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}. |
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}} |
17 |
> |
\caption[The chemical structure of lipids]{The chemical structure of |
18 |
> |
glycerophospholipids (left) and sphingophospholipids |
19 |
> |
(right).\cite{Cevc80}} |
20 |
|
\label{Infig:lipid} |
21 |
|
\end{figure} |
22 |
< |
The glycerophospholipid is the dominant phospholipid in membranes. The |
23 |
< |
types of glycerophospholipids depend on the X group, and the |
24 |
< |
chains. For example, if X is choline, the lipids are known as |
25 |
< |
phosphatidylcholine (PC), or if X is ethanolamine, the lipids are |
26 |
< |
known as phosphatidyethanolamine (PE). Table~\ref{Intab:pc} listed a |
27 |
< |
number types of phosphatidycholine with different fatty acids as the |
28 |
< |
lipid chains. |
22 |
> |
Glycerophospholipids are the dominant phospholipids in biological |
23 |
> |
membranes. The type of glycerophospholipid depends on the identity of |
24 |
> |
the X group, and the chains. For example, if X is choline |
25 |
> |
[(CH$_3$)$_3$N$^+$CH$_2$CH$_2$OH], the lipids are known as |
26 |
> |
phosphatidylcholine (PC), or if X is ethanolamine |
27 |
> |
[H$_3$N$^+$CH$_2$CH$_2$OH], the lipids are known as |
28 |
> |
phosphatidyethanolamine (PE). Table~\ref{Intab:pc} listed a number |
29 |
> |
types of phosphatidycholine with different fatty acids as the lipid |
30 |
> |
chains. |
31 |
|
\begin{table*} |
32 |
|
\begin{minipage}{\linewidth} |
33 |
|
\begin{center} |
34 |
< |
\caption{A number types of phosphatidycholine.} |
34 |
> |
\caption{A NUMBER TYPES OF PHOSPHATIDYCHOLINE} |
35 |
|
\begin{tabular}{lll} |
36 |
|
\hline |
37 |
|
& Fatty acid & Generic Name \\ |
46 |
|
\end{center} |
47 |
|
\end{minipage} |
48 |
|
\end{table*} |
49 |
< |
When dispersed in water, lipids self assemble into a mumber of |
49 |
> |
When dispersed in water, lipids self assemble into a number of |
50 |
|
topologically distinct bilayer structures. The phase behavior of lipid |
51 |
|
bilayers has been explored experimentally~\cite{Cevc80}, however, a |
52 |
|
complete understanding of the mechanism and driving forces behind the |
53 |
|
various phases has not been achieved. |
54 |
|
|
55 |
< |
\subsection{Ripple Phase\label{In:ssec:ripple}} |
55 |
> |
\subsection{The Ripple Phase\label{In:ssec:ripple}} |
56 |
|
The $P_{\beta'}$ {\it ripple phase} of lipid bilayers, named from the |
57 |
|
periodic buckling of the membrane, is an intermediate phase which is |
58 |
|
developed either from heating the gel phase $L_{\beta'}$ or cooling |
59 |
< |
the fluid phase $L_\alpha$. A Sketch is shown in |
59 |
> |
the fluid phase $L_\alpha$. A sketch of the phases is shown in |
60 |
|
figure~\ref{Infig:phaseDiagram}.~\cite{Cevc80} |
61 |
|
\begin{figure} |
62 |
|
\centering |
63 |
|
\includegraphics[width=\linewidth]{./figures/inPhaseDiagram.pdf} |
64 |
< |
\caption{A phase diagram of lipid bilayer. With increasing the |
65 |
< |
temperature, the bilayer can go through a gel ($L_{\beta'}$), ripple |
66 |
< |
($P_{\beta'}$) to fluid ($L_\alpha$) phase transition.~\cite{Cevc80}} |
64 |
> |
\caption[Phases of PC lipid bilayers]{Phases of PC lipid |
65 |
> |
bilayers. With increasing temperature, phosphotidylcholine (PC) |
66 |
> |
bilayers can go through $L_{\beta'} \rightarrow P_{\beta'}$ (gel |
67 |
> |
$\rightarrow$ ripple) and $P_{\beta'} \rightarrow L_\alpha$ (ripple |
68 |
> |
