ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/xDissertation/Introduction.tex
Revision: 3374
Committed: Thu Mar 20 22:21:43 2008 UTC (17 years, 1 month ago) by xsun
Content type: application/x-tex
File size: 34376 byte(s)
Log Message:
writing.

File Contents

# Content
1 \chapter{\label{chap:intro}INTRODUCTION AND BACKGROUND}
2
3 \section{Background on the Problem\label{In:sec:pro}}
4 Phospholipid molecules are the primary topic of this dissertation
5 because of their critical role as the foundation of biological
6 membranes. The chemical structure of phospholipids includes the polar
7 head group which is due to a large charge separation between phosphate
8 and amino alcohol, and the nonpolar tails that contains fatty acid
9 chains. Depending on the alcohol which phosphate and fatty acid chains
10 are esterified to, the phospholipids are divided into
11 glycerophospholipids and sphingophospholipids.~\cite{Cevc80} The
12 chemical structures are shown in figure~\ref{Infig:lipid}.
13 \begin{figure}
14 \centering
15 \includegraphics[width=\linewidth]{./figures/inLipid.pdf}
16 \caption{The chemical structure of glycerophospholipids (left) and
17 sphingophospholipids (right).\cite{Cevc80}}
18 \label{Infig:lipid}
19 \end{figure}
20 The glycerophospholipid is the dominant phospholipid in membranes. The
21 types of glycerophospholipids depend on the X group, and the
22 chains. For example, if X is choline, the lipids are known as
23 phosphatidylcholine (PC), or if X is ethanolamine, the lipids are
24 known as phosphatidyethanolamine (PE). Table~\ref{Intab:pc} listed a
25 number types of phosphatidycholine with different fatty acids as the
26 lipid chains.
27 \begin{table*}
28 \begin{minipage}{\linewidth}
29 \begin{center}
30 \caption{A number types of phosphatidycholine.}
31 \begin{tabular}{lll}
32 \hline
33 & Fatty acid & Generic Name \\
34 \hline
35 \textcolor{red}{DMPC} & Myristic: CH$_3$(CH$_2$)$_{12}$COOH &
36 \textcolor{red}{D}i\textcolor{red}{M}yristoyl\textcolor{red}{P}hosphatidyl\textcolor{red}{C}holine \\
37 \textcolor{red}{DPPC} & Palmitic: CH$_3$(CH$_2$)$_{14}$COOH & \textcolor{red}{D}i\textcolor{red}{P}almtoyl\textcolor{red}{P}hosphatidyl\textcolor{red}{C}holine
38 \\
39 \textcolor{red}{DSPC} & Stearic: CH$_3$(CH$_2$)$_{16}$COOH & \textcolor{red}{D}i\textcolor{red}{S}tearoyl\textcolor{red}{P}hosphatidyl\textcolor{red}{C}holine \\
40 \end{tabular}
41 \label{Intab:pc}
42 \end{center}
43 \end{minipage}
44 \end{table*}
45 When dispersed in water, lipids self assemble into a mumber of
46 topologically distinct bilayer structures. The phase behavior of lipid
47 bilayers has been explored experimentally~\cite{Cevc80}, however, a
48 complete understanding of the mechanism and driving forces behind the
49 various phases has not been achieved.
50
51 \subsection{Ripple Phase\label{In:ssec:ripple}}
52 The $P_{\beta'}$ {\it ripple phase} of lipid bilayers, named from the
53 periodic buckling of the membrane, is an intermediate phase which is
54 developed either from heating the gel phase $L_{\beta'}$ or cooling
55 the fluid phase $L_\alpha$. A Sketch is shown in
56 figure~\ref{Infig:phaseDiagram}.~\cite{Cevc80}
57 \begin{figure}
58 \centering
59 \includegraphics[width=\linewidth]{./figures/inPhaseDiagram.pdf}
60 \caption{A phase diagram of lipid bilayer. With increasing the
61 temperature, the bilayer can go through a gel ($L_{\beta'}$), ripple
62 ($P_{\beta'}$) to fluid ($L_\alpha$) phase transition.~\cite{Cevc80}}
63 \label{Infig:phaseDiagram}
64 \end{figure}
65 Most structural information of the ripple phase has been obtained by
66 the X-ray diffraction~\cite{Sun96,Katsaras00} and freeze-fracture
67 electron microscopy (FFEM).~\cite{Copeland80,Meyer96} The X-ray
68 diffraction work by Katsaras {\it et al.} showed that a rich phase
69 diagram exhibiting both {\it asymmetric} and {\it symmetric} ripples
70 is possible for lecithin bilayers.\cite{Katsaras00} Recently,
71 Kaasgaard {\it et al.} used atomic force microscopy (AFM) to observe
72 ripple phase morphology in bilayers supported on
73 mica.~\cite{Kaasgaard03}
74 \begin{figure}
75 \centering
76 \includegraphics[width=\linewidth]{./figures/inRipple.pdf}
77 \caption{The experimental observed ripple phase. The top image is
78 obtained by Sun {\it et al.} using X-ray diffraction~\cite{Sun96},
79 and the bottom one is observed by Kaasgaard {\it et al.} using
80 AFM.~\cite{Kaasgaard03}}
81 \label{Infig:ripple}
82 \end{figure}
83 Figure~\ref{Infig:ripple} shows the ripple phase oberved by X-ray
84 diffraction and AFM. The experimental results provide strong support
85 for a 2-dimensional triangular packing lattice of the lipid molecules
86 within the ripple phase. This is a notable change from the observed
87 lipid packing within the gel phase,~\cite{Cevc80} although Tenchov
88 {\it et al.} have recently observed near-hexagonal packing in some
89 phosphatidylcholine (PC) gel phases.~\cite{Tenchov2001} However, the
90 physical mechanism for the formation of the ripple phase has never
91 been explained and the microscopic structure of the ripple phase has
92 never been elucidated by experiments. Computational simulation is a
93 perfect tool to study the microscopic properties for a
94 system. However, the large length scale the ripple structure and the
95 long time scale of the formation of the ripples are crucial obstacles
96 to performing the actual work. The principal ideas explored in this
97 dissertation are attempts to break the computational task up by
98 \begin{itemize}
99 \item Simplifying the lipid model.
