ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/mattDisertation/Introduction.tex
Revision: 1105
Committed: Tue Apr 13 18:32:53 2004 UTC (20 years, 2 months ago) by mmeineke
Content type: application/x-tex
File size: 44349 byte(s)
Log Message:
started on some revisions

File Contents

# Content
1
2
3 \chapter{\label{chapt:intro}INTRODUCTION AND THEORETICAL BACKGROUND}
4
5
6 The techniques used in the course of this research fall under the two
7 main classes of molecular simulation: Molecular Dynamics and Monte
8 Carlo. Molecular Dynamics simulations are carried out by integrating
9 the equations of motion for a given system of particles, allowing the
10 researcher to gain insight into the time dependent evolution of a
11 system. Transport phenomena are readily studied with this simulation
12 technique, making Molecular Dynamics the main simulation technique
13 used in this research. Other aspects of the research fall in the
14 Monte Carlo class of simulations. In Monte Carlo, the configuration
15 space available to the collection of particles is sampled
16 stochastically, or randomly. Each configuration is chosen with a given
17 probability based on the Boltzmann distribution. These types
18 of simulations are best used to probe properties of a system that are
19 dependent only on the state of the system. Structural information
20 about a system is most readily obtained through these types of
21 methods.
22
23 Although the two techniques employed seem dissimilar, they are both
24 linked by the over-arching principles of Statistical
25 Mechanics. Statistical Mechanics governs the behavior of
26 both classes of simulations and dictates what each method can and
27 cannot do. When investigating a system, one must first analyze what
28 thermodynamic properties of the system are being probed, then choose
29 which method best suits that objective.
30
31 \section{\label{introSec:statThermo}Statistical Mechanics}
32
33 The following section serves as a brief introduction to some of the
34 Statistical Mechanics concepts presented in this dissertation. What
35 follows is a brief derivation of Boltzmann weighted statistics and an
36 explanation of how one can use the information to calculate an
37 observable for a system. This section then concludes with a brief
38 discussion of the ergodic hypothesis and its relevance to this
39 research.
40
41 \subsection{\label{introSec:boltzman}Boltzmann weighted statistics}
42
43 Consider a system, $\gamma$, with total energy $E_{\gamma}$. Let
44 $\Omega(E_{\gamma})$ represent the number of degenerate ways
45 $\boldsymbol{\Gamma}\{r_1,r_2,\ldots r_n,p_1,p_2,\ldots p_n\}$, the
46 collection of positions and conjugate momenta of system $\gamma$, can
47 be configured to give $E_{\gamma}$. Further, if $\gamma$ is a subset
48 of a larger system, $\boldsymbol{\Lambda}\{E_1,E_2,\ldots
49 E_{\gamma},\ldots E_n\}$, the total degeneracy of the system can be
50 expressed as
51 \begin{equation}
52 \Omega(\boldsymbol{\Lambda}) = \Omega(E_1) \times \Omega(E_2) \times \ldots
53 \Omega(E_{\gamma}) \times \ldots \Omega(E_n).
54 \label{introEq:SM0.1}
55 \end{equation}
56 This multiplicative combination of degeneracies is illustrated in
57 Fig.~\ref{introFig:degenProd}.
58
59 \begin{figure}
60 \centering
61 \includegraphics[width=\linewidth]{omegaFig.eps}
62 \caption[An explanation of the combination of degeneracies]{Systems A and B both have three energy levels and two indistinguishable particles. When the total energy is 2, there are two ways for each system to disperse the energy. However, for system C, the superset of A and B, the total degeneracy is the product of the degeneracy of each system. In this case $\Omega(\text{C})$ is 4.}
63 \label{introFig:degenProd}
64 \end{figure}
65
66 Next, consider the specific case of $\gamma$ in contact with a
67 bath. Exchange of energy is allowed between the bath and the system,
68 subject to the constraint that the total energy
69 ($E_{\text{bath}}+E_{\gamma}$) remain constant. $\Omega(E)$, where $E$
70 is the total energy of both systems, can be represented as
71 \begin{equation}
72 \Omega(E) = \Omega(E_{\gamma}) \times \Omega(E - E_{\gamma}).
73 \label{introEq:SM1}
74 \end{equation}
75 Or additively as
76 \begin{equation}
77 \ln \Omega(E) = \ln \Omega(E_{\gamma}) + \ln \Omega(E - E_{\gamma}).
78 \label{introEq:SM2}
79 \end{equation}
80
81 The solution to Eq.~\ref{introEq:SM2} maximizes the number of
82 degenerate configurations in $E$. \cite{Frenkel1996}
83 This gives
84 \begin{equation}
85 \frac{\partial \ln \Omega(E)}{\partial E_{\gamma}} = 0 =
86 \frac{\partial \ln \Omega(E_{\gamma})}{\partial E_{\gamma}}
87 +
88 \frac{\partial \ln \Omega(E_{\text{bath}})}{\partial E_{\text{bath}}}
89 \frac{\partial E_{\text{bath}}}{\partial E_{\gamma}},
90 \label{introEq:SM3}
91 \end{equation}
92 where $E_{\text{bath}}$ is $E-E_{\gamma}$, and
93 $\frac{\partial E_{\text{bath}}}{\partial E_{\gamma}}$ is
94 $-1$. Eq.~\ref{introEq:SM3} becomes
95 \begin{equation}
96 \frac{\partial \ln \Omega(E_{\gamma})}{\partial E_{\gamma}} =
97 \frac{\partial \ln \Omega(E_{\text{bath}})}{\partial E_{\text{bath}}}.
98 \label{introEq:SM4}
99 \end{equation}
100
101 At this point, one can draw a relationship between the maximization of
102 degeneracy in Eq.~\ref{introEq:SM3} and the second law of
103 thermodynamics. Namely, that for a closed system, entropy will
104 increase for an irreversible process.\cite{chandler:1987} Here the
105 process is the partitioning of energy among the two systems. This
106 allows the following definition of entropy:
107 \begin{equation}
108 S = k_B \ln \Omega(E),
109 \label{introEq:SM5}
110 \end{equation}
111 where $k_B$ is the Boltzmann constant. Having defined entropy, one can
112 also define the temperature of the system using the Maxwell relation,
113 \begin{equation}
114 \frac{1}{T} = \biggl ( \frac{\partial S}{\partial E} \biggr )_{N,V}.
115 \label{introEq:SM6}
116 \end{equation}
117 The temperature in the system $\gamma$ is then
118 \begin{equation}
119 \beta( E_{\gamma} ) = \frac{1}{k_B T} =
120 \frac{\partial \ln \Omega(E_{\gamma})}{\partial E_{\gamma}}.
