ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/xDissertation/Introduction.tex
Revision: 3372
Committed: Wed Mar 19 21:04:47 2008 UTC (16 years, 5 months ago) by xsun
Content type: application/x-tex
File size: 32509 byte(s)
Log Message:
writing.

File Contents

# User Rev Content
1 xsun 3336 \chapter{\label{chap:intro}INTRODUCTION AND BACKGROUND}
2 xsun 3347
3     \section{Background on the Problem\label{In:sec:pro}}
4 xsun 3365 Phospholipid molecules are the primary topic of this dissertation
5     because of their critical role as the foundation of biological
6 xsun 3372 membranes. Lipids, when dispersed in water, self assemble into a
7     mumber of topologically distinct bilayer structures. The phase
8     behavior of lipid bilayers has been explored
9 xsun 3365 experimentally~\cite{Cevc87}, however, a complete understanding of the
10 xsun 3372 mechanism and driving forces behind the various phases has not been
11     achieved.
12 xsun 3347
13     \subsection{Ripple Phase\label{In:ssec:ripple}}
14 xsun 3365 The $P_{\beta'}$ {\it ripple phase} of lipid bilayers, named from the
15 xsun 3347 periodic buckling of the membrane, is an intermediate phase which is
16     developed either from heating the gel phase $L_{\beta'}$ or cooling
17 xsun 3372 the fluid phase $L_\alpha$. A Sketch is shown in
18     figure~\ref{Infig:phaseDiagram}.~\cite{Cevc87}
19     \begin{figure}
20     \centering
21     \includegraphics[width=\linewidth]{./figures/inPhaseDiagram.pdf}
22     \caption{A phase diagram of lipid bilayer. With increasing the
23     temperature, the bilayer can go through a gel ($L_{\beta'}$), ripple
24     ($P_{\beta'}$) to fluid ($L_\alpha$) phase transition.}
25     \label{Infig:phaseDiagram}
26     \end{figure}
27     Most structural information of the ripple phase has been obtained by
28     the X-ray diffraction~\cite{Sun96,Katsaras00} and freeze-fracture
29     electron microscopy (FFEM).~\cite{Copeland80,Meyer96} The X-ray
30     diffraction work by Katsaras {\it et al.} showed that a rich phase
31     diagram exhibiting both {\it asymmetric} and {\it symmetric} ripples
32     is possible for lecithin bilayers.\cite{Katsaras00} Recently,
33     Kaasgaard {\it et al.} used atomic force microscopy (AFM) to observe
34     ripple phase morphology in bilayers supported on
35     mica.~\cite{Kaasgaard03}
36     \begin{figure}
37     \centering
38     \includegraphics[width=\linewidth]{./figures/inRipple.pdf}
39     \caption{The experimental observed ripple phase. The top image is
40     obtained by X-ray diffraction~\cite{Sun96}, and the bottom one is
41     observed by AFM.~\cite{Kaasgaard03}}
42     \label{Infig:ripple}
43     \end{figure}
44     Figure~\ref{Infig:ripple} shows the ripple phase oberved by X-ray
45     diffraction and AFM. The experimental results provide strong support
46     for a 2-dimensional triangular packing lattice of the lipid molecules
47     within the ripple phase. This is a notable change from the observed
48     lipid packing within the gel phase,~\cite{Cevc87} although Tenchov
49     {\it et al.} have recently observed near-hexagonal packing in some
50     phosphatidylcholine (PC) gel phases.~\cite{Tenchov2001} However, the
51     physical mechanism for the formation of the ripple phase has never
52     been explained and the microscopic structure of the ripple phase has
53     never been elucidated by experiments. Computational simulation is a
54     perfect tool to study the microscopic properties for a
55     system. However, the large length scale the ripple structure and the
56     long time scale of the formation of the ripples are crucial obstacles
57     to performing the actual work. The principal ideas explored in this
58     dissertation are attempts to break the computational task up by
59 xsun 3347 \begin{itemize}
60 xsun 3365 \item Simplifying the lipid model.
61     \item Improving algorithm for integrating the equations of motion.
62 xsun 3347 \end{itemize}
63 xsun 3365 In chapters~\ref{chap:mc} and~\ref{chap:md}, we use a simple point
64 xsun 3372 dipole spin model and a coarse-grained molecualr scale model to
65     perform the Monte Carlo and Molecular Dynamics simulations
66     respectively, and in chapter~\ref{chap:ld}, we develop a Langevin
67     Dynamics algorithm which excludes the explicit solvent to improve the
68     efficiency of the simulations.
69 xsun 3347
70     \subsection{Lattice Model\label{In:ssec:model}}
71 xsun 3372 The gel-like characteristic (relatively immobile molecules) exhibited
72     by the ripple phase makes it feasible to apply a lattice model to
73     study the system. The popular $2$ dimensional lattice models, {\it
74     i.e.}, the Ising, $X-Y$, and Heisenberg models, show {\it frustration}
75     on triangular lattices. The Hamiltonians of these systems are given by
76 xsun 3365 \begin{equation}
77     H =
78     \begin{cases}
79     -J \sum_n \sum_{n'} s_n s_n' & \text{Ising}, \\
80     -J \sum_n \sum_{n'} \vec s_n \cdot \vec s_{n'} & \text{$X-Y$ and
81     Heisenberg},
82     \end{cases}
83     \end{equation}
84     where $J$ has non zero value only when spins $s_n$ ($\vec s_n$) and
85 xsun 3372 $s_{n'}$ ($\vec s_{n'}$) are nearest neighbors. When $J > 0$, spins
86     prefer an aligned structure, and if $J < 0$, spins prefer an
87     anti-aligned structure.
