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 3365 by xsun, Mon Mar 10 21:52:50 2008 UTC vs.
Revision 3372 by xsun, Wed Mar 19 21:04:47 2008 UTC

# Line 3 | Line 3 | because of their critical role as the foundation of bi
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. Lipids, when dispersed in water, self assemble into bilayer
7 < structures. The phase behavior of lipid bilayers have been explored
6 > 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   experimentally~\cite{Cevc87}, however, a complete understanding of the
10 < mechanism for self-assembly has not been achieved.
10 > mechanism and driving forces behind the various phases has not been
11 > achieved.
12  
13   \subsection{Ripple Phase\label{In:ssec:ripple}}
14   The $P_{\beta'}$ {\it ripple phase} of lipid bilayers, named from the
15   periodic buckling of the membrane, is an intermediate phase which is
16   developed either from heating the gel phase $L_{\beta'}$ or cooling
17 < the fluid phase $L_\alpha$. Although the ripple phase has been
18 < observed in several
19 < experiments~\cite{Sun96,Katsaras00,Copeland80,Meyer96,Kaasgaard03},
20 < the physical mechanism for the formation of the ripple phase has never been
21 < explained and the microscopic structure of the ripple phase has never
22 < been elucidated by experiments. Computational simulation is a perfect
23 < tool to study the microscopic properties for a system. However, the
24 < large length scale the ripple structure and the long time scale
25 < of the formation of the ripples are crucial obstacles to performing
26 < the actual work. The principal ideas explored in this dissertation are
27 < attempts to break the computational task up by
17 > 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   \begin{itemize}
60   \item Simplifying the lipid model.
61   \item Improving algorithm for integrating the equations of motion.
62   \end{itemize}
63   In chapters~\ref{chap:mc} and~\ref{chap:md}, we use a simple point
64 < dipole model and a coarse-grained model to perform the Monte Carlo and
65 < Molecular Dynamics simulations respectively, and in
66 < chapter~\ref{chap:ld}, we develop a Langevin Dynamics algorithm which
67 < excludes the explicit solvent to improve the efficiency of the
68 < simulations.
64 > 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  
70   \subsection{Lattice Model\label{In:ssec:model}}
71 < The gel-like characteristic of the ripple phase makes it feasible to
72 < apply a lattice model to study the system. It is claimed that the
73 < packing of the lipid molecules in ripple phase is
74 < hexagonal~\cite{Cevc87}. The popular $2$ dimensional lattice models,
75 < {\it i.e.}, the Ising, $X-Y$, and Heisenberg models, show {\it
43 < frustration} on triangular lattices. The Hamiltonian of the systems
44 < are given by
71 > 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   \begin{equation}
77   H =
78    \begin{cases}
# Line 51 | Line 82 | where $J$ has non zero value only when spins $s_n$ ($\
82    \end{cases}
83   \end{equation}
84   where $J$ has non zero value only when spins $s_n$ ($\vec s_n$) and
85 < $s_{n'}$ ($\vec s_{n'}$) are the nearest neighbours.
85 > $s_{n'}$ ($\vec s_{n'}$) are nearest neighbors. When $J > 0$, spins
86 > prefer an aligned structure, and if $J < 0$, spins prefer an
87 > anti-aligned structure.
88 >
89   \begin{figure}
90   \centering
91 < \includegraphics[width=\linewidth]{./figures/inFrustration.pdf}
92 < \caption{Frustration on a triangular lattice, the spins are
93 < represented by arrows. No matter which direction the spin on the top
94 < of triangle points to, the Hamiltonain of the system goes up.}
91 > \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   \label{Infig:frustration}
96   \end{figure}
97 < Figure~\ref{Infig:frustration} shows an illustration of the
98 < frustration on a triangular lattice. When $J < 0$, the spins want to
99 < be anti-aligned, The direction of the spin on top of the triangle will
100 < make the energy go up no matter which direction it picks, therefore
101 < infinite possibilities for the packing of spins induce what is known
102 < as ``complete regular frustration'' which leads to disordered low
103 < temperature phases.
97 > 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 > possibilities for the packing of spins which induces what is known as
102 > ``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  
111 < The lack of translational degree of freedom in lattice models prevents
112 < their utilization in models for surface buckling which would
113 < correspond to ripple formation. In chapter~\ref{chap:mc}, a modified
114 < lattice model is introduced to tackle this specific situation.