121 \label{introEq:SM7}
122 \end{equation}
123 Applying this to Eq.~\ref{introEq:SM4} gives the following
124 \begin{equation}
125 \beta( E_{\gamma} ) = \beta( E_{\text{bath}} ).
126 \label{introEq:SM8}
127 \end{equation}
128 Eq.~\ref{introEq:SM8} shows that the partitioning of energy between
129 the two systems is actually a process of thermal
130 equilibration.\cite{Frenkel1996}
131
132 An application of these results is to formulate the expectation value
133 of an observable, $A$, in the canonical ensemble. In the canonical
134 ensemble, the number of particles, $N$, the volume, $V$, and the
135 temperature, $T$, are all held constant while the energy, $E$, is
136 allowed to fluctuate. Returning to the previous example, the bath
137 system is now an infinitely large thermal bath, whose exchange of
138 energy with the system $\gamma$ holds the temperature constant. The
139 partitioning of energy between the bath and the system is then related
140 to the total energy of both systems and the fluctuations in
141 $E_{\gamma}$:
142 \begin{equation}
143 \Omega( E_{\gamma} ) = \Omega( E - E_{\gamma} ).
144 \label{introEq:SM9}
145 \end{equation}
146 As for the expectation value, it can be defined
147 \begin{equation}
148 \langle A \rangle =
149 \int\limits_{\boldsymbol{\Gamma}} d\boldsymbol{\Gamma}
150 P_{\gamma} A(\boldsymbol{\Gamma}),
151 \label{introEq:SM10}
152 \end{equation}
153 where $\int\limits_{\boldsymbol{\Gamma}} d\boldsymbol{\Gamma}$ denotes
154 an integration over all accessible points in phase space, $P_{\gamma}$
155 is the probability of being in a given phase state and
156 $A(\boldsymbol{\Gamma})$ is an observable that is a function of the
157 phase state.
158
159 Because entropy seeks to maximize the number of degenerate states at a
160 given energy, the probability of being in a particular state in
161 $\gamma$ will be directly proportional to the number of allowable
162 states the coupled system is able to assume. Namely,
163 \begin{equation}
164 P_{\gamma} \propto \Omega( E_{\text{bath}} ) =
165 e^{\ln \Omega( E - E_{\gamma})}
166 \label{introEq:SM11}
167 \end{equation}
168 With $E_{\gamma} \ll E$, $\ln \Omega$ can be expanded in a Taylor series:
169 \begin{equation}
170 \ln \Omega ( E - E_{\gamma}) =
171 \ln \Omega (E) -
172 E_{\gamma} \frac{\partial \ln \Omega }{\partial E}
173 + \ldots
174 \label{introEq:SM12}
175 \end{equation}
176 Higher order terms are omitted as $E$ is an infinite thermal
177 bath. Further, using Eq.~\ref{introEq:SM7}, Eq.~\ref{introEq:SM11} can
178 be rewritten:
179 \begin{equation}
180 P_{\gamma} \propto e^{-\beta E_{\gamma}}
181 \label{introEq:SM13}
182 \end{equation}
183 Where $\ln \Omega(E)$ has been factored out of the proportionality as a
184 constant. Normalizing the probability ($\int\limits_{\boldsymbol{\Gamma}}
185 d\boldsymbol{\Gamma} P_{\gamma} = 1$) gives
186 \begin{equation}
187 P_{\gamma} = \frac{e^{-\beta E_{\gamma}}}
188 {\int\limits_{\boldsymbol{\Gamma}} d\boldsymbol{\Gamma} e^{-\beta E_{\gamma}}}
189 \label{introEq:SM14}
190 \end{equation}
191 This result is the standard Boltzmann statistical distribution.
192 Applying it to Eq.~\ref{introEq:SM10} one can obtain the following relation for ensemble averages:
193 \begin{equation}
194 \langle A \rangle =
195 \frac{\int\limits_{\boldsymbol{\Gamma}} d\boldsymbol{\Gamma}
196 A( \boldsymbol{\Gamma} ) e^{-\beta E_{\gamma}}}
197 {\int\limits_{\boldsymbol{\Gamma}} d\boldsymbol{\Gamma} e^{-\beta E_{\gamma}}}
198 \label{introEq:SM15}
199 \end{equation}
200
201 \subsection{\label{introSec:ergodic}The Ergodic Hypothesis}
202
203 In the case of a Molecular Dynamics simulation, rather than
204 calculating an ensemble average integral over phase space as in
205 Eq.~\ref{introEq:SM15}, it becomes easier to caclulate the time
206 average of an observable. Namely,
207 \begin{equation}
208 \langle A \rangle_t = \frac{1}{\tau}
209 \int_0^{\tau} A[\boldsymbol{\Gamma}(t)]\,dt
210 \label{introEq:SM16}
211 \end{equation}
212 Where the value of an observable is averaged over the length of time,
213 $\tau$, that the simulation is run. This type of measurement mirrors
214 the experimental measurement of an observable. In an experiment, the
215 instrument analyzing the system must average its observation over the
216 finite time of the measurement. What is required then, is a principle
217 to relate the time average to the ensemble average. This is the
218 ergodic hypothesis.