88 xsun 3371
89 xsun 3347 \begin{figure}
90     \centering
91 xsun 3371 \includegraphics[width=3in]{./figures/inFrustration.pdf}
92     \caption{Frustration on triangular lattice, the spins and dipoles are
93     represented by arrows. The multiple local minima of energy states
94     induce the frustration for spins and dipoles picking the directions.}
95 xsun 3354 \label{Infig:frustration}
96 xsun 3347 \end{figure}
97 xsun 3371 The spins in figure~\ref{Infig:frustration} shows an illustration of
98     the frustration for $J < 0$ on a triangular lattice. There are
99     multiple local minima energy states which are independent of the
100     direction of the spin on top of the triangle, therefore infinite
101 xsun 3372 possibilities for the packing of spins which induces what is known as
102 xsun 3371 ``complete regular frustration'' which leads to disordered low
103     temperature phases. The similarity goes to the dipoles on a hexagonal
104     lattice, which are shown by the dipoles in
105     figure~\ref{Infig:frustration}. In this circumstance, the dipoles want
106     to be aligned, however, due to the long wave fluctuation, at low
107     temperature, the aligned state becomes unstable, vortex is formed and
108     results in multiple local minima of energy states. The dipole on the
109     center of the hexagonal lattice is frustrated.
110 xsun 3347
111 xsun 3372 The lack of translational degrees of freedom in lattice models
112     prevents their utilization in models for surface buckling. In
113     chapter~\ref{chap:mc}, a modified lattice model is introduced to
114     tackle this specific situation.
115 xsun 3347
116     \section{Overview of Classical Statistical Mechanics\label{In:sec:SM}}
117     Statistical mechanics provides a way to calculate the macroscopic
118     properties of a system from the molecular interactions used in
119     computational simulations. This section serves as a brief introduction
120     to key concepts of classical statistical mechanics that we used in
121     this dissertation. Tolman gives an excellent introduction to the
122 xsun 3372 principles of statistical mechanics.~\cite{Tolman1979} A large part of
123 xsun 3347 section~\ref{In:sec:SM} will follow Tolman's notation.
124    
125     \subsection{Ensembles\label{In:ssec:ensemble}}
126     In classical mechanics, the state of the system is completely
127     described by the positions and momenta of all particles. If we have an
128 xsun 3372 $N$ particle system, there are $6N$ coordinates ($3N$ positions $(q_1,
129 xsun 3347 q_2, \ldots, q_{3N})$ and $3N$ momenta $(p_1, p_2, \ldots, p_{3N})$)
130     to define the instantaneous state of the system. Each single set of
131     the $6N$ coordinates can be considered as a unique point in a $6N$
132     dimensional space where each perpendicular axis is one of
133     $q_{i\alpha}$ or $p_{i\alpha}$ ($i$ is the particle and $\alpha$ is
134     the spatial axis). This $6N$ dimensional space is known as phase
135     space. The instantaneous state of the system is a single point in
136     phase space. A trajectory is a connected path of points in phase
137     space. An ensemble is a collection of systems described by the same
138     macroscopic observables but which have microscopic state
139     distributions. In phase space an ensemble is a collection of a set of
140     representive points. A density distribution $\rho(q^N, p^N)$ of the
141     representive points in phase space describes the condition of an
142     ensemble of identical systems. Since this density may change with
143     time, it is also a function of time. $\rho(q^N, p^N, t)$ describes the
144     ensemble at a time $t$, and $\rho(q^N, p^N, t')$ describes the
145     ensemble at a later time $t'$. For convenience, we will use $\rho$
146     instead of $\rho(q^N, p^N, t)$ in the following disccusion. If we
147     normalize $\rho$ to unity,
148     \begin{equation}
149     1 = \int d \vec q~^N \int d \vec p~^N \rho,
150 xsun 3354 \label{Ineq:normalized}
151 xsun 3347 \end{equation}
152     then the value of $\rho$ gives the probability of finding the system
153 xsun 3372 in a unit volume in phase space.
154 xsun 3347
155     Liouville's theorem describes the change in density $\rho$ with
156 xsun 3372 time. The number of representative points at a given volume in phase
157 xsun 3347 space at any instant can be written as:
158     \begin{equation}
159 xsun 3354 \label{Ineq:deltaN}
160 xsun 3347 \delta N = \rho~\delta q_1 \delta q_2 \ldots \delta q_N \delta p_1 \delta p_2 \ldots \delta p_N.
161     \end{equation}
162 xsun 3372 To calculate the change in the number of representative points in this
163 xsun 3347 volume, let us consider a simple condition: the change in the number
164 xsun 3372 of representative points along the $q_1$ axis. The rate of the number
165     of the representative points entering the volume at $q_1$ per unit time
166     is:
167 xsun 3347 \begin{equation}
168 xsun 3354 \label{Ineq:deltaNatq1}
169 xsun 3347 \rho~\dot q_1 \delta q_2 \ldots \delta q_N \delta p_1 \delta p_2 \ldots \delta p_N,
170     \end{equation}
171 xsun 3372 and the rate of the number of representative points leaving the volume
172 xsun 3347 at another position $q_1 + \delta q_1$ is:
173     \begin{equation}
174 xsun 3354 \label{Ineq:deltaNatq1plusdeltaq1}
175 xsun 3347 \left( \rho + \frac{\partial \rho}{\partial q_1} \delta q_1 \right)\left(\dot q_1 +
176     \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.