$\rightarrow$ fluid) phase transitions.~\cite{Cevc80}} |
69 |
|
\label{Infig:phaseDiagram} |
70 |
|
\end{figure} |
71 |
< |
Most structural information of the ripple phase has been obtained by |
72 |
< |
the X-ray diffraction~\cite{Sun96,Katsaras00} and freeze-fracture |
71 |
> |
Most structural information about the ripple phase has been obtained |
72 |
> |
by X-ray diffraction~\cite{Sun96,Katsaras00} and freeze-fracture |
73 |
|
electron microscopy (FFEM).~\cite{Copeland80,Meyer96} The X-ray |
74 |
|
diffraction work by Katsaras {\it et al.} showed that a rich phase |
75 |
|
diagram exhibiting both {\it asymmetric} and {\it symmetric} ripples |
80 |
|
\begin{figure} |
81 |
|
\centering |
82 |
|
\includegraphics[width=\linewidth]{./figures/inRipple.pdf} |
83 |
< |
\caption{The experimental observed ripple phase. The top image is |
84 |
< |
obtained by Sun {\it et al.} using X-ray diffraction~\cite{Sun96}, |
85 |
< |
and the bottom one is observed by Kaasgaard {\it et al.} using |
86 |
< |
AFM.~\cite{Kaasgaard03}} |
83 |
> |
\caption[Experimental observations of the riple phase]{Experimental |
84 |
> |
observations of the riple phase. The top image is an electrostatic |
85 |
> |
density map obtained by Sun {\it et al.} using X-ray |
86 |
> |
diffraction~\cite{Sun96}. The lower figures are the surface topology |
87 |
> |
of various ripple domains in bilayers supported in mica. The AFM |
88 |
> |
images were observed by Kaasgaard {\it et al.}.~\cite{Kaasgaard03}} |
89 |
|
\label{Infig:ripple} |
90 |
|
\end{figure} |
91 |
< |
Figure~\ref{Infig:ripple} shows the ripple phase oberved by X-ray |
91 |
> |
Figure~\ref{Infig:ripple} shows the ripple phase oberved by both X-ray |
92 |
|
diffraction and AFM. The experimental results provide strong support |
93 |
|
for a 2-dimensional triangular packing lattice of the lipid molecules |
94 |
|
within the ripple phase. This is a notable change from the observed |
95 |
|
lipid packing within the gel phase,~\cite{Cevc80} although Tenchov |
96 |
< |
{\it et al.} have recently observed near-hexagonal packing in some |
96 |
> |
{\it et al.} have recently observed near-triangular packing in some |
97 |
|
phosphatidylcholine (PC) gel phases.~\cite{Tenchov2001} However, the |
98 |
|
physical mechanism for the formation of the ripple phase has never |
99 |
|
been explained and the microscopic structure of the ripple phase has |
100 |
|
never been elucidated by experiments. Computational simulation is a |
101 |
< |
perfect tool to study the microscopic properties for a |
102 |
< |
system. However, the large length scale the ripple structure and the |
103 |
< |
long time scale of the formation of the ripples are crucial obstacles |
104 |
< |
to performing the actual work. The principal ideas explored in this |
105 |
< |
dissertation are attempts to break the computational task up by |
101 |
> |
very good tool to study the microscopic properties for a |
102 |
> |
system. However, the large length scale of the ripple structures and |
103 |
> |
the long time required for the formation of the ripples are crucial |
104 |
> |
obstacles to performing the actual work. The principal ideas explored |
105 |
> |
in this dissertation are attempts to break the computational task up |
106 |
> |
by |
107 |
|
\begin{itemize} |
108 |
|
\item Simplifying the lipid model. |
109 |
< |
\item Improving algorithm for integrating the equations of motion. |
109 |
> |
\item Improving the algorithm for integrating the equations of motion. |