219
220 Ergodicity is the assumption that given an infinite amount of time, a
221 system will visit every available point in phase
222 space.\cite{Frenkel1996} For most systems, this is a valid assumption,
223 except in cases where the system may be trapped in a local feature
224 (\emph{e.g.}~glasses). When valid, ergodicity allows the unification
225 of a time averaged observation and an ensemble averaged one. If an
226 observation is averaged over a sufficiently long time, the system is
227 assumed to visit all energetically accessible points in phase space,
228 giving a properly weighted statistical average. This allows the
229 researcher freedom of choice when deciding how best to measure a given
230 observable. When an ensemble averaged approach seems most logical,
231 the Monte Carlo techniques described in Sec.~\ref{introSec:monteCarlo}
232 can be utilized. Conversely, if a problem lends itself to a time
233 averaging approach, the Molecular Dynamics techniques in
234 Sec.~\ref{introSec:MD} can be employed.
235
236 \section{\label{introSec:monteCarlo}Monte Carlo Simulations}
237
238 The Monte Carlo method was developed by Metropolis and Ulam for their
239 work in fissionable material.\cite{metropolis:1949} The method is so
240 named, because it uses random numbers extensively.\cite{allen87:csl}
241 The Monte Carlo method allows for the solution of integrals through
242 the stochastic sampling of the values within the integral. In the
243 simplest case, the evaluation of an integral would follow a brute
244 force method of sampling.\cite{Frenkel1996} Consider the following
245 single dimensional integral:
246 \begin{equation}
247 I = \int_a^b f(x)dx
248 \label{eq:MCex1}
249 \end{equation}
250 The equation can be recast as:
251 \begin{equation}
252 I = \int^b_a \frac{f(x)}{\rho(x)} \rho(x) dx
253 \label{introEq:Importance1}
254 \end{equation}
255 Where $\rho(x)$ is an arbitrary probability distribution in $x$. If
256 one conducts $\tau$ trials selecting a random number, $\zeta_\tau$,
257 from the distribution $\rho(x)$ on the interval $[a,b]$, then
258 Eq.~\ref{introEq:Importance1} becomes
259 \begin{equation}
260 I= \lim_{\tau \rightarrow \infty}\biggl \langle \frac{f(x)}{\rho(x)} \biggr \rangle_{\text{trials}[a,b]}
261 \label{introEq:Importance2}
262 \end{equation}
263 If $\rho(x)$ is uniformly distributed over the interval $[a,b]$,
264 \begin{equation}
265 \rho(x) = \frac{1}{b-a}
266 \label{introEq:importance2b}
267 \end{equation}
268 then the integral can be rewritten as
269 \begin{equation}
270 I = (b-a)\lim_{\tau \rightarrow \infty}
271 \langle f(x) \rangle_{\text{trials}[a,b]}
272 \label{eq:MCex2}
273 \end{equation}
274
275 However, in Statistical Mechanics, one is typically interested in
276 integrals of the form:
277 \begin{equation}
278 \langle A \rangle = \frac{\int d^N \mathbf{r}~A(\mathbf{r}^N)%
279 e^{-\beta V(\mathbf{r}^N)}}%
280 {\int d^N \mathbf{r}~e^{-\beta V(\mathbf{r}^N)}}
281 \label{eq:mcEnsAvg}
282 \end{equation}
283 Where $\mathbf{r}^N$ stands for the coordinates of all $N$ particles
284 and $A$ is some observable that is only dependent on position. This is
285 the ensemble average of $A$ as presented in
286 Sec.~\ref{introSec:statThermo}, except here $A$ is independent of
287 momentum. Therefore the momenta contribution of the integral can be
288 factored out, leaving the configurational integral. Application of the
289 brute force method to this system would yield highly inefficient
290 results. Due to the Boltzmann weighting of this integral, most random
291 configurations will have a near zero contribution to the ensemble
292 average. This is where importance sampling comes into
293 play.\cite{allen87:csl}
294
295 Importance sampling is a method where the distribution, from which the
296 random configurations are chosen, is selected in such a way as to
297 efficiently sample the integral in question. Looking at
298 Eq.~\ref{eq:mcEnsAvg}, and realizing
299 \begin {equation}
300 \rho_{kT}(\mathbf{r}^N) =
301 \frac{e^{-\beta V(\mathbf{r}^N)}}
302 {\int d^N \mathbf{r}~e^{-\beta V(\mathbf{r}^N)}}
303 \label{introEq:MCboltzman}
304 \end{equation}
305 Where $\rho_{kT}$ is the Boltzmann distribution. The ensemble average
306 can be rewritten as
307 \begin{equation}
308 \langle A \rangle = \int d^N \mathbf{r}~A(\mathbf{r}^N)
309 \rho_{kT}(\mathbf{r}^N)
310 \label{introEq:Importance3}
311 \end{equation}
312 Applying Eq.~\ref{introEq:Importance1} one obtains
313 \begin{equation}
314 \langle A \rangle = \biggl \langle
315 \frac{ A \rho_{kT}(\mathbf{r}^N) }
316 {\rho(\mathbf{r}^N)} \biggr \rangle_{\text{trials}}
317 \label{introEq:Importance4}
318 \end{equation}
319 By selecting $\rho(\mathbf{r}^N)$ to be $\rho_{kT}(\mathbf{r}^N)$
320 Eq.~\ref{introEq:Importance4} becomes
321 \begin{equation}
322 \langle A \rangle = \langle A(\mathbf{r}^N) \rangle_{kT}
323 \label{introEq:Importance5}
324 \end{equation}
325 The difficulty is selecting points $\mathbf{r}^N$ such that they are
326 sampled from the distribution $\rho_{kT}(\mathbf{r}^N)$. A solution
327 was proposed by Metropolis \emph{et al}.\cite{metropolis:1953} which involved
328 the use of a Markov chain whose limiting distribution was
329 $\rho_{kT}(\mathbf{r}^N)$.
330
331 \subsection{\label{introSec:markovChains}Markov Chains}
332
333 A Markov chain is a chain of states satisfying the following
334 conditions:\cite{leach01:mm}
335 \begin{enumerate}
336 \item The outcome of each trial depends only on the outcome of the previous trial.
337 \item Each trial belongs to a finite set of outcomes called the state space.
338 \end{enumerate}
339 If given two configurations, $\mathbf{r}^N_m$ and $\mathbf{r}^N_n$,
340 $\rho_m$ and $\rho_n$ are the probabilities of being in state
341 $\mathbf{r}^N_m$ and $\mathbf{r}^N_n$ respectively. Further, the two
342 states are linked by a transition probability, $\pi_{mn}$, which is the
343 probability of going from state $m$ to state $n$.
344
345 \newcommand{\accMe}{\operatorname{acc}}
346
347 The transition probability is given by the following:
348 \begin{equation}
349 \pi_{mn} = \alpha_{mn} \times \accMe(m \rightarrow n)
350 \label{introEq:MCpi}
351 \end{equation}
352 Where $\alpha_{mn}$ is the probability of attempting the move $m
353 \rightarrow n$, and $\accMe$ is the probability of accepting the move
354 $m \rightarrow n$. Defining a probability vector,
355 $\boldsymbol{\rho}$, such that
356 \begin{equation}
357 \boldsymbol{\rho} = \{\rho_1, \rho_2, \ldots \rho_m, \rho_n,
358 \ldots \rho_N \}
359 \label{introEq:MCrhoVector}
360 \end{equation}
361 a transition matrix $\boldsymbol{\Pi}$ can be defined,
362 whose elements are $\pi_{mn}$, for each given transition. The
363 limiting distribution of the Markov chain can then be found by
364 applying the transition matrix an infinite number of times to the
365 distribution vector.
366 \begin{equation}
367 \boldsymbol{\rho}_{\text{limit}} =
368 \lim_{N \rightarrow \infty} \boldsymbol{\rho}_{\text{initial}}
369 \boldsymbol{\Pi}^N
370 \label{introEq:MCmarkovLimit}
371 \end{equation}
372 The limiting distribution of the chain is independent of the starting
373 distribution, and successive applications of the transition matrix
374 will only yield the limiting distribution again.