177     \end{equation}
178 xsun 3372 Here the higher order differentials are neglected. So the change in
179     the number of representative points is the difference between
180     eq.~\ref{Ineq:deltaNatq1} and eq.~\ref{Ineq:deltaNatq1plusdeltaq1},
181     which gives us:
182 xsun 3347 \begin{equation}
183 xsun 3354 \label{Ineq:deltaNatq1axis}
184 xsun 3347 -\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,
185     \end{equation}
186     where, higher order differetials are neglected. If we sum over all the
187 xsun 3372 axes in the phase space, we can get the change in the number of
188     representative points in a given volume with time:
189 xsun 3347 \begin{equation}
190 xsun 3354 \label{Ineq:deltaNatGivenVolume}
191 xsun 3347 \frac{d(\delta N)}{dt} = -\sum_{i=1}^N \left[\rho \left(\frac{\partial
192     {\dot q_i}}{\partial q_i} + \frac{\partial
193     {\dot p_i}}{\partial p_i}\right) + \left( \frac{\partial {\rho}}{\partial
194     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.
195     \end{equation}
196 xsun 3372 From Hamilton's equations of motion,
197 xsun 3347 \begin{equation}
198     \frac{\partial {\dot q_i}}{\partial q_i} = - \frac{\partial
199     {\dot p_i}}{\partial p_i},
200 xsun 3354 \label{Ineq:canonicalFormOfEquationOfMotion}
201 xsun 3347 \end{equation}
202     this cancels out the first term on the right side of
203 xsun 3354 eq.~\ref{Ineq:deltaNatGivenVolume}. If both sides of
204     eq.~\ref{Ineq:deltaNatGivenVolume} are divided by $\delta q_1 \delta q_2
205 xsun 3347 \ldots \delta q_N \delta p_1 \delta p_2 \ldots \delta p_N$, then we
206 xsun 3372 arrive at Liouville's theorem:
207 xsun 3347 \begin{equation}
208     \left( \frac{\partial \rho}{\partial t} \right)_{q, p} = -\sum_{i} \left(
209     \frac{\partial {\rho}}{\partial
210     q_i} \dot q_i + \frac{\partial {\rho}}{\partial p_i} \dot p_i \right).
211 xsun 3354 \label{Ineq:simpleFormofLiouville}
212 xsun 3347 \end{equation}
213     This is the basis of statistical mechanics. If we move the right
214 xsun 3354 side of equation~\ref{Ineq:simpleFormofLiouville} to the left, we
215 xsun 3347 will obtain
216     \begin{equation}
217     \left( \frac{\partial \rho}{\partial t} \right)_{q, p} + \sum_{i} \left(
218     \frac{\partial {\rho}}{\partial
219     q_i} \dot q_i + \frac{\partial {\rho}}{\partial p_i} \dot p_i \right)
220     = 0.
221 xsun 3354 \label{Ineq:anotherFormofLiouville}
222 xsun 3347 \end{equation}
223     It is easy to note that the left side of
224 xsun 3354 equation~\ref{Ineq:anotherFormofLiouville} is the total derivative of
225 xsun 3372 $\rho$ with respect to $t$, which means
226 xsun 3347 \begin{equation}
227     \frac{d \rho}{dt} = 0,
228 xsun 3354 \label{Ineq:conservationofRho}
229 xsun 3347 \end{equation}
230     and the rate of density change is zero in the neighborhood of any
231 xsun 3372 selected moving representative points in the phase space.
232 xsun 3347
233     The condition of the ensemble is determined by the density
234     distribution. If we consider the density distribution as only a
235     function of $q$ and $p$, which means the rate of change of the phase
236 xsun 3372 space density in the neighborhood of all representative points in the
237 xsun 3347 phase space is zero,
238     \begin{equation}
239     \left( \frac{\partial \rho}{\partial t} \right)_{q, p} = 0.
240 xsun 3354 \label{Ineq:statEquilibrium}
241 xsun 3347 \end{equation}
242     We may conclude the ensemble is in {\it statistical equilibrium}. An
243 xsun 3372 ensemble in statistical equilibrium means the system is also in
244 xsun 3347 macroscopic equilibrium. If $\left( \frac{\partial \rho}{\partial t}
245     \right)_{q, p} = 0$, then
246     \begin{equation}
247     \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     = 0.
251 xsun 3354 \label{Ineq:constantofMotion}
252 xsun 3347 \end{equation}
253     If $\rho$ is a function only of some constant of the motion, $\rho$ is
254     independent of time. For a conservative system, the energy of the
255 xsun 3372 system is one of the constants of the motion. There are many
256     thermodynamically relevant ensembles: when the density distribution is
257     constant everywhere in the phase space,
258 xsun 3347 \begin{equation}
259     \rho = \mathrm{const.}
260 xsun 3354 \label{Ineq:uniformEnsemble}
261 xsun 3347 \end{equation}
262 xsun 3372 the ensemble is called {\it uniform ensemble}.