100 \item Improving algorithm for integrating the equations of motion.
101 \end{itemize}
102 In chapters~\ref{chap:mc} and~\ref{chap:md}, we use a simple point
103 dipole spin model and a coarse-grained molecualr scale model to
104 perform the Monte Carlo and Molecular Dynamics simulations
105 respectively, and in chapter~\ref{chap:ld}, we develop a Langevin
106 Dynamics algorithm which excludes the explicit solvent to improve the
107 efficiency of the simulations.
108
109 \subsection{Lattice Model\label{In:ssec:model}}
110 The gel-like characteristic (relatively immobile molecules) exhibited
111 by the ripple phase makes it feasible to apply a lattice model to
112 study the system. The popular $2$ dimensional lattice models, {\it
113 i.e.}, the Ising, $X-Y$, and Heisenberg models, show {\it frustration}
114 on triangular lattices. The Hamiltonians of these systems are given by
115 \begin{equation}
116 H =
117 \begin{cases}
118 -J \sum_n \sum_{n'} s_n s_n' & \text{Ising}, \\
119 -J \sum_n \sum_{n'} \vec s_n \cdot \vec s_{n'} & \text{$X-Y$ and
120 Heisenberg},
121 \end{cases}
122 \end{equation}
123 where $J$ has non zero value only when spins $s_n$ ($\vec s_n$) and
124 $s_{n'}$ ($\vec s_{n'}$) are nearest neighbors. When $J > 0$, spins
125 prefer an aligned structure, and if $J < 0$, spins prefer an
126 anti-aligned structure.
127
128 \begin{figure}
129 \centering
130 \includegraphics[width=3in]{./figures/inFrustration.pdf}
131 \caption{Frustration on triangular lattice, the spins and dipoles are
132 represented by arrows. The multiple local minima of energy states
133 induce the frustration for spins and dipoles picking the directions.}
134 \label{Infig:frustration}
135 \end{figure}
136 The spins in figure~\ref{Infig:frustration} shows an illustration of
137 the frustration for $J < 0$ on a triangular lattice. There are
138 multiple local minima energy states which are independent of the
139 direction of the spin on top of the triangle, therefore infinite
140 possibilities for the packing of spins which induces what is known as
141 ``complete regular frustration'' which leads to disordered low
142 temperature phases. The similarity goes to the dipoles on a hexagonal
143 lattice, which are shown by the dipoles in
144 figure~\ref{Infig:frustration}. In this circumstance, the dipoles want
145 to be aligned, however, due to the long wave fluctuation, at low
146 temperature, the aligned state becomes unstable, vortex is formed and
147 results in multiple local minima of energy states. The dipole on the
148 center of the hexagonal lattice is frustrated.
149
150 The lack of translational degrees of freedom in lattice models
151 prevents their utilization in models for surface buckling. In
152 chapter~\ref{chap:mc}, a modified lattice model is introduced to
153 tackle this specific situation.
154
155 \section{Overview of Classical Statistical Mechanics\label{In:sec:SM}}
156 Statistical mechanics provides a way to calculate the macroscopic
157 properties of a system from the molecular interactions used in
158 computational simulations. This section serves as a brief introduction
159 to key concepts of classical statistical mechanics that we used in
160 this dissertation. Tolman gives an excellent introduction to the
161 principles of statistical mechanics.~\cite{Tolman1979} A large part of
162 section~\ref{In:sec:SM} will follow Tolman's notation.