375 \begin{equation}
376 \boldsymbol{\rho}_{\text{limit}} = \boldsymbol{\rho}_{\text{limit}}
377 \boldsymbol{\Pi}
378 \label{introEq:MCmarkovEquil}
379 \end{equation}
380
381 \subsection{\label{introSec:metropolisMethod}The Metropolis Method}
382
383 In the Metropolis method\cite{metropolis:1953}
384 Eq.~\ref{introEq:MCmarkovEquil} is solved such that
385 $\boldsymbol{\rho}_{\text{limit}}$ matches the Boltzmann distribution
386 of states. The method accomplishes this by imposing the strong
387 condition of microscopic reversibility on the equilibrium
388 distribution. This means that at equilibrium, the probability of going
389 from $m$ to $n$ is the same as going from $n$ to $m$.
390 \begin{equation}
391 \rho_m\pi_{mn} = \rho_n\pi_{nm}
392 \label{introEq:MCmicroReverse}
393 \end{equation}
394 Further, $\boldsymbol{\alpha}$ is chosen to be a symmetric matrix in
395 the Metropolis method. Using Eq.~\ref{introEq:MCpi},
396 Eq.~\ref{introEq:MCmicroReverse} becomes
397 \begin{equation}
398 \frac{\accMe(m \rightarrow n)}{\accMe(n \rightarrow m)} =
399 \frac{\rho_n}{\rho_m}
400 \label{introEq:MCmicro2}
401 \end{equation}
402 For a Boltzmann limiting distribution,
403 \begin{equation}
404 \frac{\rho_n}{\rho_m} = e^{-\beta[\mathcal{U}(n) - \mathcal{U}(m)]}
405 = e^{-\beta \Delta \mathcal{U}}
406 \label{introEq:MCmicro3}
407 \end{equation}
408 This allows for the following set of acceptance rules be defined:
409 \begin{equation}
410 \accMe( m \rightarrow n ) =
411 \begin{cases}
412 1& \text{if $\Delta \mathcal{U} \leq 0$,} \\
413 e^{-\beta \Delta \mathcal{U}}& \text{if $\Delta \mathcal{U} > 0$.}
414 \end{cases}
415 \label{introEq:accRules}
416 \end{equation}
417
418 Using the acceptance criteria from Eq.~\ref{introEq:accRules} the
419 Metropolis method proceeds as follows
420 \begin{enumerate}
421 \item Generate an initial configuration $\mathbf{r}^N$ which has some finite probability in $\rho_{kT}$.
422 \item Modify $\mathbf{r}^N$, to generate configuration $\mathbf{r^{\prime}}^N$.
423 \item If the new configuration lowers the energy of the system, accept the move ($\mathbf{r}^N$ becomes $\mathbf{r^{\prime}}^N$). Otherwise accept with probability $e^{-\beta \Delta \mathcal{U}}$.
424 \item Accumulate the average for the configurational observable of interest.
425 \item Repeat from step 2 until the average converges.
426 \end{enumerate}
427 One important note is that the average is accumulated whether the move
428 is accepted or not, this ensures proper weighting of the average.
429 Using Eq.~\ref{introEq:Importance4} it becomes clear that the
430 accumulated averages are the ensemble averages, as this method ensures
431 that the limiting distribution is the Boltzmann distribution.
432
433 \section{\label{introSec:MD}Molecular Dynamics Simulations}
434
435 The main simulation tool used in this research is Molecular Dynamics.
436 Molecular Dynamics is the numerical integration of the equations of
437 motion for a system in order to obtain information about the dynamic
438 changes in the positions and momentum of the constituent particles.
439 This allows the calculation of not only configurational observables,
440 but momentum dependent ones as well: diffusion constants, relaxation
441 events, folding/unfolding events, etc. With the time dependent
442 information gained from a Molecular Dynamics simulation, one can also
443 calculate time correlation functions of the form\cite{Hansen86}
444 \begin{equation}
445 \langle A(t)\,A(0)\rangle = \lim_{\tau\rightarrow\infty} \frac{1}{\tau}
446 \int_0^{\tau} A(t+t^{\prime})\,A(t^{\prime})\,dt^{\prime}
447 \label{introEq:timeCorr}
448 \end{equation}
449 These correlations can be used to measure fundamental time constants
450 of a system, such as diffusion constants from the velocity
451 autocorrelation or dipole relaxation times from the dipole
452 autocorrelation. Due to the principle of ergodicity,
453 Sec.~\ref{introSec:ergodic}, the average of static observables over
454 the length of the simulation are taken to be the ensemble averages for
455 the system.
456
457 The choice of when to use molecular dynamics over Monte Carlo
458 techniques, is normally decided by the observables in which the
459 researcher is interested. If the observables depend on time in any
460 fashion, then the only choice is molecular dynamics in some form.
461 However, when the observable is dependent only on the configuration,
462 then for most small systems, Monte Carlo techniques will be more
463 efficient. The focus of research in the second half of this
464 dissertation is centered on the dynamic properties of phospholipid
465 bilayers, making molecular dynamics key in the simulation of those
466 properties.
467
468 \subsection{\label{introSec:mdAlgorithm}Molecular Dynamics Algorithms}
469
470 To illustrate how the molecular dynamics technique is applied, the
471 following sections will describe the sequence involved in a
472 simulation. Sec.~\ref{introSec:mdInit} deals with the initialization
473 of a simulation. Sec.~\ref{introSec:mdForce} discusses issues
474 involved with the calculation of the forces.
475 Sec.~\ref{introSec:mdIntegrate} concludes the algorithm discussion
476 with the integration of the equations of motion.\cite{Frenkel1996}
477
478 \subsection{\label{introSec:mdInit}Initialization}
479
480 When selecting the initial configuration for the simulation it is
481 important to consider what dynamics one is hoping to observe.
482 Ch.~\ref{chapt:lipid} deals with the formation and equilibrium dynamics of
483 phospholipid membranes. Therefore in these simulations initial
484 positions were selected that in some cases dispersed the lipids in
485 water, and in other cases structured the lipids into preformed
486 bilayers. Important considerations at this stage of the simulation are:
487 \begin{itemize}
488 \item There are no overlapping molecules or atoms.
489 \item Velocities are chosen in such a way as to give the system zero total momentum and angular momentum.
490 \item It is also sometimes desirable to select the velocities to correctly sample the ensemble at a particular target temperature.