263    
264     \subsubsection{The Microcanonical Ensemble\label{In:sssec:microcanonical}}
265     Another useful ensemble is the {\it microcanonical ensemble}, for
266     which:
267 xsun 3347 \begin{equation}
268     \rho = \delta \left( H(q^N, p^N) - E \right) \frac{1}{\Sigma (N, V, E)}
269 xsun 3354 \label{Ineq:microcanonicalEnsemble}
270 xsun 3347 \end{equation}
271     where $\Sigma(N, V, E)$ is a normalization constant parameterized by
272     $N$, the total number of particles, $V$, the total physical volume and
273     $E$, the total energy. The physical meaning of $\Sigma(N, V, E)$ is
274     the phase space volume accessible to a microcanonical system with
275     energy $E$ evolving under Hamilton's equations. $H(q^N, p^N)$ is the
276     Hamiltonian of the system. The Gibbs entropy is defined as
277     \begin{equation}
278     S = - k_B \int d \vec q~^N \int d \vec p~^N \rho \ln [C^N \rho],
279 xsun 3354 \label{Ineq:gibbsEntropy}
280 xsun 3347 \end{equation}
281     where $k_B$ is the Boltzmann constant and $C^N$ is a number which
282 xsun 3372 makes the argument of $\ln$ dimensionless. In this case, $C^N$ is the
283     total phase space volume of one state. The entropy of a microcanonical
284 xsun 3347 ensemble is given by
285     \begin{equation}
286     S = k_B \ln \left(\frac{\Sigma(N, V, E)}{C^N}\right).
287 xsun 3354 \label{Ineq:entropy}
288 xsun 3347 \end{equation}
289 xsun 3372
290     \subsubsection{The Canonical Ensemble\label{In:sssec:canonical}}
291 xsun 3347 If the density distribution $\rho$ is given by
292     \begin{equation}
293     \rho = \frac{1}{Z_N}e^{-H(q^N, p^N) / k_B T},
294 xsun 3354 \label{Ineq:canonicalEnsemble}
295 xsun 3347 \end{equation}
296     the ensemble is known as the {\it canonical ensemble}. Here,
297     \begin{equation}
298     Z_N = \int d \vec q~^N \int_\Gamma d \vec p~^N e^{-H(q^N, p^N) / k_B T},
299 xsun 3354 \label{Ineq:partitionFunction}
300 xsun 3347 \end{equation}
301 xsun 3372 which is also known as the canonical{\it partition function}. $\Gamma$
302     indicates that the integral is over all phase space. In the canonical
303 xsun 3347 ensemble, $N$, the total number of particles, $V$, total volume, and
304 xsun 3372 $T$, the temperature, are constants. The systems with the lowest
305 xsun 3347 energies hold the largest population. According to maximum principle,
306 xsun 3372 thermodynamics maximizes the entropy $S$, implying that
307 xsun 3347 \begin{equation}
308     \begin{array}{ccc}
309     \delta S = 0 & \mathrm{and} & \delta^2 S < 0.
310     \end{array}
311 xsun 3354 \label{Ineq:maximumPrinciple}
312 xsun 3347 \end{equation}
313 xsun 3372 From Eq.~\ref{Ineq:maximumPrinciple} and two constrains on the
314     canonical ensemble, {\it i.e.}, total probability and average energy
315     must be conserved, the partition function can be shown to be
316     equivalent to
317 xsun 3347 \begin{equation}
318     Z_N = e^{-A/k_B T},
319 xsun 3354 \label{Ineq:partitionFunctionWithFreeEnergy}
320 xsun 3347 \end{equation}
321     where $A$ is the Helmholtz free energy. The significance of
322 xsun 3354 Eq.~\ref{Ineq:entropy} and~\ref{Ineq:partitionFunctionWithFreeEnergy} is
323 xsun 3347 that they serve as a connection between macroscopic properties of the
324 xsun 3372 system and the distribution of microscopic states.
325 xsun 3347
326     There is an implicit assumption that our arguments are based on so
327 xsun 3372 far. A representative point in the phase space is equally likely to be
328     found in any energetically allowed region. In other words, all
329     energetically accessible states are represented equally, the
330     probabilities to find the system in any of the accessible states is
331     equal. This is called the principle of equal a {\it priori}
332     probabilities.
333 xsun 3347
334 xsun 3372 \subsection{Statistical Averages\label{In:ssec:average}}
335     Given a density distribution $\rho$ in phase space, the average
336 xsun 3347 of any quantity ($F(q^N, p^N$)) which depends on the coordinates
337     ($q^N$) and the momenta ($p^N$) for all the systems in the ensemble
338     can be calculated based on the definition shown by
339 xsun 3354 Eq.~\ref{Ineq:statAverage1}
340 xsun 3347 \begin{equation}
341 xsun 3372 \langle F(q^N, p^N) \rangle = \frac{\int d \vec q~^N \int d \vec p~^N
342     F(q^N, p^N) \rho}{\int d \vec q~^N \int d \vec p~^N \rho}.
343 xsun 3354 \label{Ineq:statAverage1}
344 xsun 3347 \end{equation}
345     Since the density distribution $\rho$ is normalized to unity, the mean
346     value of $F(q^N, p^N)$ is simplified to
347     \begin{equation}
348 xsun 3372 \langle F(q^N, p^N) \rangle = \int d \vec q~^N \int d \vec p~^N F(q^N,
349     p^N) \rho,
350 xsun 3354 \label{Ineq:statAverage2}
351 xsun 3347 \end{equation}
352 xsun 3372 called the {\it ensemble average}. However, the quantity is often
353     averaged for a finite time in real experiments,
354 xsun 3347 \begin{equation}
355     \langle F(q^N, p^N) \rangle_t = \lim_{T \rightarrow \infty}
356 xsun 3372 \frac{1}{T} \int_{t_0}^{t_0+T} F[q^N(t), p^N(t)] dt.