163
164 \subsection{Ensembles\label{In:ssec:ensemble}}
165 In classical mechanics, the state of the system is completely
166 described by the positions and momenta of all particles. If we have an
167 $N$ particle system, there are $6N$ coordinates ($3N$ positions $(q_1,
168 q_2, \ldots, q_{3N})$ and $3N$ momenta $(p_1, p_2, \ldots, p_{3N})$)
169 to define the instantaneous state of the system. Each single set of
170 the $6N$ coordinates can be considered as a unique point in a $6N$
171 dimensional space where each perpendicular axis is one of
172 $q_{i\alpha}$ or $p_{i\alpha}$ ($i$ is the particle and $\alpha$ is
173 the spatial axis). This $6N$ dimensional space is known as phase
174 space. The instantaneous state of the system is a single point in
175 phase space. A trajectory is a connected path of points in phase
176 space. An ensemble is a collection of systems described by the same
177 macroscopic observables but which have microscopic state
178 distributions. In phase space an ensemble is a collection of a set of
179 representive points. A density distribution $\rho(q^N, p^N)$ of the
180 representive points in phase space describes the condition of an
181 ensemble of identical systems. Since this density may change with
182 time, it is also a function of time. $\rho(q^N, p^N, t)$ describes the
183 ensemble at a time $t$, and $\rho(q^N, p^N, t')$ describes the
184 ensemble at a later time $t'$. For convenience, we will use $\rho$
185 instead of $\rho(q^N, p^N, t)$ in the following disccusion. If we
186 normalize $\rho$ to unity,
187 \begin{equation}
188 1 = \int d \vec q~^N \int d \vec p~^N \rho,
189 \label{Ineq:normalized}
190 \end{equation}
191 then the value of $\rho$ gives the probability of finding the system
192 in a unit volume in phase space.
193
194 Liouville's theorem describes the change in density $\rho$ with
195 time. The number of representative points at a given volume in phase
196 space at any instant can be written as:
197 \begin{equation}
198 \label{Ineq:deltaN}
199 \delta N = \rho~\delta q_1 \delta q_2 \ldots \delta q_N \delta p_1 \delta p_2 \ldots \delta p_N.
200 \end{equation}
201 To calculate the change in the number of representative points in this
202 volume, let us consider a simple condition: the change in the number
203 of representative points along the $q_1$ axis. The rate of the number
204 of the representative points entering the volume at $q_1$ per unit time
205 is:
206 \begin{equation}
207 \label{Ineq:deltaNatq1}
208 \rho~\dot q_1 \delta q_2 \ldots \delta q_N \delta p_1 \delta p_2 \ldots \delta p_N,
209 \end{equation}
210 and the rate of the number of representative points leaving the volume
211 at another position $q_1 + \delta q_1$ is:
212 \begin{equation}
213 \label{Ineq:deltaNatq1plusdeltaq1}
214 \left( \rho + \frac{\partial \rho}{\partial q_1} \delta q_1 \right)\left(\dot q_1 +
215 \frac{\partial \dot q_1}{\partial q_1} \delta q_1 \right)\delta q_2 \ldots \delta q_N \delta p_1 \delta p_2 \ldots \delta p_N.
216 \end{equation}
217 Here the higher order differentials are neglected. So the change in
218 the number of representative points is the difference between
219 eq.~\ref{Ineq:deltaNatq1} and eq.~\ref{Ineq:deltaNatq1plusdeltaq1},
220 which gives us:
221 \begin{equation}
222 \label{Ineq:deltaNatq1axis}
223 -\left(\rho \frac{\partial {\dot q_1}}{\partial q_1} + \frac{\partial {\rho}}{\partial q_1} \dot q_1 \right)\delta q_1 \delta q_2 \ldots \delta q_N \delta p_1 \delta p_2 \ldots \delta p_N,
224 \end{equation}
225 where, higher order differetials are neglected. If we sum over all the
226 axes in the phase space, we can get the change in the number of
227 representative points in a given volume with time:
228 \begin{equation}
229 \label{Ineq:deltaNatGivenVolume}
230 \frac{d(\delta N)}{dt} = -\sum_{i=1}^N \left[\rho \left(\frac{\partial
231 {\dot q_i}}{\partial q_i} + \frac{\partial
232 {\dot p_i}}{\partial p_i}\right) + \left( \frac{\partial {\rho}}{\partial
233 q_i} \dot q_i + \frac{\partial {\rho}}{\partial p_i} \dot p_i\right)\right]\delta q_1 \delta q_2 \ldots \delta q_N \delta p_1 \delta p_2 \ldots \delta p_N.
234 \end{equation}
235 From Hamilton's equations of motion,
236 \begin{equation}
237 \frac{\partial {\dot q_i}}{\partial q_i} = - \frac{\partial
238 {\dot p_i}}{\partial p_i},
239 \label{Ineq:canonicalFormOfEquationOfMotion}
240 \end{equation}
241 this cancels out the first term on the right side of
242 eq.~\ref{Ineq:deltaNatGivenVolume}. If both sides of
243 eq.~\ref{Ineq:deltaNatGivenVolume} are divided by $\delta q_1 \delta q_2
244 \ldots \delta q_N \delta p_1 \delta p_2 \ldots \delta p_N$, then we
245 arrive at Liouville's theorem:
246 \begin{equation}
247 \left( \frac{\partial \rho}{\partial t} \right)_{q, p} = -\sum_{i} \left(
248 \frac{\partial {\rho}}{\partial
249 q_i} \dot q_i + \frac{\partial {\rho}}{\partial p_i} \dot p_i \right).
250 \label{Ineq:simpleFormofLiouville}
251 \end{equation}
252 This is the basis of statistical mechanics. If we move the right
253 side of equation~\ref{Ineq:simpleFormofLiouville} to the left, we
254 will obtain
255 \begin{equation}
256 \left( \frac{\partial \rho}{\partial t} \right)_{q, p} + \sum_{i} \left(
257 \frac{\partial {\rho}}{\partial
258 q_i} \dot q_i + \frac{\partial {\rho}}{\partial p_i} \dot p_i \right)
259 = 0.