491 \end{itemize}
492
493 The first point is important due to the forces generated when two
494 particles are too close together. If overlap occurs, the first
495 evaluation of forces will return numbers so large as to render the
496 numerical integration of the equations motion meaningless. The second
497 consideration keeps the system from drifting or rotating as a whole.
498 This arises from the fact that most simulations are of systems in
499 equilibrium in the absence of outside forces. Therefore any net
500 movement would be unphysical and an artifact of the simulation method
501 used. The final point addresses the selection of the magnitude of the
502 initial velocities. For many simulations it is convenient to use this
503 opportunity to scale the amount of kinetic energy to reflect the
504 desired thermal distribution of the system. However, it must be noted
505 that due to equipartition, most systems will require further velocity
506 rescaling after the first few initial simulation steps due to either
507 loss or gain of kinetic energy from energy stored as potential energy.
508
509 \subsection{\label{introSec:mdForce}Force Evaluation}
510
511 The evaluation of forces is the most computationally expensive portion
512 of any molecular dynamics simulation. This is due entirely to the
513 evaluation of long range forces in a simulation, which are typically
514 pair potentials. These forces are most commonly the van der Waals
515 force, and sometimes Coulombic forces as well. For a pair-wise
516 interaction, there are $N(N-1)/ 2$ pairs to be evaluated, where $N$ is
517 the number of particles in the system. This leads to calculations which
518 scale as $N^2$, making large simulations prohibitive in the absence
519 of any computation saving techniques.
520
521 Another consideration one must resolve, is that in a given simulation,
522 a disproportionate number of the particles will feel the effects of
523 the surface.\cite{allen87:csl} For a cubic system of 1000 particles
524 arranged in a $10 \times 10 \times 10$ cube, 488 particles will be
525 exposed to the surface. Unless one is simulating an isolated cluster
526 in a vacuum, the behavior of the system will be far from the desired
527 bulk characteristics. To offset this, simulations employ the use of
528 periodic images.\cite{born:1912}
529
530 The technique involves the use of an algorithm that replicates the
531 simulation box on an infinite lattice in Cartesian space. Any
532 particle leaving the simulation box on one side will have an image of
533 itself enter on the opposite side (see
534 Fig.~\ref{introFig:pbc}). Therefore, a pair of particles have an
535 image of each other, real or periodic, within $\text{box}/2$. A
536 discussion of the method used to calculate the periodic image can be
537 found in Sec.\ref{oopseSec:pbc}.
538
539 \begin{figure}
540 \centering
541 \includegraphics[width=\linewidth]{pbcFig.eps}
542 \caption[An illustration of periodic boundary conditions]{A 2-D illustration of periodic boundary conditions. As one particle leaves the right of the simulation box, an image of it enters the left.}
543 \label{introFig:pbc}
544 \end{figure}
545
546 Returning to the topic of the computational scaling of the force
547 evaluation, the use of periodic boundary conditions requires that a
548 cutoff radius be employed. Using a cutoff radius improves the
549 efficiency of the force evaluation, as particles farther than a
550 predetermined distance, $r_{\text{cut}}$, are not included in the
551 calculation.\cite{Frenkel1996} In a simulation with periodic images,
552 there are two methods to choose from, both with their own cutoff
553 limits. In the minimum image convention, $r_{\text{cut}}$ has a
554 maximum value of $\text{box}/2$. This is because each atom has only
555 one image that is seen by other atoms. The image used is the one that
556 minimizes the distance between the two atoms. A system of wrapped
557 images about a central atom therefore has a maximum length scale of
558 box on a side (Fig.~\ref{introFig:rMaxMin}). The second convention,
559 multiple image convention, has a maximum $r_{\text{cut}}$ of the box
560 length itself. Here multiple images of each atom are replicated in the
561 periodic cells surrounding the central atom, this causes the atom to
562 see multiple copies of several atoms. If the cutoff radius is larger
563 than box, however, then the atom will see an image of itself, and
564 attempt to calculate an unphysical self-self force interaction
565 (Fig.~\ref{introFig:rMaxMult}). Due to the increased complexity and
566 commputaional ineffeciency of the multiple image method, the minimum
567 image method is the used throughout this research.
568
569 \begin{figure}
570 \centering
571 \includegraphics[width=\linewidth]{rCutMaxFig.eps}
572 \caption[An explanation of minimum image convention]{The yellow atom has all other images wrapped to itself as the center. If $r_{\text{cut}}=\text{box}/2$, then the distribution is uniform (blue atoms). However, when $r_{\text{cut}}>\text{box}/2$ the corners are disproportionately weighted (green atoms) vs the axial directions (shaded regions).}
573 \label{introFig:rMaxMin}
574 \end{figure}
575
576 \begin{figure}
577 \centering
578 \includegraphics[width=\linewidth]{rCutMaxMultFig.eps}
579 \caption[An explanation of multiple image convention]{The yellow atom is the central wrapping point. The blue atoms are the minimum images of the system about the central atom. The boxes with the green atoms are multiple images of the central box. If $r_{\text{cut}} \geq \text{box}$ then the central atom sees multiple images of itself (red atom), creating a self-self force evaluation.}
580 \label{introFig:rMaxMult}
581 \end{figure}
582
583 With the use of a cutoff radius, however, comes a discontinuity in
584 the potential energy curve (Fig.~\ref{introFig:shiftPot}). To fix this
585 discontinuity, one calculates the potential energy at the
586 $r_{\text{cut}}$, and adds that value to the potential, causing
587 the function to go smoothly to zero at the cutoff radius. This
588 shifted potential ensures conservation of energy when integrating the
589 Newtonian equations of motion.
590
591 \begin{figure}
592 \centering
593 \includegraphics[width=\linewidth]{shiftedPot.eps}
594 \caption[Shifting the Lennard-Jones Potential]{The Lennard-Jones potential (blue line) is shifted (red line) to remove the discontinuity at $r_{\text{cut}}$.}
595 \label{introFig:shiftPot}
596 \end{figure}
597
598 The second main simplification used in this research is the Verlet
599 neighbor list.\cite{allen87:csl} In the Verlet method, one generates
600 a list of all neighbor atoms, $j$, surrounding atom $i$ within some
601 cutoff $r_{\text{list}}$, where $r_{\text{list}}>r_{\text{cut}}$.
602 This list is created the first time forces are evaluated, then on
603 subsequent force evaluations, pair calculations are only calculated
604 from the neighbor lists. The lists are updated if any particle
605 in the system moves farther than $r_{\text{list}}-r_{\text{cut}}$,
606 which indeicates the possibility that a particle has left or joined the
607 neighbor list.