110 |
|
\end{itemize} |
111 |
|
In chapters~\ref{chap:mc} and~\ref{chap:md}, we use a simple point |
112 |
< |
dipole spin model and a coarse-grained molecualr scale model to |
112 |
> |
dipole spin model and a coarse-grained molecular scale model to |
113 |
|
perform the Monte Carlo and Molecular Dynamics simulations |
114 |
|
respectively, and in chapter~\ref{chap:ld}, we develop a Langevin |
115 |
|
Dynamics algorithm which excludes the explicit solvent to improve the |
116 |
|
efficiency of the simulations. |
117 |
|
|
118 |
< |
\subsection{Lattice Model\label{In:ssec:model}} |
119 |
< |
The gel-like characteristic (relatively immobile molecules) exhibited |
118 |
> |
\subsection{Lattice Models\label{In:ssec:model}} |
119 |
> |
The gel-like characteristic (laterally immobile molecules) exhibited |
120 |
|
by the ripple phase makes it feasible to apply a lattice model to |
121 |
|
study the system. The popular $2$ dimensional lattice models, {\it |
122 |
|
i.e.}, the Ising, $X-Y$, and Heisenberg models, show {\it frustration} |
137 |
|
\begin{figure} |
138 |
|
\centering |
139 |
|
\includegraphics[width=3in]{./figures/inFrustration.pdf} |
140 |
< |
\caption{Frustration on triangular lattice, the spins and dipoles are |
141 |
< |
represented by arrows. The multiple local minima of energy states |
142 |
< |
induce the frustration for spins and dipoles picking the directions.} |
140 |
> |
\caption[Frustration on triangular lattice]{Frustration on triangular |
141 |
> |
lattice, the spins and dipoles are represented by arrows. The multiple |
142 |
> |
local minima of energy states induce frustration for spins and dipoles |
143 |
> |
resulting in disordered low-temperature phases.} |
144 |
|
\label{Infig:frustration} |
145 |
|
\end{figure} |
146 |
< |
The spins in figure~\ref{Infig:frustration} shows an illustration of |
147 |
< |
the frustration for $J < 0$ on a triangular lattice. There are |
148 |
< |
multiple local minima energy states which are independent of the |
149 |
< |
direction of the spin on top of the triangle, therefore infinite |
150 |
< |
possibilities for the packing of spins which induces what is known as |
151 |
< |
``complete regular frustration'' which leads to disordered low |
152 |
< |
temperature phases. The similarity goes to the dipoles on a hexagonal |
153 |
< |
lattice, which are shown by the dipoles in |
154 |
< |
figure~\ref{Infig:frustration}. In this circumstance, the dipoles want |
155 |
< |
to be aligned, however, due to the long wave fluctuation, at low |
156 |
< |
temperature, the aligned state becomes unstable, vortex is formed and |
157 |
< |
results in multiple local minima of energy states. The dipole on the |
148 |
< |
center of the hexagonal lattice is frustrated. |
146 |
> |
The spins in figure~\ref{Infig:frustration} illustrate frustration for |
147 |
> |
$J < 0$ on a triangular lattice. There are multiple local minima |
148 |
> |
energy states which are independent of the direction of the spin on |
149 |
> |
top of the triangle, therefore infinite possibilities for orienting |
150 |
> |
large numbers spins. This induces what is known as ``complete regular |
151 |
> |
frustration'' which leads to disordered low temperature phases. This |
152 |
> |
behavior extends to dipoles on a triangular lattices, which are shown |
153 |
> |
in the lower portion of figure~\ref{Infig:frustration}. In this case, |
154 |
> |
