ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/xDissertation/Introduction.tex
(Generate patch)

Comparing trunk/xDissertation/Introduction.tex (file contents):
Revision 3336 by xsun, Wed Jan 30 16:01:02 2008 UTC vs.
Revision 3375 by xsun, Mon Mar 24 20:12:52 2008 UTC

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

Diff Legend

Removed lines
+ Added lines
< Changed lines
> Changed lines