357 xsun 3354 \label{Ineq:timeAverage1}
358 xsun 3347 \end{equation}
359     Usually this time average is independent of $t_0$ in statistical
360 xsun 3354 mechanics, so Eq.~\ref{Ineq:timeAverage1} becomes
361 xsun 3347 \begin{equation}
362     \langle F(q^N, p^N) \rangle_t = \lim_{T \rightarrow \infty}
363 xsun 3372 \frac{1}{T} \int_{0}^{T} F[q^N(t), p^N(t)] dt
364 xsun 3354 \label{Ineq:timeAverage2}
365 xsun 3347 \end{equation}
366     for an infinite time interval.
367    
368 xsun 3372 \subsubsection{Ergodicity\label{In:sssec:ergodicity}}
369     The {\it ergodic hypothesis}, an important hypothesis governing modern
370     computer simulations states that the system will eventually pass
371 xsun 3347 arbitrarily close to any point that is energetically accessible in
372     phase space. Mathematically, this leads to
373     \begin{equation}
374 xsun 3372 \langle F(q^N, p^N) \rangle = \langle F(q^N, p^N) \rangle_t.
375 xsun 3354 \label{Ineq:ergodicity}
376 xsun 3347 \end{equation}
377 xsun 3372 Eq.~\ref{Ineq:ergodicity} validates Molecular Dynamics as a form of
378     averaging for sufficiently ergodic systems. Also Monte Carlo may be
379     used to obtain ensemble averages of a quantity which are related to
380     time averages measured in experiments.
381 xsun 3347
382 xsun 3372 \subsection{Correlation Functions\label{In:ssec:corr}}
383     Thermodynamic properties can be computed by equilibrium statistical
384     mechanics. On the other hand, {\it Time correlation functions} are a
385     powerful tool to understand the evolution of a dynamical
386     systems. Imagine that property $A(q^N, p^N, t)$ as a function of
387     coordinates $q^N$ and momenta $p^N$ has an intial value at $t_0$, and
388     at a later time $t_0 + \tau$ this value has changed. If $\tau$ is very
389     small, the change of the value is minor, and the later value of
390     $A(q^N, p^N, t_0 + \tau)$ is correlated to its initial value. However,
391     when $\tau$ is large, this correlation is lost. A time correlation
392     function measures this relationship and is defined
393     by~\cite{Berne90}
394 xsun 3347 \begin{equation}
395 xsun 3372 C_{AA}(\tau) = \langle A(0)A(\tau) \rangle = \lim_{T \rightarrow
396     \infty}
397 xsun 3347 \frac{1}{T} \int_{0}^{T} dt A(t) A(t + \tau).
398 xsun 3354 \label{Ineq:autocorrelationFunction}
399 xsun 3347 \end{equation}
400 xsun 3372 Eq.~\ref{Ineq:autocorrelationFunction} is the correlation function of
401     a single variable, called an {\it autocorrelation function}. The
402     definition of the correlation function for two different variables is
403     similar to that of autocorrelation function, which is
404 xsun 3347 \begin{equation}
405 xsun 3372 C_{AB}(\tau) = \langle A(0)B(\tau) \rangle = \lim_{T \rightarrow \infty}
406 xsun 3347 \frac{1}{T} \int_{0}^{T} dt A(t) B(t + \tau),
407 xsun 3354 \label{Ineq:crosscorrelationFunction}
408 xsun 3347 \end{equation}
409     and called {\it cross correlation function}.
410    
411 xsun 3372 We know from the ergodic hypothesis that there is a relationship
412     between time average and ensemble average. We can put the correlation
413     function in a classical mechanics form,
414 xsun 3347 \begin{equation}
415 xsun 3372 C_{AA}(\tau) = \int d \vec q~^N \int d \vec p~^N A[(q^N(t), p^N(t)]
416     A[q^N(t+\tau), q^N(t+\tau)] \rho(q, p)
417 xsun 3354 \label{Ineq:autocorrelationFunctionCM}
418 xsun 3347 \end{equation}
419     and
420     \begin{equation}
421 xsun 3372 C_{AB}(\tau) = \int d \vec q~^N \int d \vec p~^N A[(q^N(t), p^N(t)]
422     B[q^N(t+\tau), q^N(t+\tau)] \rho(q, p)
423 xsun 3354 \label{Ineq:crosscorrelationFunctionCM}
424 xsun 3347 \end{equation}
425     as autocorrelation function and cross correlation function
426     respectively. $\rho(q, p)$ is the density distribution at equillibrium
427     in phase space. In many cases, the correlation function decay is a
428     single exponential
429     \begin{equation}
430     C(t) \sim e^{-t / \tau_r},
431 xsun 3354 \label{Ineq:relaxation}
432 xsun 3347 \end{equation}
433     where $\tau_r$ is known as relaxation time which discribes the rate of
434     the decay.
435    
436     \section{Methodolody\label{In:sec:method}}
437     The simulations performed in this dissertation are branched into two
438     main catalog, Monte Carlo and Molecular Dynamics. There are two main
439     difference between Monte Carlo and Molecular Dynamics simulations. One
440     is that the Monte Carlo simulation is time independent, and Molecular
441     Dynamics simulation is time involved. Another dissimilar is that the
442     Monte Carlo is a stochastic process, the configuration of the system
443     is not determinated by its past, however, using Moleuclar Dynamics,
444     the system is propagated by Newton's equation of motion, the
445     trajectory of the system evolved in the phase space is determined. A
446     brief introduction of the two algorithms are given in
447     section~\ref{In:ssec:mc} and~\ref{In:ssec:md}. An extension of the
448     Molecular Dynamics, Langevin Dynamics, is introduced by
449     section~\ref{In:ssec:ld}.