dipole-aligned structures are energetically favorable, however, at low |
155 |
> |
temperature, vortices are easily formed, and, this results in multiple |
156 |
> |
local minima of energy states for a central dipole. The dipole on the |
157 |
> |
center of the hexagonal lattice is therefore frustrated. |
158 |
|
|
159 |
|
The lack of translational degrees of freedom in lattice models |
160 |
|
prevents their utilization in models for surface buckling. In |
168 |
|
to key concepts of classical statistical mechanics that we used in |
169 |
|
this dissertation. Tolman gives an excellent introduction to the |
170 |
|
principles of statistical mechanics.~\cite{Tolman1979} A large part of |
171 |
< |
section~\ref{In:sec:SM} will follow Tolman's notation. |
171 |
> |
section~\ref{In:sec:SM} follows Tolman's notation. |
172 |
|
|
173 |
|
\subsection{Ensembles\label{In:ssec:ensemble}} |
174 |
|
In classical mechanics, the state of the system is completely |
278 |
|
and the rate of density change is zero in the neighborhood of any |
279 |
|
selected moving representative points in the phase space. |
280 |
|
|
281 |
< |
The condition of the ensemble is determined by the density |
281 |
> |
The type of thermodynamic ensemble is determined by the density |
282 |
|
distribution. If we consider the density distribution as only a |
283 |
|
function of $q$ and $p$, which means the rate of change of the phase |
284 |
|
space density in the neighborhood of all representative points in the |
307 |
|
\rho = \mathrm{const.} |
308 |
|
\label{Ineq:uniformEnsemble} |
309 |
|
\end{equation} |
310 |
< |
the ensemble is called {\it uniform ensemble}. |
310 |
> |
the ensemble is called {\it uniform ensemble}, but this ensemble has |
311 |
> |
little relevance for physical chemistry. It is an ensemble with |
312 |
> |
essentially infinite temperature. |
313 |
|
|
314 |
|
\subsubsection{The Microcanonical Ensemble\label{In:sssec:microcanonical}} |
315 |
< |
Another useful ensemble is the {\it microcanonical ensemble}, for |
316 |
< |
which: |
315 |
> |
The most useful ensemble for Molecular Dynamics is the {\it |
316 |
> |
microcanonical ensemble}, for which: |
317 |
|
\begin{equation} |
318 |
|
\rho = \delta \left( H(q^N, p^N) - E \right) \frac{1}{\Sigma (N, V, E)} |
319 |
|
\label{Ineq:microcanonicalEnsemble} |
330 |
|
\end{equation} |
331 |
|
where $k_B$ is the Boltzmann constant and $C^N$ is a number which |
332 |
|
makes the argument of $\ln$ dimensionless. In this case, $C^N$ is the |
333 |
< |
total phase space volume of one state. The entropy of a microcanonical |
334 |
< |
ensemble is given by |
333 |
> |
total phase space volume of one state which has the same units as |
334 |
> |
$\Sigma(N, V, E)$. The entropy of a microcanonical ensemble is given |
335 |
> |
by |
336 |
|
\begin{equation} |
337 |
|
S = k_B \ln \left(\frac{\Sigma(N, V, E)}{C^N}\right). |
338 |
|
\label{Ineq:entropy} |
349 |
|
Z_N = \int d \vec q~^N \int_\Gamma d \vec p~^N e^{-H(q^N, p^N) / k_B T}, |
350 |
|
\label{Ineq:partitionFunction} |
351 |
|
\end{equation} |
352 |
< |
which is also known as the canonical{\it partition function}. $\Gamma$ |
353 |
< |
indicates that the integral is over all phase space. In the canonical |
354 |
< |
ensemble, $N$, the total number of particles, $V$, total volume, and |
355 |
< |
$T$, the temperature, are constants. The systems with the lowest |
356 |
< |
energies hold the largest population. According to maximum principle, |