111 > 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  
116   \section{Overview of Classical Statistical Mechanics\label{In:sec:SM}}
117   Statistical mechanics provides a way to calculate the macroscopic
# Line 79 | Line 119 | this dissertation. Tolman gives an excellent introduct
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 < principles of statistical mechanics~\cite{Tolman1979}. A large part of
122 > principles of statistical mechanics.~\cite{Tolman1979} A large part of
123   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 < $N$ particle system, there are $6N$ coordinates ($3N$ position $(q_1,
128 > $N$ particle system, there are $6N$ coordinates ($3N$ positions $(q_1,
129   q_2, \ldots, q_{3N})$ 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 $6N$
# Line 110 | Line 150 | then the value of $\rho$ gives the probability of find
150   \label{Ineq:normalized}
151   \end{equation}
152   then the value of $\rho$ gives the probability of finding the system
153 < in a unit volume in the phase space.
153 > in a unit volume in phase space.
154  
155   Liouville's theorem describes the change in density $\rho$ with
156 < time. The number of representive points at a given volume in the phase
156 > time. The number of representative points at a given volume in phase
157   space at any instant can be written as:
158   \begin{equation}
159   \label{Ineq:deltaN}
160   \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 < To calculate the change in the number of representive points in this
162 > To calculate the change in the number of representative points in this
163   volume, let us consider a simple condition: the change in the number
164 < of representive points in $q_1$ axis. The rate of the number of the
165 < representive points entering the volume at $q_1$ per unit time is:
164 > 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   \begin{equation}
168   \label{Ineq:deltaNatq1}
169   \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 < and the rate of the number of representive points leaving the volume
171 > and the rate of the number of representative points leaving the volume
172   at another position $q_1 + \delta q_1$ is:
173   \begin{equation}
174   \label{Ineq:deltaNatq1plusdeltaq1}
175   \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 < Here the higher order differentials are neglected. So the change of
179 < the number of the representive points is the difference of
180 < eq.~\ref{Ineq:deltaNatq1} and eq.~\ref{Ineq:deltaNatq1plusdeltaq1}, which
181 < gives us:
178 > 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   \begin{equation}
183   \label{Ineq:deltaNatq1axis}
184   -\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 < axes in the phase space, we can get the change of the number of
188 < representive points in a given volume with time:
187 > axes in the phase space, we can get the change in the number of
188 > representative points in a given volume with time:
189   \begin{equation}
190   \label{Ineq:deltaNatGivenVolume}
191   \frac{d(\delta N)}{dt} = -\sum_{i=1}^N \left[\rho \left(\frac{\partial
# Line 152 | Line 193 | q_i} \dot q_i + \frac{\partial {\rho}}{\partial p_i} \
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 < From Hamilton's equation of motion,
196 > From Hamilton's equations of motion,
197   \begin{equation}
198   \frac{\partial {\dot q_i}}{\partial q_i} = - \frac{\partial
199   {\dot p_i}}{\partial p_i},
# Line 162 | Line 203 | eq.~\ref{Ineq:deltaNatGivenVolume} are divided by $\de
203   eq.~\ref{Ineq:deltaNatGivenVolume}. If both sides of
204   eq.~\ref{Ineq:deltaNatGivenVolume} are divided by $\delta q_1 \delta q_2
205   \ldots \delta q_N \delta p_1 \delta p_2 \ldots \delta p_N$, then we
206 < can derive Liouville's theorem:
206 > arrive at Liouville's theorem:
207   \begin{equation}
208   \left( \frac{\partial \rho}{\partial t} \right)_{q, p} = -\sum_{i} \left(
209   \frac{\partial {\rho}}{\partial
# Line 181 | Line 222 | equation~\ref{Ineq:anotherFormofLiouville} is the tota
222   \end{equation}
223   It is easy to note that the left side of
224   equation~\ref{Ineq:anotherFormofLiouville} is the total derivative of
225 < $\rho$ with respect of $t$, which means
225 > $\rho$ with respect to $t$, which means
226   \begin{equation}
227   \frac{d \rho}{dt} = 0,
228   \label{Ineq:conservationofRho}
229   \end{equation}
230   and the rate of density change is zero in the neighborhood of any
231 < selected moving representive points in the phase space.
231 > selected moving representative points in the phase space.
232  
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 < space density in the neighborhood of all representive points in the
236 > space density in the neighborhood of all representative points in the
237   phase space is zero,
238   \begin{equation}
239   \left( \frac{\partial \rho}{\partial t} \right)_{q, p} = 0.