450    
451     \subsection{Monte Carlo\label{In:ssec:mc}}
452     Monte Carlo algorithm was first introduced by Metropolis {\it et
453     al.}.~\cite{Metropolis53} Basic Monte Carlo algorithm is usually
454     applied to the canonical ensemble, a Boltzmann-weighted ensemble, in
455     which the $N$, the total number of particles, $V$, total volume, $T$,
456     temperature are constants. The average energy is given by substituding
457 xsun 3354 Eq.~\ref{Ineq:canonicalEnsemble} into Eq.~\ref{Ineq:statAverage2},
458 xsun 3347 \begin{equation}
459     \langle E \rangle = \frac{1}{Z_N} \int d \vec q~^N \int d \vec p~^N E e^{-H(q^N, p^N) / k_B T}.
460 xsun 3354 \label{Ineq:energyofCanonicalEnsemble}
461 xsun 3347 \end{equation}
462     So are the other properties of the system. The Hamiltonian is the
463     summation of Kinetic energy $K(p^N)$ as a function of momenta and
464     Potential energy $U(q^N)$ as a function of positions,
465     \begin{equation}
466     H(q^N, p^N) = K(p^N) + U(q^N).
467 xsun 3354 \label{Ineq:hamiltonian}
468 xsun 3347 \end{equation}
469     If the property $A$ is only a function of position ($ A = A(q^N)$),
470     the mean value of $A$ is reduced to
471     \begin{equation}
472     \langle A \rangle = \frac{\int d \vec q~^N \int d \vec p~^N A e^{-U(q^N) / k_B T}}{\int d \vec q~^N \int d \vec p~^N e^{-U(q^N) / k_B T}},
473 xsun 3354 \label{Ineq:configurationIntegral}
474 xsun 3347 \end{equation}
475     The kinetic energy $K(p^N)$ is factored out in
476 xsun 3354 Eq.~\ref{Ineq:configurationIntegral}. $\langle A
477 xsun 3347 \rangle$ is a configuration integral now, and the
478 xsun 3354 Eq.~\ref{Ineq:configurationIntegral} is equivalent to
479 xsun 3347 \begin{equation}
480     \langle A \rangle = \int d \vec q~^N A \rho(q^N).
481 xsun 3354 \label{Ineq:configurationAve}
482 xsun 3347 \end{equation}
483    
484     In a Monte Carlo simulation of canonical ensemble, the probability of
485     the system being in a state $s$ is $\rho_s$, the change of this
486     probability with time is given by
487     \begin{equation}
488     \frac{d \rho_s}{dt} = \sum_{s'} [ -w_{ss'}\rho_s + w_{s's}\rho_{s'} ],
489 xsun 3354 \label{Ineq:timeChangeofProb}
490 xsun 3347 \end{equation}
491     where $w_{ss'}$ is the tansition probability of going from state $s$
492     to state $s'$. Since $\rho_s$ is independent of time at equilibrium,
493     \begin{equation}
494     \frac{d \rho_{s}^{equilibrium}}{dt} = 0,
495 xsun 3354 \label{Ineq:equiProb}
496 xsun 3347 \end{equation}
497     which means $\sum_{s'} [ -w_{ss'}\rho_s + w_{s's}\rho_{s'} ]$ is $0$
498     for all $s'$. So
499     \begin{equation}
500     \frac{\rho_s^{equilibrium}}{\rho_{s'}^{equilibrium}} = \frac{w_{s's}}{w_{ss'}}.
501 xsun 3354 \label{Ineq:relationshipofRhoandW}
502 xsun 3347 \end{equation}
503     If
504     \begin{equation}
505     \frac{w_{s's}}{w_{ss'}} = e^{-(U_s - U_{s'}) / k_B T},
506 xsun 3354 \label{Ineq:conditionforBoltzmannStatistics}
507 xsun 3347 \end{equation}
508     then
509     \begin{equation}
510     \frac{\rho_s^{equilibrium}}{\rho_{s'}^{equilibrium}} = e^{-(U_s - U_{s'}) / k_B T}.
511 xsun 3354 \label{Ineq:satisfyofBoltzmannStatistics}
512 xsun 3347 \end{equation}
513 xsun 3354 Eq.~\ref{Ineq:satisfyofBoltzmannStatistics} implies that
514 xsun 3347 $\rho^{equilibrium}$ satisfies Boltzmann statistics. An algorithm,
515     shows how Monte Carlo simulation generates a transition probability
516 xsun 3354 governed by \ref{Ineq:conditionforBoltzmannStatistics}, is schemed as
517 xsun 3347 \begin{enumerate}
518 xsun 3354 \item\label{Initm:oldEnergy} Choose an particle randomly, calculate the energy.
519     \item\label{Initm:newEnergy} Make a random displacement for particle,
520 xsun 3347 calculate the new energy.
521     \begin{itemize}
522 xsun 3354 \item Keep the new configuration and return to step~\ref{Initm:oldEnergy} if energy
523 xsun 3347 goes down.
524     \item Pick a random number between $[0,1]$ if energy goes up.
525     \begin{itemize}
526     \item Keep the new configuration and return to
527 xsun 3354 step~\ref{Initm:oldEnergy} if the random number smaller than
528 xsun 3347 $e^{-(U_{new} - U_{old})} / k_B T$.