357 |
< |
thermodynamics maximizes the entropy $S$, implying that |
352 |
> |
which is also known as the canonical {\it partition |
353 |
> |
function}. $\Gamma$ indicates that the integral is over all phase |
354 |
> |
space. In the canonical ensemble, $N$, the total number of particles, |
355 |
> |
$V$, total volume, and $T$, the temperature, are constants. The |
356 |
> |
systems with the lowest energies hold the largest |
357 |
> |
population. Thermodynamics maximizes the entropy, $S$, implying that |
358 |
|
\begin{equation} |
359 |
|
\begin{array}{ccc} |
360 |
|
\delta S = 0 & \mathrm{and} & \delta^2 S < 0. |
375 |
|
system and the distribution of microscopic states. |
376 |
|
|
377 |
|
There is an implicit assumption that our arguments are based on so |
378 |
< |
far. A representative point in the phase space is equally likely to be |
379 |
< |
found in any energetically allowed region. In other words, all |
380 |
< |
energetically accessible states are represented equally, the |
381 |
< |
probabilities to find the system in any of the accessible states is |
382 |
< |
equal. This is called the principle of equal a {\it priori} |
378 |
> |
far. Tow representative points in phase space are equally likely to be |
379 |
> |
found if they have the same energy. In other words, all energetically |
380 |
> |
accessible states are represented , and the probabilities to find the |
381 |
> |
system in any of the accessible states is equal to that states |
382 |
> |
Boltzmann weight. This is called the principle of equal a {\it priori} |
383 |
|
probabilities. |
384 |
|
|
385 |
|
\subsection{Statistical Averages\label{In:ssec:average}} |
414 |
|
\frac{1}{T} \int_{0}^{T} F[q^N(t), p^N(t)] dt |
415 |
|
\label{Ineq:timeAverage2} |
416 |
|
\end{equation} |
417 |
< |
for an infinite time interval. |
417 |
> |
for an finite time interval, $T$. |
418 |
|
|
419 |
|
\subsubsection{Ergodicity\label{In:sssec:ergodicity}} |
420 |
|
The {\it ergodic hypothesis}, an important hypothesis governing modern |
457 |
|
\frac{1}{T} \int_{0}^{T} dt A(t) B(t + \tau), |
458 |
|
\label{Ineq:crosscorrelationFunction} |
459 |
|
\end{equation} |
460 |
< |
and called {\it cross correlation function}. |
460 |
> |
and is called a {\it cross correlation function}. |
461 |
|
|
462 |
|
We know from the ergodic hypothesis that there is a relationship |
463 |
|
between time average and ensemble average. We can put the correlation |
464 |
< |
function in a classical mechanics form, |
464 |
> |
function in a classical mechanical form, |
465 |
|
\begin{equation} |
466 |
< |
C_{AA}(\tau) = \int d \vec q~^N \int d \vec p~^N A[(q^N(t), p^N(t)] |
467 |
< |
A[q^N(t+\tau), p^N(t+\tau)] \rho(q, p) |
466 |
> |
C_{AA}(\tau) = \int d \vec q~^N \int d \vec p~^N A[(q^N, p^N] |
467 |
> |
A[q^N(\tau), p^N(\tau)] \rho(q^N, p^N) |
468 |
|
\label{Ineq:autocorrelationFunctionCM} |
469 |
|
\end{equation} |
470 |
< |
and |
470 |
> |
where $q^N(\tau)$, $p^N(\tau)$ is the phase space point that follows |
471 |
> |
classical evolution of the point $q^N$, $p^N$ after a tme $\tau$ has |
472 |
> |
elapsed, and |
473 |
|
\begin{equation} |
474 |
< |
C_{AB}(\tau) = \int d \vec q~^N \int d \vec p~^N A[(q^N(t), p^N(t)] |
475 |
< |
B[q^N(t+\tau), p^N(t+\tau)] \rho(q, p) |
474 |
> |
C_{AB}(\tau) = \int d \vec q~^N \int d \vec p~^N A[(q^N, p^N] |
475 |
> |
B[q^N(\tau), p^N(\tau)] \rho(q^N, p^N) |
476 |
|
\label{Ineq:crosscorrelationFunctionCM} |
477 |
|
\end{equation} |
