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 3347 by xsun, Fri Feb 29 15:20:55 2008 UTC vs.
Revision 3371 by xsun, Fri Mar 14 23:16:11 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 chosen to be studied in this dissertation
5 < because of their critical role as a foundation of the bio-membrane
6 < construction. The self assembled bilayer of the lipids when dispersed
7 < in water is the micro structure of the membrane. The phase behavior of
8 < lipid bilayer is explored experimentally~\cite{Cevc87}, however, fully
9 < understanding on the mechanism is far beyond accomplished.
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
8 > experimentally~\cite{Cevc87}, however, a complete understanding of the
9 > mechanism for self-assembly has not been achieved.
10  
11   \subsection{Ripple Phase\label{In:ssec:ripple}}
12 < The {\it ripple phase} $P_{\beta'}$ of lipid bilayers, named from the
12 > The $P_{\beta'}$ {\it ripple phase} of lipid bilayers, named from the
13   periodic buckling of the membrane, is an intermediate phase which is
14   developed either from heating the gel phase $L_{\beta'}$ or cooling
15 < the fluid phase $L_\alpha$. Although the ripple phase is observed in
16 < different
15 > the fluid phase $L_\alpha$. Although the ripple phase has been
16 > observed in several
17   experiments~\cite{Sun96,Katsaras00,Copeland80,Meyer96,Kaasgaard03},
18 < the mechanism of the formation of the ripple phase has never been
18 > the physical mechanism for the formation of the ripple phase has never been
19   explained and the microscopic structure of the ripple phase has never
20   been elucidated by experiments. Computational simulation is a perfect
21 < tool to study the microscopic properties for a system, however, the
22 < long range dimension of the ripple structure and the long time scale
21 > tool to study the microscopic properties for a system. However, the
22 > large length scale the ripple structure and the long time scale
23   of the formation of the ripples are crucial obstacles to performing
24 < the actual work. The idea to break through this dilemma forks into:
24 > the actual work. The principal ideas explored in this dissertation are
25 > attempts to break the computational task up by
26   \begin{itemize}
27 < \item Simplify the lipid model.
28 < \item Improve the integrating algorithm.
27 > \item Simplifying the lipid model.
28 > \item Improving algorithm for integrating the equations of motion.
29   \end{itemize}
30 < In Ch.~\ref{chap:mc} and~\ref{chap:md}, we use a simple point dipole
31 < model and a coarse-grained model to perform the Monte Carlo and
32 < Molecular Dynamics simulations respectively, and in Ch.~\ref{chap:ld},
33 < we implement a Langevin Dynamics algorithm to exclude the explicit
34 < solvent to improve the efficiency of the simulations.
30 > In chapters~\ref{chap:mc} and~\ref{chap:md}, we use a simple point
31 > dipole model and a coarse-grained model to perform the Monte Carlo and
32 > Molecular Dynamics simulations respectively, and in
33 > chapter~\ref{chap:ld}, we develop a Langevin Dynamics algorithm which
34 > excludes the explicit solvent to improve the efficiency of the
35 > simulations.
36  
37   \subsection{Lattice Model\label{In:ssec:model}}
38 < The gel like characteristic of the ripple phase ensures the feasiblity
39 < of applying the lattice model to study the system. It is claimed that
40 < the packing of the lipid molecules in ripple phase is
38 > The gel-like characteristic of the ripple phase makes it feasible to
39 > apply a lattice model to study the system. It is claimed that the
40 > packing of the lipid molecules in ripple phase is
41   hexagonal~\cite{Cevc87}. The popular $2$ dimensional lattice models,
42 < {\it i.e.}, Ising model, Heisenberg model and $X-Y$ model, show
43 < {\it frustration} on triangular lattice.
42 > {\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
45 > \begin{equation}
46 > H =
47 >  \begin{cases}
48 >    -J \sum_n \sum_{n'} s_n s_n' & \text{Ising}, \\
49 >    -J \sum_n \sum_{n'} \vec s_n \cdot \vec s_{n'} & \text{$X-Y$ and
50 > Heisenberg},
51 >  \end{cases}
52 > \end{equation}
53 > where $J$ has non zero value only when spins $s_n$ ($\vec s_n$) and
54 > $s_{n'}$ ($\vec s_{n'}$) are the nearest neighbours. When $J > 0$, the
55 > spins prefer aligned with each other, and if $J < 0$, the spins want
56 > to be anti-aligned.