529     \item Keep the old configuration and return to
530 xsun 3354 step~\ref{Initm:oldEnergy} if the random number larger than
531 xsun 3347 $e^{-(U_{new} - U_{old})} / k_B T$.
532     \end{itemize}
533     \end{itemize}
534 xsun 3354 \item\label{Initm:accumulateAvg} Accumulate the average after it converges.
535 xsun 3347 \end{enumerate}
536     It is important to notice that the old configuration has to be sampled
537     again if it is kept.
538    
539     \subsection{Molecular Dynamics\label{In:ssec:md}}
540     Although some of properites of the system can be calculated from the
541     ensemble average in Monte Carlo simulations, due to the nature of
542     lacking in the time dependence, it is impossible to gain information
543     of those dynamic properties from Monte Carlo simulations. Molecular
544     Dynamics is a measurement of the evolution of the positions and
545     momenta of the particles in the system. The evolution of the system
546     obeys laws of classical mechanics, in most situations, there is no
547     need for the count of the quantum effects. For a real experiment, the
548     instantaneous positions and momenta of the particles in the system are
549     neither important nor measurable, the observable quantities are
550     usually a average value for a finite time interval. These quantities
551     are expressed as a function of positions and momenta in Melecular
552     Dynamics simulations. Like the thermal temperature of the system is
553     defined by
554     \begin{equation}
555     \frac{1}{2} k_B T = \langle \frac{1}{2} m v_\alpha \rangle,
556 xsun 3354 \label{Ineq:temperature}
557 xsun 3347 \end{equation}
558     here $m$ is the mass of the particle and $v_\alpha$ is the $\alpha$
559     component of the velocity of the particle. The right side of
560 xsun 3354 Eq.~\ref{Ineq:temperature} is the average kinetic energy of the
561 xsun 3347 system. A simple Molecular Dynamics simulation scheme
562     is:~\cite{Frenkel1996}
563     \begin{enumerate}
564 xsun 3354 \item\label{Initm:initialize} Assign the initial positions and momenta
565 xsun 3347 for the particles in the system.
566 xsun 3354 \item\label{Initm:calcForce} Calculate the forces.
567     \item\label{Initm:equationofMotion} Integrate the equation of motion.
568 xsun 3347 \begin{itemize}
569 xsun 3354 \item Return to step~\ref{Initm:calcForce} if the equillibrium is
570 xsun 3347 not achieved.
571 xsun 3354 \item Go to step~\ref{Initm:calcAvg} if the equillibrium is
572 xsun 3347 achieved.
573     \end{itemize}
574 xsun 3354 \item\label{Initm:calcAvg} Compute the quantities we are interested in.
575 xsun 3347 \end{enumerate}
576     The initial positions of the particles are chosen as that there is no
577     overlap for the particles. The initial velocities at first are
578     distributed randomly to the particles, and then shifted to make the
579     momentum of the system $0$, at last scaled to the desired temperature
580 xsun 3354 of the simulation according Eq.~\ref{Ineq:temperature}.
581 xsun 3347
582 xsun 3354 The core of Molecular Dynamics simulations is step~\ref{Initm:calcForce}
583     and~\ref{Initm:equationofMotion}. The calculation of the forces are
584 xsun 3347 often involved numerous effort, this is the most time consuming step
585     in the Molecular Dynamics scheme. The evaluation of the forces is
586     followed by
587     \begin{equation}
588     f(q) = - \frac{\partial U(q)}{\partial q},
589 xsun 3354 \label{Ineq:force}
590 xsun 3347 \end{equation}
591     $U(q)$ is the potential of the system. Once the forces computed, are
592     the positions and velocities updated by integrating Newton's equation
593     of motion,
594     \begin{equation}
595     f(q) = \frac{dp}{dt} = \frac{m dv}{dt}.
596 xsun 3354 \label{Ineq:newton}
597 xsun 3347 \end{equation}
598     Here is an example of integrating algorithms, Verlet algorithm, which
599     is one of the best algorithms to integrate Newton's equation of
600     motion. The Taylor expension of the position at time $t$ is
601     \begin{equation}
602     q(t+\Delta t)= q(t) + v(t) \Delta t + \frac{f(t)}{2m}\Delta t^2 +
603     \frac{\Delta t^3}{3!}\frac{\partial^3 q(t)}{\partial t^3} +
604     \mathcal{O}(\Delta t^4)
605 xsun 3354 \label{Ineq:verletFuture}
606 xsun 3347 \end{equation}
607     for a later time $t+\Delta t$, and
608     \begin{equation}
609     q(t-\Delta t)= q(t) - v(t) \Delta t + \frac{f(t)}{2m}\Delta t^2 -
610     \frac{\Delta t^3}{3!}\frac{\partial^3 q(t)}{\partial t^3} +
611     \mathcal{O}(\Delta t^4) ,
612 xsun 3354 \label{Ineq:verletPrevious}
613 xsun 3347 \end{equation}
614     for a previous time $t-\Delta t$. The summation of the
615 xsun 3354 Eq.~\ref{Ineq:verletFuture} and~\ref{Ineq:verletPrevious} gives
616 xsun 3347 \begin{equation}
617     q(t+\Delta t)+q(t-\Delta t) =
618     2q(t) + \frac{f(t)}{m}\Delta t^2 + \mathcal{O}(\Delta t^4),
619 xsun 3354 \label{Ineq:verletSum}
620 xsun 3347 \end{equation}
621     so, the new position can be expressed as
622     \begin{equation}
623     q(t+\Delta t) \approx
624     2q(t) - q(t-\Delta t) + \frac{f(t)}{m}\Delta t^2.