478 |
|
as the autocorrelation function and cross correlation functions |
493 |
|
simulations. One is that the Monte Carlo simulations are time |
494 |
|
independent methods of sampling structural features of an ensemble, |
495 |
|
while Molecular Dynamics simulations provide dynamic |
496 |
< |
information. Additionally, Monte Carlo methods are a stochastic |
497 |
< |
processes, the future configurations of the system are not determined |
496 |
> |
information. Additionally, Monte Carlo methods are stochastic |
497 |
> |
processes; the future configurations of the system are not determined |
498 |
|
by its past. However, in Molecular Dynamics, the system is propagated |
499 |
< |
by Newton's equation of motion, and the trajectory of the system |
499 |
> |
by Hamilton's equations of motion, and the trajectory of the system |
500 |
|
evolving in phase space is deterministic. Brief introductions of the |
501 |
|
two algorithms are given in section~\ref{In:ssec:mc} |
502 |
|
and~\ref{In:ssec:md}. Langevin Dynamics, an extension of the Molecular |
507 |
|
A Monte Carlo integration algorithm was first introduced by Metropolis |
508 |
|
{\it et al.}~\cite{Metropolis53} The basic Metropolis Monte Carlo |
509 |
|
algorithm is usually applied to the canonical ensemble, a |
510 |
< |
Boltzmann-weighted ensemble, in which the $N$, the total number of |
511 |
< |
particles, $V$, total volume, $T$, temperature are constants. An |
512 |
< |
average in this ensemble is given |
510 |
> |
Boltzmann-weighted ensemble, in which $N$, the total number of |
511 |
> |
particles, $V$, the total volume, and $T$, the temperature are |
512 |
> |
constants. An average in this ensemble is given |
513 |
|
\begin{equation} |
514 |
|
\langle A \rangle = \frac{1}{Z_N} \int d \vec q~^N \int d \vec p~^N |
515 |
|
A(q^N, p^N) e^{-H(q^N, p^N) / k_B T}. |
533 |
|
\rangle$ is now a configuration integral, and |
534 |
|
Eq.~\ref{Ineq:configurationIntegral} is equivalent to |
535 |
|
\begin{equation} |
536 |
< |
\langle A \rangle = \int d \vec q~^N A \rho(q^N). |
536 |
> |
\langle A \rangle = \int d \vec q~^N A \rho(q^N), |
537 |
|
\label{Ineq:configurationAve} |
538 |
|
\end{equation} |
539 |
+ |
where $\rho(q^N)$ is a configurational probability |
540 |
+ |
\begin{equation} |
541 |
+ |
\rho(q^N) = \frac{e^{-U(q^N) / k_B T}}{\int d \vec q~^N e^{-U(q^N) / k_B T}}. |
542 |
+ |
\label{Ineq:configurationProb} |
543 |
+ |
\end{equation} |
544 |
|
|
545 |
|
In a Monte Carlo simulation of a system in the canonical ensemble, the |
546 |
|
probability of the system being in a state $s$ is $\rho_s$, the change |
555 |
|
\frac{d \rho_{s}^{equilibrium}}{dt} = 0, |
556 |
|
\label{Ineq:equiProb} |
557 |
|
\end{equation} |
558 |
< |
which means $\sum_{s'} [ -w_{ss'}\rho_s + w_{s's}\rho_{s'} ]$ is $0$ |
559 |
< |
for all $s'$. So |
558 |
> |
the sum of transition probabilities $\sum_{s'} [ -w_{ss'}\rho_s + |
559 |
> |
w_{s's}\rho_{s'} ]$ is $0$ for all $s'$. So |
560 |
|
\begin{equation} |
561 |
|
\frac{\rho_s^{equilibrium}}{\rho_{s'}^{equilibrium}} = \frac{w_{s's}}{w_{ss'}}. |
562 |
|
\label{Ineq:relationshipofRhoandW} |
571 |
|
\frac{w_{s's}}{w_{ss'}} = e^{-(U_s - U_{s'}) / k_B T}, |
572 |
|
\label{Ineq:conditionforBoltzmannStatistics} |
573 |
|
\end{equation} |
574 |
< |
An algorithm that shows how Monte Carlo simulation generates a |
574 |
> |
An algorithm that indicates how a Monte Carlo simulation generates a |
575 |
|