240   \label{Ineq:statEquilibrium}
241   \end{equation}
242   We may conclude the ensemble is in {\it statistical equilibrium}. An
243 < ensemble in statistical equilibrium often means the system is also in
243 > ensemble in statistical equilibrium means the system is also in
244   macroscopic equilibrium. If $\left( \frac{\partial \rho}{\partial t}
245   \right)_{q, p} = 0$, then
246   \begin{equation}
# Line 211 | Line 252 | independent of time. For a conservative system, the en
252   \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 < system is one of the constants of the motion. Here are several
256 < examples: when the density distribution is constant everywhere in the
257 < phase space,
255 > 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   \begin{equation}
259   \rho = \mathrm{const.}
260   \label{Ineq:uniformEnsemble}
261   \end{equation}
262 < the ensemble is called {\it uniform ensemble}.  Another useful
263 < ensemble is called {\it microcanonical ensemble}, for which:
262 > 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   \begin{equation}
268   \rho = \delta \left( H(q^N, p^N) - E \right) \frac{1}{\Sigma (N, V, E)}
269   \label{Ineq:microcanonicalEnsemble}
# Line 235 | Line 279 | where $k_B$ is the Boltzmann constant and $C^N$ is a n
279   \label{Ineq:gibbsEntropy}
280   \end{equation}
281   where $k_B$ is the Boltzmann constant and $C^N$ is a number which
282 < makes the argument of $\ln$ dimensionless, in this case, it is the
283 < total phase space volume of one state. The entropy in microcanonical
282 > 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   ensemble is given by
285   \begin{equation}
286   S = k_B \ln \left(\frac{\Sigma(N, V, E)}{C^N}\right).
287   \label{Ineq:entropy}
288   \end{equation}
289 +
290 + \subsubsection{The Canonical Ensemble\label{In:sssec:canonical}}
291   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},
# Line 252 | Line 298 | Z_N = \int d \vec q~^N \int_\Gamma d \vec p~^N  e^{-H(
298   Z_N = \int d \vec q~^N \int_\Gamma d \vec p~^N  e^{-H(q^N, p^N) / k_B T},
299   \label{Ineq:partitionFunction}
300   \end{equation}
301 < which is also known as {\it partition function}. $\Gamma$ indicates
302 < that the integral is over all the phase space. In the canonical
301 > 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   ensemble, $N$, the total number of particles, $V$, total volume, and
304 < $T$, the temperature are constants. The systems with the lowest
304 > $T$, the temperature, are constants. The systems with the lowest
305   energies hold the largest population. According to maximum principle,
306 < the thermodynamics maximizes the entropy $S$,
306 > thermodynamics maximizes the entropy $S$, implying that
307   \begin{equation}
308   \begin{array}{ccc}
309   \delta S = 0 & \mathrm{and} & \delta^2 S < 0.
310   \end{array}
311   \label{Ineq:maximumPrinciple}
312   \end{equation}
313 < From Eq.~\ref{Ineq:maximumPrinciple} and two constrains of the canonical
314 < ensemble, {\it i.e.}, total probability and average energy conserved,
315 < the partition function is calculated as
313 > 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   \begin{equation}
318   Z_N = e^{-A/k_B T},
319   \label{Ineq:partitionFunctionWithFreeEnergy}
# Line 274 | Line 321 | that they serve as a connection between macroscopic pr
321   where $A$ is the Helmholtz free energy. The significance of
322   Eq.~\ref{Ineq:entropy} and~\ref{Ineq:partitionFunctionWithFreeEnergy} is
323   that they serve as a connection between macroscopic properties of the
324 < system and the distribution of the microscopic states.
324 > system and the distribution of microscopic states.
325  
326   There is an implicit assumption that our arguments are based on so
327 < far. A representive point in the phase space is equally to be found in
328 < any same extent of the regions. In other words, all energetically
329 < accessible states are represented equally, the probabilities to find
330 < the system in any of the accessible states is equal. This is called
331 < equal a {\it priori} probabilities.
327 > 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  
334 < \subsection{Statistical Average\label{In:ssec:average}}
335 < Given the density distribution $\rho$ in the phase space, the average
334 > \subsection{Statistical Averages\label{In:ssec:average}}
335 > Given a density distribution $\rho$ in phase space, the average
336   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   Eq.~\ref{Ineq:statAverage1}
340   \begin{equation}
341 < \langle F(q^N, p^N, t) \rangle = \frac{\int d \vec q~^N \int d \vec p~^N
342 < F(q^N, p^N, t) \rho}{\int d \vec q~^N \int d \vec p~^N \rho}.
341 > \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   \label{Ineq:statAverage1}
344   \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 < \langle F(q^N, p^N, t) \rangle = \int d \vec q~^N \int d \vec p~^N F(q^N,
349 < p^N, t) \rho,
348 > \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   \label{Ineq:statAverage2}
351   \end{equation}
352 < called {\it ensemble average}. However, the quantity is often averaged
353 < for a finite time in real experiments,
352 > called the {\it ensemble average}. However, the quantity is often
353 > averaged for a finite time in real experiments,
354   \begin{equation}
355   \langle F(q^N, p^N) \rangle_t = \lim_{T \rightarrow \infty}
356 < \frac{1}{T} \int_{t_0}^{t_0+T} F(q^N, p^N, t) dt.