260 \label{Ineq:anotherFormofLiouville}
261 \end{equation}
262 It is easy to note that the left side of
263 equation~\ref{Ineq:anotherFormofLiouville} is the total derivative of
264 $\rho$ with respect to $t$, which means
265 \begin{equation}
266 \frac{d \rho}{dt} = 0,
267 \label{Ineq:conservationofRho}
268 \end{equation}
269 and the rate of density change is zero in the neighborhood of any
270 selected moving representative points in the phase space.
271
272 The condition of the ensemble is determined by the density
273 distribution. If we consider the density distribution as only a
274 function of $q$ and $p$, which means the rate of change of the phase
275 space density in the neighborhood of all representative points in the
276 phase space is zero,
277 \begin{equation}
278 \left( \frac{\partial \rho}{\partial t} \right)_{q, p} = 0.
279 \label{Ineq:statEquilibrium}
280 \end{equation}
281 We may conclude the ensemble is in {\it statistical equilibrium}. An
282 ensemble in statistical equilibrium means the system is also in
283 macroscopic equilibrium. If $\left( \frac{\partial \rho}{\partial t}
284 \right)_{q, p} = 0$, then
285 \begin{equation}
286 \sum_{i} \left(
287 \frac{\partial {\rho}}{\partial
288 q_i} \dot q_i + \frac{\partial {\rho}}{\partial p_i} \dot p_i \right)
289 = 0.
290 \label{Ineq:constantofMotion}
291 \end{equation}
292 If $\rho$ is a function only of some constant of the motion, $\rho$ is
293 independent of time. For a conservative system, the energy of the
294 system is one of the constants of the motion. There are many
295 thermodynamically relevant ensembles: when the density distribution is
296 constant everywhere in the phase space,
297 \begin{equation}
298 \rho = \mathrm{const.}
299 \label{Ineq:uniformEnsemble}
300 \end{equation}
301 the ensemble is called {\it uniform ensemble}.
302
303 \subsubsection{The Microcanonical Ensemble\label{In:sssec:microcanonical}}
304 Another useful ensemble is the {\it microcanonical ensemble}, for
305 which:
306 \begin{equation}
307 \rho = \delta \left( H(q^N, p^N) - E \right) \frac{1}{\Sigma (N, V, E)}
308 \label{Ineq:microcanonicalEnsemble}
309 \end{equation}
310 where $\Sigma(N, V, E)$ is a normalization constant parameterized by
311 $N$, the total number of particles, $V$, the total physical volume and
312 $E$, the total energy. The physical meaning of $\Sigma(N, V, E)$ is
313 the phase space volume accessible to a microcanonical system with
314 energy $E$ evolving under Hamilton's equations. $H(q^N, p^N)$ is the
315 Hamiltonian of the system. The Gibbs entropy is defined as
316 \begin{equation}
317 S = - k_B \int d \vec q~^N \int d \vec p~^N \rho \ln [C^N \rho],
318 \label{Ineq:gibbsEntropy}
319 \end{equation}
320 where $k_B$ is the Boltzmann constant and $C^N$ is a number which
321 makes the argument of $\ln$ dimensionless. In this case, $C^N$ is the
322 total phase space volume of one state. The entropy of a microcanonical
323 ensemble is given by
324 \begin{equation}
325 S = k_B \ln \left(\frac{\Sigma(N, V, E)}{C^N}\right).
326 \label{Ineq:entropy}
327 \end{equation}
328
329 \subsubsection{The Canonical Ensemble\label{In:sssec:canonical}}
330 If the density distribution $\rho$ is given by
331 \begin{equation}
332 \rho = \frac{1}{Z_N}e^{-H(q^N, p^N) / k_B T},
333 \label{Ineq:canonicalEnsemble}
334 \end{equation}
335 the ensemble is known as the {\it canonical ensemble}. Here,
336 \begin{equation}
337 Z_N = \int d \vec q~^N \int_\Gamma d \vec p~^N e^{-H(q^N, p^N) / k_B T},
338 \label{Ineq:partitionFunction}
339 \end{equation}
340 which is also known as the canonical{\it partition function}. $\Gamma$
341 indicates that the integral is over all phase space. In the canonical
342 ensemble, $N$, the total number of particles, $V$, total volume, and
343 $T$, the temperature, are constants. The systems with the lowest
344 energies hold the largest population. According to maximum principle,
345 thermodynamics maximizes the entropy $S$, implying that
346 \begin{equation}
347 \begin{array}{ccc}
348 \delta S = 0 & \mathrm{and} & \delta^2 S < 0.
349 \end{array}
350 \label{Ineq:maximumPrinciple}
351 \end{equation}
352 From Eq.~\ref{Ineq:maximumPrinciple} and two constrains on the
353 canonical ensemble, {\it i.e.}, total probability and average energy
354 must be conserved, the partition function can be shown to be
355 equivalent to
356 \begin{equation}
357 Z_N = e^{-A/k_B T},
358 \label{Ineq:partitionFunctionWithFreeEnergy}
359 \end{equation}
360 where $A$ is the Helmholtz free energy. The significance of
361 Eq.~\ref{Ineq:entropy} and~\ref{Ineq:partitionFunctionWithFreeEnergy} is
362 that they serve as a connection between macroscopic properties of the
363 system and the distribution of microscopic states.