transition probability governed by |
576 |
|
\ref{Ineq:conditionforBoltzmannStatistics}, is given schematically as, |
577 |
|
\begin{enumerate} |
578 |
|
\item\label{Initm:oldEnergy} Choose a particle randomly, and calculate |
579 |
< |
the energy of the rest of the system due to the location of the particle. |
579 |
> |
the energy of the rest of the system due to the current location of |
580 |
> |
the particle. |
581 |
|
\item\label{Initm:newEnergy} Make a random displacement of the particle, |
582 |
|
calculate the new energy taking the movement of the particle into account. |
583 |
|
\begin{itemize} |
584 |
< |
\item If the energy goes down, keep the new configuration and return to |
565 |
< |
step~\ref{Initm:oldEnergy}. |
584 |
> |
\item If the energy goes down, keep the new configuration. |
585 |
|
\item If the energy goes up, pick a random number between $[0,1]$. |
586 |
|
\begin{itemize} |
587 |
|
\item If the random number smaller than |
588 |
< |
$e^{-(U_{new} - U_{old})} / k_B T$, keep the new configuration and return to |
570 |
< |
step~\ref{Initm:oldEnergy}. |
588 |
> |
$e^{-(U_{new} - U_{old})} / k_B T$, keep the new configuration. |
589 |
|
\item If the random number is larger than |
590 |
< |
$e^{-(U_{new} - U_{old})} / k_B T$, keep the old configuration and return to |
573 |
< |
step~\ref{Initm:oldEnergy}. |
590 |
> |
$e^{-(U_{new} - U_{old})} / k_B T$, keep the old configuration. |
591 |
|
\end{itemize} |
592 |
|
\end{itemize} |
593 |
|
\item\label{Initm:accumulateAvg} Accumulate the averages based on the |
594 |
< |
current configuartion. |
594 |
> |
current configuration. |
595 |
|
\item Go to step~\ref{Initm:oldEnergy}. |
596 |
|
\end{enumerate} |
597 |
< |
It is important for sampling accurately that the old configuration is |
597 |
> |
It is important for sampling accuracy that the old configuration is |
598 |
|
sampled again if it is kept. |
599 |
|
|
600 |
|
\subsection{Molecular Dynamics\label{In:ssec:md}} |
606 |
|
evolution of the system obeys the laws of classical mechanics, and in |
607 |
|
most situations, there is no need to account for quantum effects. In a |
608 |
|
real experiment, the instantaneous positions and momenta of the |
609 |
< |
particles in the system are neither important nor measurable, the |
610 |
< |
observable quantities are usually an average value for a finite time |
611 |
< |
interval. These quantities are expressed as a function of positions |
612 |
< |
and momenta in Molecular Dynamics simulations. For example, |
609 |
> |
particles in the system are ofter neither important nor measurable, |
610 |
> |
the observable quantities are usually an average value for a finite |
611 |
> |
time interval. These quantities are expressed as a function of |
612 |
> |
positions and momenta in Molecular Dynamics simulations. For example, |
613 |
|
temperature of the system is defined by |
614 |
|
\begin{equation} |
615 |
|
\frac{3}{2} N k_B T = \langle \sum_{i=1}^N \frac{1}{2} m_i v_i \rangle, |
622 |
|
The initial positions of the particles are chosen so that there is no |
623 |
|
overlap of the particles. The initial velocities at first are |
624 |
|
distributed randomly to the particles using a Maxwell-Boltzmann |
625 |
< |
ditribution, and then shifted to make the total linear momentum of the |
626 |
< |
system $0$. |
625 |
> |
distribution, and then shifted to make the total linear momentum of |
626 |
> |
the system $0$. |
627 |
|
|
628 |
|
