \item Use the half step velocities to move positions one whole step, $\Delta t$.
537    
538     \item Evaluate the forces at the new positions, $\mathbf{r}(\Delta t)$, and use the new forces to complete the velocity move.
539    
540     \item Repeat from step 1 with the new position, velocities, and forces assuming the roles of the initial values.
541     \end{enumerate}
542    
543     \subsubsection{\label{introSec:MDsymplecticRot} Symplectic Propagation of the Rotation Matrix}
544    
545     Based on the factorization from the previous section,
546     Dullweber\emph{et al.}\cite{Dullweber:1997}~ proposed a scheme for the
547     symplectic propagation of the rotation matrix, $\mathbf{A}$, as an
548     alternative method for the integration of orientational degrees of
549     freedom. The method starts with a straightforward splitting of the
550     Liouville operator:
551     \begin{equation}
552     eq here
553     \label{introEq:SR1}
554     \end{equation}
555     Where $\boldsymbol{\tau}(\mathbf{A})$ are the tourques of the system
556     due to the configuration, and $\boldsymbol{/pi}$ are the conjugate
557     angular momenta of the system. The propagator, $G(\Delta t)$, becomes
558     \begin{equation}
559     eq here
560     \label{introEq:SR2}
561     \end{equation}
562     Propagation fo the linear and angular momenta follows as in the Verlet
563     scheme. The propagation of positions also follows the verlet scheme
564     with the addition of a further symplectic splitting of the rotation
565     matrix propagation, $\mathcal{G}_{\text{rot}}(\Delta t)$.
566     \begin{equation}
567     eq here
568     \label{introEq:SR3}
569     \end{equation}
570     Where $\mathcal{G}_j$ is a unitary rotation of $\mathbf{A}$ and
571     $\boldsymbol{\pi}$ about each axis $j$. As all propagations are now
572     unitary and symplectic, the entire integration scheme is also
573     symplectic and time reversible.
574    
575 mmeineke 914 \section{\label{introSec:chapterLayout}Chapter Layout}
576    
577     \subsection{\label{introSec:RSA}Random Sequential Adsorption}
578    
579     \subsection{\label{introSec:OOPSE}The OOPSE Simulation Package}
580    
581 mmeineke 978 \subsection{\label{introSec:bilayers}A Mesoscale Model for
582     Phospholipid Bilayers}