364
365 There is an implicit assumption that our arguments are based on so
366 far. A representative point in the phase space is equally likely to be
367 found in any energetically allowed region. In other words, all
368 energetically accessible states are represented equally, the
369 probabilities to find the system in any of the accessible states is
370 equal. This is called the principle of equal a {\it priori}
371 probabilities.
372
373 \subsection{Statistical Averages\label{In:ssec:average}}
374 Given a density distribution $\rho$ in phase space, the average
375 of any quantity ($F(q^N, p^N$)) which depends on the coordinates
376 ($q^N$) and the momenta ($p^N$) for all the systems in the ensemble
377 can be calculated based on the definition shown by
378 Eq.~\ref{Ineq:statAverage1}
379 \begin{equation}
380 \langle F(q^N, p^N) \rangle = \frac{\int d \vec q~^N \int d \vec p~^N
381 F(q^N, p^N) \rho}{\int d \vec q~^N \int d \vec p~^N \rho}.
382 \label{Ineq:statAverage1}
383 \end{equation}
384 Since the density distribution $\rho$ is normalized to unity, the mean
385 value of $F(q^N, p^N)$ is simplified to
386 \begin{equation}
387 \langle F(q^N, p^N) \rangle = \int d \vec q~^N \int d \vec p~^N F(q^N,
388 p^N) \rho,
389 \label{Ineq:statAverage2}
390 \end{equation}
391 called the {\it ensemble average}. However, the quantity is often
392 averaged for a finite time in real experiments,
393 \begin{equation}
394 \langle F(q^N, p^N) \rangle_t = \lim_{T \rightarrow \infty}
395 \frac{1}{T} \int_{t_0}^{t_0+T} F[q^N(t), p^N(t)] dt.
396 \label{Ineq:timeAverage1}
397 \end{equation}
398 Usually this time average is independent of $t_0$ in statistical
399 mechanics, so Eq.~\ref{Ineq:timeAverage1} becomes
400 \begin{equation}
401 \langle F(q^N, p^N) \rangle_t = \lim_{T \rightarrow \infty}
402 \frac{1}{T} \int_{0}^{T} F[q^N(t), p^N(t)] dt
403 \label{Ineq:timeAverage2}
404 \end{equation}
405 for an infinite time interval.
406
407 \subsubsection{Ergodicity\label{In:sssec:ergodicity}}
408 The {\it ergodic hypothesis}, an important hypothesis governing modern
409 computer simulations states that the system will eventually pass
410 arbitrarily close to any point that is energetically accessible in
411 phase space. Mathematically, this leads to
412 \begin{equation}
413 \langle F(q^N, p^N) \rangle = \langle F(q^N, p^N) \rangle_t.
414 \label{Ineq:ergodicity}
415 \end{equation}
416 Eq.~\ref{Ineq:ergodicity} validates Molecular Dynamics as a form of
417 averaging for sufficiently ergodic systems. Also Monte Carlo may be
418 used to obtain ensemble averages of a quantity which are related to
419 time averages measured in experiments.
420
421 \subsection{Correlation Functions\label{In:ssec:corr}}
422 Thermodynamic properties can be computed by equilibrium statistical
423 mechanics. On the other hand, {\it Time correlation functions} are a
424 powerful tool to understand the evolution of a dynamical
425 systems. Imagine that property $A(q^N, p^N, t)$ as a function of
426 coordinates $q^N$ and momenta $p^N$ has an intial value at $t_0$, and
427 at a later time $t_0 + \tau$ this value has changed. If $\tau$ is very
428 small, the change of the value is minor, and the later value of
429 $A(q^N, p^N, t_0 + \tau)$ is correlated to its initial value. However,
430 when $\tau$ is large, this correlation is lost. A time correlation
431 function measures this relationship and is defined
432 by~\cite{Berne90}
433 \begin{equation}
434 C_{AA}(\tau) = \langle A(0)A(\tau) \rangle = \lim_{T \rightarrow
435 \infty}
436 \frac{1}{T} \int_{0}^{T} dt A(t) A(t + \tau).
437 \label{Ineq:autocorrelationFunction}
438 \end{equation}
439 Eq.~\ref{Ineq:autocorrelationFunction} is the correlation function of
440 a single variable, called an {\it autocorrelation function}. The
441 definition of the correlation function for two different variables is
442 similar to that of autocorrelation function, which is
443 \begin{equation}
444 C_{AB}(\tau) = \langle A(0)B(\tau) \rangle = \lim_{T \rightarrow \infty}
445 \frac{1}{T} \int_{0}^{T} dt A(t) B(t + \tau),
446 \label{Ineq:crosscorrelationFunction}
447 \end{equation}
448 and called {\it cross correlation function}.