608
609 \subsection{\label{introSec:mdIntegrate} Integration of the Equations of Motion}
610
611 A starting point for the discussion of molecular dynamics integrators
612 is the Verlet algorithm.\cite{Frenkel1996} It begins with a Taylor
613 expansion of position in time:
614 \begin{equation}
615 q(t+\Delta t)= q(t) + v(t)\Delta t + \frac{F(t)}{2m}\Delta t^2 +
616 \frac{\Delta t^3}{3!}\frac{\partial^3 q(t)}{\partial t^3} +
617 \mathcal{O}(\Delta t^4)
618 \label{introEq:verletForward}
619 \end{equation}
620 As well as,
621 \begin{equation}
622 q(t-\Delta t)= q(t) - v(t)\Delta t + \frac{F(t)}{2m}\Delta t^2 -
623 \frac{\Delta t^3}{3!}\frac{\partial^3 q(t)}{\partial t^3} +
624 \mathcal{O}(\Delta t^4)
625 \label{introEq:verletBack}
626 \end{equation}
627 Where $m$ is the mass of the particle, $q(t)$ is the position at time
628 $t$, $v(t)$ the velocity, and $F(t)$ the force acting on the
629 particle. Adding together Eq.~\ref{introEq:verletForward} and
630 Eq.~\ref{introEq:verletBack} results in,
631 \begin{equation}
632 q(t+\Delta t)+q(t-\Delta t) =
633 2q(t) + \frac{F(t)}{m}\Delta t^2 + \mathcal{O}(\Delta t^4)
634 \label{introEq:verletSum}
635 \end{equation}
636 Or equivalently,
637 \begin{equation}
638 q(t+\Delta t) \approx
639 2q(t) - q(t-\Delta t) + \frac{F(t)}{m}\Delta t^2
640 \label{introEq:verletFinal}
641 \end{equation}
642 Which contains an error in the estimate of the new positions on the
643 order of $\Delta t^4$.
644
645 In practice, however, the simulations in this research were integrated
646 with the velocity reformulation of the Verlet method.\cite{allen87:csl}
647 \begin{align}
648 q(t+\Delta t) &= q(t) + v(t)\Delta t + \frac{F(t)}{2m}\Delta t^2 %
649 \label{introEq:MDvelVerletPos} \\%
650 %
651 v(t+\Delta t) &= v(t) + \frac{\Delta t}{2m}[F(t) + F(t+\Delta t)] %
652 \label{introEq:MDvelVerletVel}
653 \end{align}
654 The original Verlet algorithm can be regained by substituting the
655 velocity back into Eq.~\ref{introEq:MDvelVerletPos}. The Verlet
656 formulations are chosen in this research because the algorithms have
657 very little long time drift in energy. Energy conservation in a
658 molecular dynamics simulation is of extreme importance, as it is a
659 measure of how closely one is following the ``true'' trajectory with
660 the finite integration scheme. An exact solution to the integration
661 will conserve area in phase space (i.e.~will be symplectic), as well
662 as be reversible in time, that is, the trajectory integrated forward
663 or backwards will exactly match itself. Having a finite algorithm
664 that both conserves area in phase space and is time reversible,
665 therefore increases the reliability, but does not guarantee the
666 ``correctness'' of the integrated trajectory.
667
668 It can be shown,\cite{Frenkel1996} that although the Verlet algorithm
669 does not rigorously preserve the actual Hamiltonian, it does preserve
670 a pseudo-Hamiltonian which shadows the real one in phase space. This
671 pseudo-Hamiltonian is provably area-conserving as well as time
672 reversible. The fact that it shadows the true Hamiltonian in phase
673 space is acceptable in actual simulations as one is interested in the
674 ensemble average of the observable being measured. From the ergodic
675 hypothesis (Sec.~\ref{introSec:statThermo}), it is known that the time
676 average will match the ensemble average, therefore two similar
677 trajectories in phase space should give matching statistical averages.
678
679 \subsection{\label{introSec:MDfurther}Further Considerations}
680
681 In the simulations presented in this research, a few additional
682 parameters are needed to describe the motions. In the simulations
683 involving water and phospholipids in Ch.~\ref{chapt:lipid}, we are
684 required to integrate the equations of motions for dipoles on atoms.
685 This involves an additional three parameters be specified for each
686 dipole atom: $\phi$, $\theta$, and $\psi$. These three angles are
687 taken to be the Euler angles, where $\phi$ is a rotation about the
688 $z$-axis, and $\theta$ is a rotation about the new $x$-axis, and
689 $\psi$ is a final rotation about the new $z$-axis (see
690 Fig.~\ref{introFig:eulerAngles}). This sequence of rotations can be
691 accumulated into a single $3 \times 3$ matrix, $\mathsf{A}$,
692 defined as follows:
693 \begin{equation}
694 \mathsf{A} =
695 \begin{bmatrix}
696 \cos\phi\cos\psi-\sin\phi\cos\theta\sin\psi &%
697 \sin\phi\cos\psi+\cos\phi\cos\theta\sin\psi &%
698 \sin\theta\sin\psi \\%
699 %
700 -\cos\phi\sin\psi-\sin\phi\cos\theta\cos\psi &%
701 -\sin\phi\sin\psi+\cos\phi\cos\theta\cos\psi &%
702 \sin\theta\cos\psi \\%
703 %
704 \sin\phi\sin\theta &%
705 -\cos\phi\sin\theta &%
706 \cos\theta
707 \end{bmatrix}
708 \label{introEq:EulerRotMat}
709 \end{equation}
710
711 \begin{figure}
712 \centering
713 \includegraphics[width=\linewidth]{eulerRotFig.eps}
714 \caption[Euler rotation of Cartesian coordinates]{The rotation scheme for Euler angles. First is a rotation of $\phi$ about the $z$ axis (blue rotation). Next is a rotation of $\theta$ about the new $x$ axis (green rotation). Lastly is a final rotation of $\psi$ about the new $z$ axis (red rotation).}