The core of Molecular Dynamics simulations is the calculation of |
629 |
|
forces and the integration algorithm. Calculation of the forces often |
649 |
|
equations of motion. The Taylor expansion of the position at time $t$ |
650 |
|
is |
651 |
|
\begin{equation} |
652 |
< |
q(t+\Delta t)= q(t) + v(t) \Delta t + \frac{f(t)}{2m}\Delta t^2 + |
652 |
> |
q(t+\Delta t)= q(t) + \frac{p(t)}{m} \Delta t + \frac{f(t)}{2m}\Delta t^2 + |
653 |
|
\frac{\Delta t^3}{3!}\frac{\partial^3 q(t)}{\partial t^3} + |
654 |
|
\mathcal{O}(\Delta t^4) |
655 |
|
\label{Ineq:verletFuture} |
656 |
|
\end{equation} |
657 |
|
for a later time $t+\Delta t$, and |
658 |
|
\begin{equation} |
659 |
< |
q(t-\Delta t)= q(t) - v(t) \Delta t + \frac{f(t)}{2m}\Delta t^2 - |
659 |
> |
q(t-\Delta t)= q(t) - \frac{p(t)}{m} \Delta t + \frac{f(t)}{2m}\Delta t^2 - |
660 |
|
\frac{\Delta t^3}{3!}\frac{\partial^3 q(t)}{\partial t^3} + |
661 |
|
\mathcal{O}(\Delta t^4) , |
662 |
|
\label{Ineq:verletPrevious} |
683 |
|
code can be found in Ref.~\cite{Meineke2005}. |
684 |
|
|
685 |
|
\subsection{Langevin Dynamics\label{In:ssec:ld}} |
686 |
< |
In many cases, the properites of the solvent in a system, like the |
687 |
< |
water in the lipid-water system studied in this dissertation, are less |
688 |
< |
interesting to the researchers than the behavior of the |
689 |
< |
solute. However, the major computational expense is ofter the |
690 |
< |
solvent-solvent interation, this is due to the large number of the |
691 |
< |
solvent molecules when compared to the number of solute molecules, the |
692 |
< |
ratio of the number of lipid molecules to the number of water |
693 |
< |
molecules is $1:25$ in our lipid-water system. The efficiency of the |
694 |
< |
Molecular Dynamics simulations is greatly reduced, with up to 85\% of |
695 |
< |
CPU time spent calculating only water-water interactions. |
686 |
> |
In many cases, the properites of the solvent (like the water in the |
687 |
> |
lipid-water system studied in this dissertation) are less interesting |
688 |
> |
to the researchers than the behavior of the solute. However, the major |
689 |
> |
computational expense is ofter the solvent-solvent interactions, this |
690 |
> |
is due to the large number of the solvent molecules when compared to |
691 |
> |
the number of solute molecules. The ratio of the number of lipid |
692 |
> |
molecules to the number of water molecules is $1:25$ in our |
693 |
> |
lipid-water system. The efficiency of the Molecular Dynamics |
694 |
> |
simulations is greatly reduced, with up to 85\% of CPU time spent |
695 |
> |
calculating only water-water interactions. |
696 |
|
|
697 |
|
As an extension of the Molecular Dynamics methodologies, Langevin |
698 |
|
Dynamics seeks a way to avoid integrating the equations of motion for |
745 |
|
q_\alpha(0)}{\omega_\alpha} \sin(\omega_\alpha t), |
746 |
|
\label{Ineq:randomForceforGLE} |
747 |
|
\end{equation} |
748 |
< |
depends only on the initial conditions. The relationship of friction |
749 |
< |
kernel $\xi(t)$ and random force $R(t)$ is given by the second |
750 |
< |
fluctuation dissipation theorem, |
748 |
> |
that depends only on the initial conditions. The relationship of |
749 |
> |
friction kernel $\xi(t)$ and random force $R(t)$ is given by the |
750 |
> |
second fluctuation dissipation theorem, |
751 |
|
\begin{equation} |
752 |
|
\xi(t) = \frac{1}{k_B T} \langle R(t)R(0) \rangle. |
753 |
|
\label{Ineq:relationshipofXiandR} |