449
450 We know from the ergodic hypothesis that there is a relationship
451 between time average and ensemble average. We can put the correlation
452 function in a classical mechanics form,
453 \begin{equation}
454 C_{AA}(\tau) = \int d \vec q~^N \int d \vec p~^N A[(q^N(t), p^N(t)]
455 A[q^N(t+\tau), p^N(t+\tau)] \rho(q, p)
456 \label{Ineq:autocorrelationFunctionCM}
457 \end{equation}
458 and
459 \begin{equation}
460 C_{AB}(\tau) = \int d \vec q~^N \int d \vec p~^N A[(q^N(t), p^N(t)]
461 B[q^N(t+\tau), p^N(t+\tau)] \rho(q, p)
462 \label{Ineq:crosscorrelationFunctionCM}
463 \end{equation}
464 as the autocorrelation function and cross correlation functions
465 respectively. $\rho(q, p)$ is the density distribution at equillibrium
466 in phase space. In many cases, correlation functions decay as a
467 single exponential in time
468 \begin{equation}
469 C(t) \sim e^{-t / \tau_r},
470 \label{Ineq:relaxation}
471 \end{equation}
472 where $\tau_r$ is known as relaxation time which describes the rate of
473 the decay.
474
475 \section{Methodology\label{In:sec:method}}
476 The simulations performed in this dissertation branch into two main
477 categories, Monte Carlo and Molecular Dynamics. There are two main
478 differences between Monte Carlo and Molecular Dynamics
479 simulations. One is that the Monte Carlo simulations are time
480 independent methods of sampling structural features of an ensemble,
481 while Molecular Dynamics simulations provide dynamic
482 information. Additionally, Monte Carlo methods are a stochastic
483 processes, the future configurations of the system are not determined
484 by its past. However, in Molecular Dynamics, the system is propagated
485 by Newton's equation of motion, and the trajectory of the system
486 evolving in phase space is deterministic. Brief introductions of the
487 two algorithms are given in section~\ref{In:ssec:mc}
488 and~\ref{In:ssec:md}. Langevin Dynamics, an extension of the Molecular
489 Dynamics that includes implicit solvent effects, is introduced by
490 section~\ref{In:ssec:ld}.
491
492 \subsection{Monte Carlo\label{In:ssec:mc}}
493 A Monte Carlo integration algorithm was first introduced by Metropolis
494 {\it et al.}~\cite{Metropolis53} The basic Metropolis Monte Carlo
495 algorithm is usually applied to the canonical ensemble, a
496 Boltzmann-weighted ensemble, in which the $N$, the total number of
497 particles, $V$, total volume, $T$, temperature are constants. An
498 average in this ensemble is given
499 \begin{equation}
500 \langle A \rangle = \frac{1}{Z_N} \int d \vec q~^N \int d \vec p~^N
501 A(q^N, p^N) e^{-H(q^N, p^N) / k_B T}.
502 \label{Ineq:energyofCanonicalEnsemble}
503 \end{equation}
504 The Hamiltonian is the sum of the kinetic energy, $K(p^N)$, a function
505 of momenta and the potential energy, $U(q^N)$, a function of
506 positions,
507 \begin{equation}
508 H(q^N, p^N) = K(p^N) + U(q^N).
509 \label{Ineq:hamiltonian}
510 \end{equation}
511 If the property $A$ is a function only of position ($ A = A(q^N)$),
512 the mean value of $A$ can be reduced to
513 \begin{equation}
514 \langle A \rangle = \frac{\int d \vec q~^N A e^{-U(q^N) / k_B T}}{\int d \vec q~^N e^{-U(q^N) / k_B T}},
515 \label{Ineq:configurationIntegral}
516 \end{equation}
517 The kinetic energy $K(p^N)$ is factored out in
518 Eq.~\ref{Ineq:configurationIntegral}. $\langle A
519 \rangle$ is now a configuration integral, and
520 Eq.~\ref{Ineq:configurationIntegral} is equivalent to
521 \begin{equation}
522 \langle A \rangle = \int d \vec q~^N A \rho(q^N).
523 \label{Ineq:configurationAve}
524 \end{equation}
525
526 In a Monte Carlo simulation of a system in the canonical ensemble, the
527 probability of the system being in a state $s$ is $\rho_s$, the change
528 of this probability with time is given by
529 \begin{equation}
530 \frac{d \rho_s}{dt} = \sum_{s'} [ -w_{ss'}\rho_s + w_{s's}\rho_{s'} ],
531 \label{Ineq:timeChangeofProb}
532 \end{equation}
533 where $w_{ss'}$ is the tansition probability of going from state $s$
534 to state $s'$. Since $\rho_s$ is independent of time at equilibrium,
535 \begin{equation}
536 \frac{d \rho_{s}^{equilibrium}}{dt} = 0,
537 \label{Ineq:equiProb}
538 \end{equation}
539 which means $\sum_{s'} [ -w_{ss'}\rho_s + w_{s's}\rho_{s'} ]$ is $0$
540 for all $s'$. So
541 \begin{equation}
542 \frac{\rho_s^{equilibrium}}{\rho_{s'}^{equilibrium}} = \frac{w_{s's}}{w_{ss'}}.
543 \label{Ineq:relationshipofRhoandW}
544 \end{equation}
545 If the ratio of state populations
546 \begin{equation}
547 \frac{\rho_s^{equilibrium}}{\rho_{s'}^{equilibrium}} = e^{-(U_s - U_{s'}) / k_B T}.
548 \label{Ineq:satisfyofBoltzmannStatistics}
549 \end{equation}
550 then the ratio of transition probabilities,
551 \begin{equation}
552 \frac{w_{s's}}{w_{ss'}} = e^{-(U_s - U_{s'}) / k_B T},
553 \label{Ineq:conditionforBoltzmannStatistics}
554 \end{equation}
555 An algorithm that shows how Monte Carlo simulation generates a
556 transition probability governed by
557 \ref{Ineq:conditionforBoltzmannStatistics}, is given schematically as,
558 \begin{enumerate}
559 \item\label{Initm:oldEnergy} Choose a particle randomly, and calculate
560 the energy of the rest of the system due to the location of the particle.