715 \label{introFig:eulerAngles}
716 \end{figure}
717
718 The equations of motion for Euler angles can be written down
719 as\cite{allen87:csl}
720 \begin{align}
721 \dot{\phi} &= -\omega^s_x \frac{\sin\phi\cos\theta}{\sin\theta} +
722 \omega^s_y \frac{\cos\phi\cos\theta}{\sin\theta} +
723 \omega^s_z
724 \label{introEq:MDeulerPhi} \\%
725 %
726 \dot{\theta} &= \omega^s_x \cos\phi + \omega^s_y \sin\phi
727 \label{introEq:MDeulerTheta} \\%
728 %
729 \dot{\psi} &= \omega^s_x \frac{\sin\phi}{\sin\theta} -
730 \omega^s_y \frac{\cos\phi}{\sin\theta}
731 \label{introEq:MDeulerPsi}
732 \end{align}
733 Where $\omega^s_{\alpha}$ is the angular velocity in the lab space frame
734 along Cartesian coordinate $\alpha$. However, a difficulty arises when
735 attempting to integrate Eq.~\ref{introEq:MDeulerPhi} and
736 Eq.~\ref{introEq:MDeulerPsi}. The $\frac{1}{\sin \theta}$ present in
737 both equations means there is a non-physical instability present when
738 $\theta$ is 0 or $\pi$. To correct for this, the simulations integrate
739 the rotation matrix, $\mathsf{A}$, directly, thus avoiding the
740 instability. This method was proposed by Dullweber
741 \emph{et. al.}\cite{Dullweber1997}, and is presented in
742 Sec.~\ref{introSec:MDsymplecticRot}.
743
744 \subsection{\label{introSec:MDliouville}Liouville Propagator}
745
746 Before discussing the integration of the rotation matrix, it is
747 necessary to understand the construction of a ``good'' integration
748 scheme. It has been previously discussed
749 (Sec.~\ref{introSec:mdIntegrate}) how it is desirable for an
750 integrator to be symplectic and time reversible. The following is an
751 outline of the Trotter factorization of the Liouville Propagator as a
752 scheme for generating symplectic, time-reversible
753 integrators.\cite{Tuckerman92}
754
755 For a system with $f$ degrees of freedom the Liouville operator can be
756 defined as,
757 \begin{equation}
758 iL=\sum^f_{j=1} \biggl [\dot{q}_j\frac{\partial}{\partial q_j} +
759 F_j\frac{\partial}{\partial p_j} \biggr ]
760 \label{introEq:LiouvilleOperator}
761 \end{equation}
762 Here, $q_j$ and $p_j$ are the position and conjugate momenta of a
763 degree of freedom, and $F_j$ is the force on that degree of freedom.
764 $\Gamma$ is defined as the set of all positions and conjugate momenta,
765 $\{q_j,p_j\}$, and the propagator, $U(t)$, is defined
766 \begin {equation}
767 U(t) = e^{iLt}
768 \label{introEq:Lpropagator}
769 \end{equation}
770 This allows the specification of $\Gamma$ at any time $t$ as
771 \begin{equation}
772 \Gamma(t) = U(t)\Gamma(0)
773 \label{introEq:Lp2}
774 \end{equation}
775 It is important to note, $U(t)$ is a unitary operator meaning
776 \begin{equation}
777 U(-t)=U^{-1}(t)
778 \label{introEq:Lp3}
779 \end{equation}
780
781 Decomposing $L$ into two parts, $iL_1$ and $iL_2$, one can use the
782 Trotter theorem to yield
783 \begin{align}
784 e^{iLt} &= e^{i(L_1 + L_2)t} \notag \\%
785 %
786 &= \biggl [ e^{i(L_1 +L_2)\frac{t}{P}} \biggr]^P \notag \\%
787 %
788 &= \biggl [ e^{iL_1\frac{\Delta t}{2}}\, e^{iL_2\Delta t}\,
789 e^{iL_1\frac{\Delta t}{2}} \biggr ]^P +
790 \mathcal{O}\biggl (\frac{t^3}{P^2} \biggr ) \label{introEq:Lp4}
791 \end{align}
792 Where $\Delta t = t/P$.
793 With this, a discrete time operator $G(\Delta t)$ can be defined:
794 \begin{align}
795 G(\Delta t) &= e^{iL_1\frac{\Delta t}{2}}\, e^{iL_2\Delta t}\,
796 e^{iL_1\frac{\Delta t}{2}} \notag \\%
797 %
798 &= U_1 \biggl ( \frac{\Delta t}{2} \biggr )\, U_2 ( \Delta t )\,
799 U_1 \biggl ( \frac{\Delta t}{2} \biggr )
800 \label{introEq:Lp5}
801 \end{align}
802 Because $U_1(t)$ and $U_2(t)$ are unitary, $G(\Delta t)$ is also
803 unitary. This means that an integrator based on this factorization will be
804 reversible in time.
805
806 As an example, consider the following decomposition of $L$:
807 \begin{align}
808 iL_1 &= \dot{q}\frac{\partial}{\partial q}%
809 \label{introEq:Lp6a} \\%
810 %
811 iL_2 &= F(q)\frac{\partial}{\partial p}%
812 \label{introEq:Lp6b}
813 \end{align}
814 This leads to propagator $G( \Delta t )$ as,
815 \begin{equation}
816 G(\Delta t) = e^{\frac{\Delta t}{2} F(q)\frac{\partial}{\partial p}} \,
817 e^{\Delta t\,\dot{q}\frac{\partial}{\partial q}} \,
818 e^{\frac{\Delta t}{2} F(q)\frac{\partial}{\partial p}}
819 \label{introEq:Lp7}
820 \end{equation}
821 Operating $G(\Delta t)$ on $\Gamma(0)$, and utilizing the operator property
822 \begin{equation}
823 e^{c\frac{\partial}{\partial x}}\, f(x) = f(x+c)
824 \label{introEq:Lp8}
825 \end{equation}
826 Where $c$ is independent of $x$. One obtains the following:
827 \begin{align}
828 \dot{q}\biggl (\frac{\Delta t}{2}\biggr ) &=
829 \dot{q}(0) + \frac{\Delta t}{2m}\, F[q(0)] \label{introEq:Lp9a}\\%
830 %
831 q(\Delta t) &= q(0) + \Delta t\, \dot{q}\biggl (\frac{\Delta t}{2}\biggr )%
832 \label{introEq:Lp9b}\\%
833 %
834 \dot{q}(\Delta t) &= \dot{q}\biggl (\frac{\Delta t}{2}\biggr ) +
835 \frac{\Delta t}{2m}\, F[q(0)] \label{introEq:Lp9c}
836 \end{align}
837 Or written another way,
838 \begin{align}
839 q(t+\Delta t) &= q(0) + \dot{q}(0)\Delta t +
840 \frac{F[q(0)]}{m}\frac{\Delta t^2}{2} %
841 \label{introEq:Lp10a} \\%
842 %
843 \dot{q}(\Delta t) &= \dot{q}(0) + \frac{\Delta t}{2m}
844 \biggl [F[q(0)] + F[q(\Delta t)] \biggr] %
845 \label{introEq:Lp10b}
846 \end{align}
847 This is the velocity Verlet formulation presented in
848 Sec.~\ref{introSec:mdIntegrate}. Because this integration scheme is
849 comprised of unitary propagators, it is symplectic, and therefore area
850 preserving in phase space. From the preceding factorization, one can
851 see that the integration of the equations of motion would follow:
852 \begin{enumerate}
853 \item calculate the velocities at the half step, $\frac{\Delta t}{2}$, from the forces calculated at the initial position.