356 > \frac{1}{T} \int_{t_0}^{t_0+T} F[q^N(t), p^N(t)] dt.
357   \label{Ineq:timeAverage1}
358   \end{equation}
359   Usually this time average is independent of $t_0$ in statistical
360   mechanics, so Eq.~\ref{Ineq:timeAverage1} becomes
361   \begin{equation}
362   \langle F(q^N, p^N) \rangle_t = \lim_{T \rightarrow \infty}
363 < \frac{1}{T} \int_{0}^{T} F(q^N, p^N, t) dt
363 > \frac{1}{T} \int_{0}^{T} F[q^N(t), p^N(t)] dt
364   \label{Ineq:timeAverage2}
365   \end{equation}
366   for an infinite time interval.
367  
368 < {\it ergodic hypothesis}, an important hypothesis from the statistical
369 < mechanics point of view, states that the system will eventually pass
368 > \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   arbitrarily close to any point that is energetically accessible in
372   phase space. Mathematically, this leads to
373   \begin{equation}
374 < \langle F(q^N, p^N, t) \rangle = \langle F(q^N, p^N) \rangle_t.
374 > \langle F(q^N, p^N) \rangle = \langle F(q^N, p^N) \rangle_t.
375   \label{Ineq:ergodicity}
376   \end{equation}
377 < Eq.~\ref{Ineq:ergodicity} validates the Monte Carlo method which we will
378 < discuss in section~\ref{In:ssec:mc}. An ensemble average of a quantity
379 < can be related to the time average measured in the experiments.
377 > 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  
382 < \subsection{Correlation Function\label{In:ssec:corr}}
383 < Thermodynamic properties can be computed by equillibrium statistical
384 < mechanics. On the other hand, {\it Time correlation function} is a
385 < powerful method to understand the evolution of a dynamic system in
386 < non-equillibrium statistical mechanics. Imagine a property $A(q^N,
387 < p^N, t)$ as a function of coordinates $q^N$ and momenta $p^N$ has an
388 < intial value at $t_0$, at a later time $t_0 + \tau$ this value is
389 < changed. If $\tau$ is very small, the change of the value is minor,
390 < and the later value of $A(q^N, p^N, t_0 +
391 < \tau)$ is correlated to its initial value. Howere, when $\tau$ is large,
392 < this correlation is lost. The correlation function is a measurement of
393 < this relationship and is defined by~\cite{Berne90}
382 > \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   \begin{equation}
395 < C(t) = \langle A(0)A(\tau) \rangle = \lim_{T \rightarrow \infty}
395 > C_{AA}(\tau) = \langle A(0)A(\tau) \rangle = \lim_{T \rightarrow
396 > \infty}
397   \frac{1}{T} \int_{0}^{T} dt A(t) A(t + \tau).
398   \label{Ineq:autocorrelationFunction}
399   \end{equation}
400 < Eq.~\ref{Ineq:autocorrelationFunction} is the correlation function of a
401 < single variable, called {\it autocorrelation function}. The defination
402 < of the correlation function for two different variables is similar to
403 < that of autocorrelation function, which is
400 > 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   \begin{equation}
405 < C(t) = \langle A(0)B(\tau) \rangle = \lim_{T \rightarrow \infty}
405 > C_{AB}(\tau) = \langle A(0)B(\tau) \rangle = \lim_{T \rightarrow \infty}
406   \frac{1}{T} \int_{0}^{T} dt A(t) B(t + \tau),
407   \label{Ineq:crosscorrelationFunction}
408   \end{equation}
409   and called {\it cross correlation function}.
410  
411 < In section~\ref{In:ssec:average} we know from Eq.~\ref{Ineq:ergodicity}
412 < the relationship between time average and ensemble average. We can put
413 < the correlation function in a classical mechanics form,
411 > 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   \begin{equation}
415 < C(t) = \langle A(0)A(\tau) \rangle = \int d \vec q~^N \int d \vec p~^N A(t) A(t + \tau) \rho(q, p)
415 > 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   \label{Ineq:autocorrelationFunctionCM}
418   \end{equation}
419   and
420   \begin{equation}
421 < C(t) = \langle A(0)B(\tau) \rangle = \int d \vec q~^N \int d \vec p~^N A(t) B(t + \tau)
422 < \rho(q, p)
421 > 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   \label{Ineq:crosscorrelationFunctionCM}
424   \end{equation}
425   as autocorrelation function and cross correlation function

Diff Legend

Removed lines
+ Added lines
< Changed lines
> Changed lines