561 \item\label{Initm:newEnergy} Make a random displacement of the particle,
562 calculate the new energy taking the movement of the particle into account.
563 \begin{itemize}
564 \item If the energy goes down, keep the new configuration and return to
565 step~\ref{Initm:oldEnergy}.
566 \item If the energy goes up, pick a random number between $[0,1]$.
567 \begin{itemize}
568 \item If the random number smaller than
569 $e^{-(U_{new} - U_{old})} / k_B T$, keep the new configuration and return to
570 step~\ref{Initm:oldEnergy}.
571 \item If the random number is larger than
572 $e^{-(U_{new} - U_{old})} / k_B T$, keep the old configuration and return to
573 step~\ref{Initm:oldEnergy}.
574 \end{itemize}
575 \end{itemize}
576 \item\label{Initm:accumulateAvg} Accumulate the averages based on the
577 current configuartion.
578 \item Go to step~\ref{Initm:oldEnergy}.
579 \end{enumerate}
580 It is important for sampling accurately that the old configuration is
581 sampled again if it is kept.
582
583 \subsection{Molecular Dynamics\label{In:ssec:md}}
584 Although some of properites of the system can be calculated from the
585 ensemble average in Monte Carlo simulations, due to the absence of the
586 time dependence, it is impossible to gain information on dynamic
587 properties from Monte Carlo simulations. Molecular Dynamics evolves
588 the positions and momenta of the particles in the system. The
589 evolution of the system obeys the laws of classical mechanics, and in
590 most situations, there is no need to account for quantum effects. In a
591 real experiment, the instantaneous positions and momenta of the
592 particles in the system are neither important nor measurable, the
593 observable quantities are usually an average value for a finite time
594 interval. These quantities are expressed as a function of positions
595 and momenta in Molecular Dynamics simulations. For example,
596 temperature of the system is defined by
597 \begin{equation}
598 \frac{3}{2} N k_B T = \langle \sum_{i=1}^N \frac{1}{2} m_i v_i \rangle,
599 \label{Ineq:temperature}
600 \end{equation}
601 here $m_i$ is the mass of particle $i$ and $v_i$ is the velocity of
602 particle $i$. The right side of Eq.~\ref{Ineq:temperature} is the
603 average kinetic energy of the system.
604
605 The initial positions of the particles are chosen so that there is no
606 overlap of the particles. The initial velocities at first are
607 distributed randomly to the particles using a Maxwell-Boltzmann
608 ditribution, and then shifted to make the total linear momentum of the
609 system $0$.
610
611 The core of Molecular Dynamics simulations is the calculation of
612 forces and the integration algorithm. Calculation of the forces often
613 involves enormous effort. This is the most time consuming step in the
614 Molecular Dynamics scheme. Evaluation of the forces is mathematically
615 simple,
616 \begin{equation}
617 f(q) = - \frac{\partial U(q)}{\partial q},
618 \label{Ineq:force}
619 \end{equation}
620 where $U(q)$ is the potential of the system. However, the numerical
621 details of this computation are often quite complex. Once the forces
622 computed, the positions and velocities are updated by integrating
623 Hamilton's equations of motion,
624 \begin{eqnarray}
625 \dot p_i & = & -\frac{\partial H}{\partial q_i} = -\frac{\partial
626 U(q_i)}{\partial q_i} = f(q_i) \\
627 \dot q_i & = & p_i
628 \label{Ineq:newton}
629 \end{eqnarray}
630 The classic example of an integrating algorithm is the Verlet
631 algorithm, which is one of the simplest algorithms for integrating the
632 equations of motion. The Taylor expansion of the position at time $t$
633 is
634 \begin{equation}
635 q(t+\Delta t)= q(t) + v(t) \Delta t + \frac{f(t)}{2m}\Delta t^2 +
636 \frac{\Delta t^3}{3!}\frac{\partial^3 q(t)}{\partial t^3} +
637 \mathcal{O}(\Delta t^4)
638 \label{Ineq:verletFuture}
639 \end{equation}
640 for a later time $t+\Delta t$, and
641 \begin{equation}
642 q(t-\Delta t)= q(t) - v(t) \Delta t + \frac{f(t)}{2m}\Delta t^2 -
643 \frac{\Delta t^3}{3!}\frac{\partial^3 q(t)}{\partial t^3} +
644 \mathcal{O}(\Delta t^4) ,
645 \label{Ineq:verletPrevious}
646 \end{equation}
647 for a previous time $t-\Delta t$. Adding Eq.~\ref{Ineq:verletFuture}
648 and~\ref{Ineq:verletPrevious} gives
649 \begin{equation}
650 q(t+\Delta t)+q(t-\Delta t) =
651 2q(t) + \frac{f(t)}{m}\Delta t^2 + \mathcal{O}(\Delta t^4),
652 \label{Ineq:verletSum}
653 \end{equation}
654 so, the new position can be expressed as
655 \begin{equation}
656 q(t+\Delta t) \approx
657 2q(t) - q(t-\Delta t) + \frac{f(t)}{m}\Delta t^2.