854
855 \item Use the half step velocities to move positions one whole step, $\Delta t$.
856
857 \item Evaluate the forces at the new positions, $\mathbf{r}(\Delta t)$, and use the new forces to complete the velocity move.
858
859 \item Repeat from step 1 with the new position, velocities, and forces assuming the roles of the initial values.
860 \end{enumerate}
861
862 \subsection{\label{introSec:MDsymplecticRot} Symplectic Propagation of the Rotation Matrix}
863
864 Based on the factorization from the previous section,
865 Dullweber\emph{et al}.\cite{Dullweber1997}~ proposed a scheme for the
866 symplectic propagation of the rotation matrix, $\mathsf{A}$, as an
867 alternative method for the integration of orientational degrees of
868 freedom. The method starts with a straightforward splitting of the
869 Liouville operator:
870 \begin{align}
871 iL_{\text{pos}} &= \dot{q}\frac{\partial}{\partial q} +
872 \mathsf{\dot{A}}\frac{\partial}{\partial \mathsf{A}}
873 \label{introEq:SR1a} \\%
874 %
875 iL_F &= F(q)\frac{\partial}{\partial p}
876 \label{introEq:SR1b} \\%
877 iL_{\tau} &= \tau(\mathsf{A})\frac{\partial}{\partial j}
878 \label{introEq:SR1b} \\%
879 \end{align}
880 Where $\tau(\mathsf{A})$ is the torque of the system
881 due to the configuration, and $j$ is the conjugate
882 angular momenta of the system. The propagator, $G(\Delta t)$, becomes
883 \begin{equation}
884 G(\Delta t) = e^{\frac{\Delta t}{2} F(q)\frac{\partial}{\partial p}} \,
885 e^{\frac{\Delta t}{2} \tau(\mathsf{A})\frac{\partial}{\partial j}} \,
886 e^{\Delta t\,iL_{\text{pos}}} \,
887 e^{\frac{\Delta t}{2} \tau(\mathsf{A})\frac{\partial}{\partial j}} \,
888 e^{\frac{\Delta t}{2} F(q)\frac{\partial}{\partial p}}
889 \label{introEq:SR2}
890 \end{equation}
891 Propagation of the linear and angular momenta follows as in the Verlet
892 scheme. The propagation of positions also follows the Verlet scheme
893 with the addition of a further symplectic splitting of the rotation
894 matrix propagation, $\mathcal{U}_{\text{rot}}(\Delta t)$, within
895 $U_{\text{pos}}(\Delta t)$.
896 \begin{equation}
897 \mathcal{U}_{\text{rot}}(\Delta t) =
898 \mathcal{U}_x \biggl(\frac{\Delta t}{2}\biggr)\,
899 \mathcal{U}_y \biggl(\frac{\Delta t}{2}\biggr)\,
900 \mathcal{U}_z (\Delta t)\,
901 \mathcal{U}_y \biggl(\frac{\Delta t}{2}\biggr)\,
902 \mathcal{U}_x \biggl(\frac{\Delta t}{2}\biggr)\,
903 \label{introEq:SR3}
904 \end{equation}
905 Where $\mathcal{U}_{\alpha}$ is a unitary rotation of $\mathsf{A}$ and
906 $j$ about each axis $\alpha$. As all propagations are now
907 unitary and symplectic, the entire integration scheme is also
908 symplectic and time reversible.
909
910 \section{\label{introSec:layout}Dissertation Layout}
911
912 This dissertation is divided as follows: Ch.~\ref{chapt:RSA}
913 presents the random sequential adsorption simulations of related
914 pthalocyanines on a gold (111) surface. Ch.~\ref{chapt:oopse}
915 is about our molecular dynamics simulation package
916 {\sc oopse}. Ch.~\ref{chapt:lipid} regards the simulations of
917 phospholipid bilayers using a mesoscale model. And lastly,
918 Ch.~\ref{chapt:conclusion} concludes this dissertation with a
919 summary of all results. The chapters are arranged in chronological
920 order, and reflect the progression of techniques I employed during my
921 research.
922
923 The chapter concerning random sequential adsorption simulations is a
924 study in applying Statistical Mechanics simulation techniques in order
925 to obtain a simple model capable of explaining the results. My
926 advisor, Dr. Gezelter, and I were approached by a colleague,
927 Dr. Lieberman, about possible explanations for the partial coverage of
928 a gold surface by a particular compound synthesized in her group. We
929 suggested it might be due to the statistical packing fraction of disks
930 on a plane, and set about simulating this system. As the events in
931 our model were not dynamic in nature, a Monte Carlo method was
932 employed. Here, if a molecule landed on the surface without
933 overlapping with another, then its landing was accepted. However, if there
934 was overlap, the landing we rejected and a new random landing location
935 was chosen. This defined our acceptance rules and allowed us to
936 construct a Markov chain whose limiting distribution was the surface
937 coverage in which we were interested.
938
939 The following chapter, about the simulation package {\sc oopse},
940 describes in detail the large body of scientific code that had to be
941 written in order to study phospholipid bilayers. Although there are
942 pre-existing molecular dynamic simulation packages available, none
943 were capable of implementing the models we were developing. {\sc
944 oopse} is a unique package capable of not only integrating the
945 equations of motion in Cartesian space, but is also able to integrate
946 the rotational motion of rigid bodies and dipoles. It also has the
947 ability to perform calculations across distributed parallel processors
948 and contains a flexible script syntax for creating systems. {\sc
949 oopse} has become a very powerful scientific instrument for the
950 exploration of our model.
951
952 In Ch.~\ref{chapt:lipid}, utilizing {\sc oopse}, I have been
953 able to parameterize a mesoscale model for phospholipid simulations.
954 This model retains information about solvent ordering around the
955 bilayer, as well as information regarding the interaction of the
956 phospholipid head groups' dipoles with each other and the surrounding
957 solvent. These simulations give us insight into the dynamic events
958 that lead to the formation of phospholipid bilayers, as well as
959 provide the foundation for future exploration of bilayer phase
960 behavior with this model.
961
962 In the last chapter, I discuss future directions
963 for both {\sc oopse} and this mesoscale model. Additionally, I will
964 give a summary of the results found in this dissertation.
965
966