57 >
58   \begin{figure}
59   \centering
60 < \includegraphics[width=\linewidth]{./figures/frustration.pdf}
61 < \caption{Sketch to illustrate the frustration on triangular
62 < lattice. Spins are represented by arrows, no matter which direction
63 < the spin on the top of triangle points to, the Hamiltonian of the
64 < system is the same, hence there are infinite possibilities for the
49 < packing of the spins.}
50 < \label{fig:frustration}
60 > \includegraphics[width=3in]{./figures/inFrustration.pdf}
61 > \caption{Frustration on triangular lattice, the spins and dipoles are
62 > represented by arrows. The multiple local minima of energy states
63 > induce the frustration for spins and dipoles picking the directions.}
64 > \label{Infig:frustration}
65   \end{figure}
66 < Figure~\ref{fig:frustration} shows an illustration of the frustration
67 < on a triangular lattice. The direction of the spin on top of the
68 < triangle has no effects on the Hamiltonian of the system, therefore
69 < infinite possibilities for the packing of spins induce the frustration
70 < of the lattice.
66 > The spins in figure~\ref{Infig:frustration} shows an illustration of
67 > the frustration for $J < 0$ on a triangular lattice. There are
68 > multiple local minima energy states which are independent of the
69 > direction of the spin on top of the triangle, therefore infinite
70 > possibilities for the packing of spins induce what is known as
71 > ``complete regular frustration'' which leads to disordered low
72 > temperature phases. The similarity goes to the dipoles on a hexagonal
73 > lattice, which are shown by the dipoles in
74 > figure~\ref{Infig:frustration}. In this circumstance, the dipoles want
75 > to be aligned, however, due to the long wave fluctuation, at low
76 > temperature, the aligned state becomes unstable, vortex is formed and
77 > results in multiple local minima of energy states. The dipole on the
78 > center of the hexagonal lattice is frustrated.
79  
80   The lack of translational degree of freedom in lattice models prevents
81 < their utilization on investigating the emergence of the surface
82 < buckling which is the imposition of the ripple formation. In this
83 < dissertation, a modified lattice model is introduced to this specific
62 < situation in Ch.~\ref{chap:mc}.
81 > their utilization in models for surface buckling which would
82 > correspond to ripple formation. In chapter~\ref{chap:mc}, a modified
83 > lattice model is introduced to tackle this specific situation.
84  
85   \section{Overview of Classical Statistical Mechanics\label{In:sec:SM}}
86   Statistical mechanics provides a way to calculate the macroscopic
# Line 95 | Line 116 | normalize $\rho$ to unity,
116   normalize $\rho$ to unity,
117   \begin{equation}
118   1 = \int d \vec q~^N \int d \vec p~^N \rho,
119 < \label{normalized}
119 > \label{Ineq:normalized}
120   \end{equation}
121   then the value of $\rho$ gives the probability of finding the system
122   in a unit volume in the phase space.
# Line 104 | Line 125 | space at any instant can be written as:
125   time. The number of representive points at a given volume in the phase
126   space at any instant can be written as:
127   \begin{equation}
128 < \label{eq:deltaN}
128 > \label{Ineq:deltaN}
129   \delta N = \rho~\delta q_1 \delta q_2 \ldots \delta q_N \delta p_1 \delta p_2 \ldots \delta p_N.
130   \end{equation}
131   To calculate the change in the number of representive points in this
# Line 112 | Line 133 | representive points entering the volume at $q_1$ per u
133   of representive points in $q_1$ axis. The rate of the number of the
134   representive points entering the volume at $q_1$ per unit time is:
135   \begin{equation}
136 < \label{eq:deltaNatq1}
136 > \label{Ineq:deltaNatq1}
137   \rho~\dot q_1 \delta q_2 \ldots \delta q_N \delta p_1 \delta p_2 \ldots \delta p_N,
138   \end{equation}
139   and the rate of the number of representive points leaving the volume
140   at another position $q_1 + \delta q_1$ is:
141   \begin{equation}
142 < \label{eq:deltaNatq1plusdeltaq1}
142 > \label{Ineq:deltaNatq1plusdeltaq1}
143   \left( \rho + \frac{\partial \rho}{\partial q_1} \delta q_1 \right)\left(\dot q_1 +
144   \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.