658 \label{Ineq:newPosition}
659 \end{equation}
660 The higher order terms in $\Delta t$ are omitted.
661
662 Numerous techniques and tricks have been applied to Molecular Dynamics
663 simulations to gain greater efficiency or accuracy. The engine used in
664 this dissertation for the Molecular Dynamics simulations is {\sc
665 oopse}. More details of the algorithms and techniques used in this
666 code can be found in Ref.~\cite{Meineke2005}.
667
668 \subsection{Langevin Dynamics\label{In:ssec:ld}}
669 In many cases, the properites of the solvent in a system, like the
670 water in the lipid-water system studied in this dissertation, are less
671 interesting to the researchers than the behavior of the
672 solute. However, the major computational expense is ofter the
673 solvent-solvent interation, this is due to the large number of the
674 solvent molecules when compared to the number of solute molecules, the
675 ratio of the number of lipid molecules to the number of water
676 molecules is $1:25$ in our lipid-water system. The efficiency of the
677 Molecular Dynamics simulations is greatly reduced, with up to 85\% of
678 CPU time spent calculating only water-water interactions.
679
680 As an extension of the Molecular Dynamics methodologies, Langevin
681 Dynamics seeks a way to avoid integrating the equations of motion for
682 solvent particles without losing the solvent effects on the solute
683 particles. One common approximation is to express the coupling of the
684 solute and solvent as a set of harmonic oscillators. The Hamiltonian
685 of such a system is written as
686 \begin{equation}
687 H = \frac{p^2}{2m} + U(q) + H_B + \Delta U(q),
688 \label{Ineq:hamiltonianofCoupling}
689 \end{equation}
690 where $H_B$ is the Hamiltonian of the bath which is a set of N
691 harmonic oscillators
692 \begin{equation}
693 H_B = \sum_{\alpha = 1}^{N} \left\{ \frac{p_\alpha^2}{2m_\alpha} +
694 \frac{1}{2} m_\alpha \omega_\alpha^2 q_\alpha^2\right\},
695 \label{Ineq:hamiltonianofBath}
696 \end{equation}
697 $\alpha$ runs over all the degree of freedoms of the bath,
698 $\omega_\alpha$ is the bath frequency of oscillator $\alpha$, and
699 $\Delta U(q)$ is the bilinear coupling given by
700 \begin{equation}
701 \Delta U = -\sum_{\alpha = 1}^{N} g_\alpha q_\alpha q,
702 \label{Ineq:systemBathCoupling}
703 \end{equation}
704 where $g_\alpha$ is the coupling constant for oscillator $\alpha$. By
705 solving the Hamilton's equations of motion, the {\it Generalized
706 Langevin Equation} for this system is derived as
707 \begin{equation}
708 m \ddot q = -\frac{\partial W(q)}{\partial q} - \int_0^t \xi(t) \dot q(t-t')dt' + R(t),
709 \label{Ineq:gle}
710 \end{equation}
711 with mean force,
712 \begin{equation}
713 W(q) = U(q) - \sum_{\alpha = 1}^N \frac{g_\alpha^2}{2m_\alpha
714 \omega_\alpha^2}q^2.
715 \label{Ineq:meanForce}
716 \end{equation}
717 The {\it friction kernel}, $\xi(t)$, depends only on the coordinates
718 of the solute particles,
719 \begin{equation}
720 \xi(t) = \sum_{\alpha = 1}^N \frac{-g_\alpha}{m_\alpha
721 \omega_\alpha} \cos(\omega_\alpha t),
722 \label{Ineq:xiforGLE}
723 \end{equation}
724 and a ``random'' force,
725 \begin{equation}
726 R(t) = \sum_{\alpha = 1}^N \left( g_\alpha q_\alpha(0)-\frac{g_\alpha}{m_\alpha
727 \omega_\alpha^2}q(0)\right) \cos(\omega_\alpha t) + \frac{\dot
728 q_\alpha(0)}{\omega_\alpha} \sin(\omega_\alpha t),
729 \label{Ineq:randomForceforGLE}
730 \end{equation}
731 depends only on the initial conditions. The relationship of friction
732 kernel $\xi(t)$ and random force $R(t)$ is given by the second
733 fluctuation dissipation theorem,
734 \begin{equation}
735 \xi(t) = \frac{1}{k_B T} \langle R(t)R(0) \rangle.
736 \label{Ineq:relationshipofXiandR}
737 \end{equation}
738 In the harmonic bath this relation is exact and provable from the
739 definitions of these quantities. In the limit of static friction,
740 \begin{equation}
741 \xi(t) = 2 \xi_0 \delta(t).
742 \label{Ineq:xiofStaticFriction}
743 \end{equation}
744 After substituting $\xi(t)$ into Eq.~\ref{Ineq:gle} with
745 Eq.~\ref{Ineq:xiofStaticFriction}, the {\it Langevin Equation} is
746 extracted,
747 \begin{equation}
748 m \ddot q = -\frac{\partial U(q)}{\partial q} - \xi \dot q(t) + R(t).
749 \label{Ineq:langevinEquation}
750 \end{equation}
751 Application of the Langevin Equation to dynamic simulations is
752 discussed in Ch.~\ref{chap:ld}.