625 xsun 3354 \label{Ineq:newPosition}
626 xsun 3347 \end{equation}
627     The higher order of the $\Delta t$ is omitted.
628    
629     Numerous technics and tricks are applied to Molecular Dynamics
630     simulation to gain more efficiency or more accuracy. The simulation
631     engine used in this dissertation for the Molecular Dynamics
632     simulations is {\sc oopse}, more details of the algorithms and
633     technics can be found in~\cite{Meineke2005}.
634    
635     \subsection{Langevin Dynamics\label{In:ssec:ld}}
636     In many cases, the properites of the solvent in a system, like the
637     lipid-water system studied in this dissertation, are less important to
638     the researchers. However, the major computational expense is spent on
639     the solvent in the Molecular Dynamics simulations because of the large
640     number of the solvent molecules compared to that of solute molecules,
641     the ratio of the number of lipid molecules to the number of water
642     molecules is $1:25$ in our lipid-water system. The efficiency of the
643     Molecular Dynamics simulations is greatly reduced.
644    
645     As an extension of the Molecular Dynamics simulations, the Langevin
646     Dynamics seeks a way to avoid integrating equation of motion for
647     solvent particles without losing the Brownian properites of solute
648     particles. A common approximation is that the coupling of the solute
649     and solvent is expressed as a set of harmonic oscillators. So the
650     Hamiltonian of such a system is written as
651     \begin{equation}
652     H = \frac{p^2}{2m} + U(q) + H_B + \Delta U(q),
653 xsun 3354 \label{Ineq:hamiltonianofCoupling}
654 xsun 3347 \end{equation}
655     where $H_B$ is the Hamiltonian of the bath which equals to
656     \begin{equation}
657     H_B = \sum_{\alpha = 1}^{N} \left\{ \frac{p_\alpha^2}{2m_\alpha} +
658     \frac{1}{2} m_\alpha \omega_\alpha^2 q_\alpha^2\right\},
659 xsun 3354 \label{Ineq:hamiltonianofBath}
660 xsun 3347 \end{equation}
661     $\alpha$ is all the degree of freedoms of the bath, $\omega$ is the
662     bath frequency, and $\Delta U(q)$ is the bilinear coupling given by
663     \begin{equation}
664     \Delta U = -\sum_{\alpha = 1}^{N} g_\alpha q_\alpha q,
665 xsun 3354 \label{Ineq:systemBathCoupling}
666 xsun 3347 \end{equation}
667     where $g$ is the coupling constant. By solving the Hamilton's equation
668     of motion, the {\it Generalized Langevin Equation} for this system is
669     derived to
670     \begin{equation}
671     m \ddot q = -\frac{\partial W(q)}{\partial q} - \int_0^t \xi(t) \dot q(t-t')dt' + R(t),
672 xsun 3354 \label{Ineq:gle}
673 xsun 3347 \end{equation}
674     with mean force,
675     \begin{equation}
676     W(q) = U(q) - \sum_{\alpha = 1}^N \frac{g_\alpha^2}{2m_\alpha
677     \omega_\alpha^2}q^2,
678 xsun 3354 \label{Ineq:meanForce}
679 xsun 3347 \end{equation}
680     being only a dependence of coordinates of the solute particles, {\it
681     friction kernel},
682     \begin{equation}
683     \xi(t) = \sum_{\alpha = 1}^N \frac{-g_\alpha}{m_\alpha
684     \omega_\alpha} \cos(\omega_\alpha t),
685 xsun 3354 \label{Ineq:xiforGLE}
686 xsun 3347 \end{equation}
687     and the random force,
688     \begin{equation}
689     R(t) = \sum_{\alpha = 1}^N \left( g_\alpha q_\alpha(0)-\frac{g_\alpha}{m_\alpha
690     \omega_\alpha^2}q(0)\right) \cos(\omega_\alpha t) + \frac{\dot
691     q_\alpha(0)}{\omega_\alpha} \sin(\omega_\alpha t),
692 xsun 3354 \label{Ineq:randomForceforGLE}
693 xsun 3347 \end{equation}
694     as only a dependence of the initial conditions. The relationship of
695     friction kernel $\xi(t)$ and random force $R(t)$ is given by
696     \begin{equation}
697     \xi(t) = \frac{1}{k_B T} \langle R(t)R(0) \rangle
698 xsun 3354 \label{Ineq:relationshipofXiandR}
699 xsun 3347 \end{equation}
700     from their definitions. In Langevin limit, the friction is treated
701     static, which means
702     \begin{equation}
703     \xi(t) = 2 \xi_0 \delta(t).
704 xsun 3354 \label{Ineq:xiofStaticFriction}
705 xsun 3347 \end{equation}
706 xsun 3354 After substitude $\xi(t)$ into Eq.~\ref{Ineq:gle} with
707     Eq.~\ref{Ineq:xiofStaticFriction}, {\it Langevin Equation} is extracted
708 xsun 3347 to
709     \begin{equation}
710     m \ddot q = -\frac{\partial U(q)}{\partial q} - \xi \dot q(t) + R(t).
711 xsun 3354 \label{Ineq:langevinEquation}
712 xsun 3347 \end{equation}
713     The applying of Langevin Equation to dynamic simulations is discussed
714     in Ch.~\ref{chap:ld}.