145   \end{equation}
146   Here the higher order differentials are neglected. So the change of
147   the number of the representive points is the difference of
148 < eq.~\ref{eq:deltaNatq1} and eq.~\ref{eq:deltaNatq1plusdeltaq1}, which
148 > eq.~\ref{Ineq:deltaNatq1} and eq.~\ref{Ineq:deltaNatq1plusdeltaq1}, which
149   gives us:
150   \begin{equation}
151 < \label{eq:deltaNatq1axis}
151 > \label{Ineq:deltaNatq1axis}
152   -\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,
153   \end{equation}
154   where, higher order differetials are neglected. If we sum over all the
155   axes in the phase space, we can get the change of the number of
156   representive points in a given volume with time:
157   \begin{equation}
158 < \label{eq:deltaNatGivenVolume}
158 > \label{Ineq:deltaNatGivenVolume}
159   \frac{d(\delta N)}{dt} = -\sum_{i=1}^N \left[\rho \left(\frac{\partial
160   {\dot q_i}}{\partial q_i} + \frac{\partial
161   {\dot p_i}}{\partial p_i}\right) + \left( \frac{\partial {\rho}}{\partial
# Line 144 | Line 165 | From Hamilton's equation of motion,
165   \begin{equation}
166   \frac{\partial {\dot q_i}}{\partial q_i} = - \frac{\partial
167   {\dot p_i}}{\partial p_i},
168 < \label{eq:canonicalFormOfEquationOfMotion}
168 > \label{Ineq:canonicalFormOfEquationOfMotion}
169   \end{equation}
170   this cancels out the first term on the right side of
171 < eq.~\ref{eq:deltaNatGivenVolume}. If both sides of
172 < eq.~\ref{eq:deltaNatGivenVolume} are divided by $\delta q_1 \delta q_2
171 > eq.~\ref{Ineq:deltaNatGivenVolume}. If both sides of
172 > eq.~\ref{Ineq:deltaNatGivenVolume} are divided by $\delta q_1 \delta q_2
173   \ldots \delta q_N \delta p_1 \delta p_2 \ldots \delta p_N$, then we
174   can derive Liouville's theorem:
175   \begin{equation}
176   \left( \frac{\partial \rho}{\partial t} \right)_{q, p} = -\sum_{i} \left(
177   \frac{\partial {\rho}}{\partial
178   q_i} \dot q_i + \frac{\partial {\rho}}{\partial p_i} \dot p_i \right).
179 < \label{eq:simpleFormofLiouville}
179 > \label{Ineq:simpleFormofLiouville}
180   \end{equation}
181   This is the basis of statistical mechanics. If we move the right
182 < side of equation~\ref{eq:simpleFormofLiouville} to the left, we
182 > side of equation~\ref{Ineq:simpleFormofLiouville} to the left, we
183   will obtain
184   \begin{equation}
185   \left( \frac{\partial \rho}{\partial t} \right)_{q, p} + \sum_{i} \left(
186   \frac{\partial {\rho}}{\partial
187   q_i} \dot q_i + \frac{\partial {\rho}}{\partial p_i} \dot p_i \right)
188   = 0.
189 < \label{eq:anotherFormofLiouville}
189 > \label{Ineq:anotherFormofLiouville}
190   \end{equation}
191   It is easy to note that the left side of
192 < equation~\ref{eq:anotherFormofLiouville} is the total derivative of
192 > equation~\ref{Ineq:anotherFormofLiouville} is the total derivative of
193   $\rho$ with respect of $t$, which means
194   \begin{equation}
195   \frac{d \rho}{dt} = 0,
196 < \label{eq:conservationofRho}
196 > \label{Ineq:conservationofRho}
197   \end{equation}
198   and the rate of density change is zero in the neighborhood of any
199   selected moving representive points in the phase space.
# Line 184 | Line 205 | phase space is zero,
205   phase space is zero,
206   \begin{equation}
207   \left( \frac{\partial \rho}{\partial t} \right)_{q, p} = 0.
208 < \label{eq:statEquilibrium}
208 > \label{Ineq:statEquilibrium}
209   \end{equation}
210   We may conclude the ensemble is in {\it statistical equilibrium}. An
211   ensemble in statistical equilibrium often means the system is also in
# Line 195 | Line 216 | q_i} \dot q_i + \frac{\partial {\rho}}{\partial p_i} \
216   \frac{\partial {\rho}}{\partial
217   q_i} \dot q_i + \frac{\partial {\rho}}{\partial p_i} \dot p_i \right)
218   = 0.
219 < \label{constantofMotion}
219 > \label{Ineq:constantofMotion}
220   \end{equation}
221   If $\rho$ is a function only of some constant of the motion, $\rho$ is
222   independent of time. For a conservative system, the energy of the
# Line 204 | Line 225 | phase space,
225   phase space,
226   \begin{equation}
227   \rho = \mathrm{const.}
228 < \label{eq:uniformEnsemble}
228 > \label{Ineq:uniformEnsemble}
229   \end{equation}
230   the ensemble is called {\it uniform ensemble}.  Another useful
231   ensemble is called {\it microcanonical ensemble}, for which:
232   \begin{equation}
233   \rho = \delta \left( H(q^N, p^N) - E \right) \frac{1}{\Sigma (N, V, E)}
234 < \label{eq:microcanonicalEnsemble}
234 > \label{Ineq:microcanonicalEnsemble}
235   \end{equation}
236   where $\Sigma(N, V, E)$ is a normalization constant parameterized by
237   $N$, the total number of particles, $V$, the total physical volume and
# Line 220 | Line 241 | S = - k_B \int d \vec q~^N \int d \vec p~^N \rho \ln [
241   Hamiltonian of the system. The Gibbs entropy is defined as
242   \begin{equation}
243   S = - k_B \int d \vec q~^N \int d \vec p~^N \rho \ln [C^N \rho],
244 < \label{eq:gibbsEntropy}
244 > \label{Ineq:gibbsEntropy}
245   \end{equation}
246   where $k_B$ is the Boltzmann constant and $C^N$ is a number which
247   makes the argument of $\ln$ dimensionless, in this case, it is the
# Line 228 | Line 249 | S = k_B \ln \left(\frac{\Sigma(N, V, E)}{C^N}\right).
249   ensemble is given by
250   \begin{equation}
251   S = k_B \ln \left(\frac{\Sigma(N, V, E)}{C^N}\right).
252 < \label{eq:entropy}
252 > \label{Ineq:entropy}
253   \end{equation}
254   If the density distribution $\rho$ is given by
255   \begin{equation}
256   \rho = \frac{1}{Z_N}e^{-H(q^N, p^N) / k_B T},
257 < \label{eq:canonicalEnsemble}
257 > \label{Ineq:canonicalEnsemble}
258   \end{equation}
259   the ensemble is known as the {\it canonical ensemble}. Here,
260   \begin{equation}
261   Z_N = \int d \vec q~^N \int_\Gamma d \vec p~^N  e^{-H(q^N, p^N) / k_B T},
262 < \label{eq:partitionFunction}
262 > \label{Ineq:partitionFunction}
263   \end{equation}
264   which is also known as {\it partition function}. $\Gamma$ indicates
265   that the integral is over all the phase space. In the canonical
# Line 250 | Line 271 | the thermodynamics maximizes the entropy $S$,
271   \begin{array}{ccc}
272   \delta S = 0 & \mathrm{and} & \delta^2 S < 0.
273   \end{array}
274 < \label{eq:maximumPrinciple}
274 > \label{Ineq:maximumPrinciple}
275   \end{equation}
276 < From Eq.~\ref{eq:maximumPrinciple} and two constrains of the canonical
276 > From Eq.~\ref{Ineq:maximumPrinciple} and two constrains of the canonical
277   ensemble, {\it i.e.}, total probability and average energy conserved,
278   the partition function is calculated as
279   \begin{equation}
280   Z_N = e^{-A/k_B T},
281 < \label{eq:partitionFunctionWithFreeEnergy}
281 > \label{Ineq:partitionFunctionWithFreeEnergy}
282   \end{equation}
283   where $A$ is the Helmholtz free energy. The significance of
284 < Eq.~\ref{eq:entropy} and~\ref{eq:partitionFunctionWithFreeEnergy} is
284 > Eq.~\ref{Ineq:entropy} and~\ref{Ineq:partitionFunctionWithFreeEnergy} is
285   that they serve as a connection between macroscopic properties of the
286   system and the distribution of the microscopic states.
287  
# Line 276 | Line 297 | can be calculated based on the definition shown by
297   of any quantity ($F(q^N, p^N$)) which depends on the coordinates
298   ($q^N$) and the momenta ($p^N$) for all the systems in the ensemble
299   can be calculated based on the definition shown by
300 < Eq.~\ref{eq:statAverage1}
300 > Eq.~\ref{Ineq:statAverage1}
301   \begin{equation}
302   \langle F(q^N, p^N, t) \rangle = \frac{\int d \vec q~^N \int d \vec p~^N
303   F(q^N, p^N, t) \rho}{\int d \vec q~^N \int d \vec p~^N \rho}.
304 < \label{eq:statAverage1}
304 > \label{Ineq:statAverage1}
305   \end{equation}
306   Since the density distribution $\rho$ is normalized to unity, the mean
307   value of $F(q^N, p^N)$ is simplified to
308   \begin{equation}
309   \langle F(q^N, p^N, t) \rangle = \int d \vec q~^N \int d \vec p~^N F(q^N,
310   p^N, t) \rho,
311 < \label{eq:statAverage2}
311 > \label{Ineq:statAverage2}
312   \end{equation}
313   called {\it ensemble average}. However, the quantity is often averaged
314   for a finite time in real experiments,
315   \begin{equation}
316   \langle F(q^N, p^N) \rangle_t = \lim_{T \rightarrow \infty}
317   \frac{1}{T} \int_{t_0}^{t_0+T} F(q^N, p^N, t) dt.
318 < \label{eq:timeAverage1}
318 > \label{Ineq:timeAverage1}
319   \end{equation}
320   Usually this time average is independent of $t_0$ in statistical
321 < mechanics, so Eq.~\ref{eq:timeAverage1} becomes
321 > mechanics, so Eq.~\ref{Ineq:timeAverage1} becomes
322   \begin{equation}
323   \langle F(q^N, p^N) \rangle_t = \lim_{T \rightarrow \infty}
324   \frac{1}{T} \int_{0}^{T} F(q^N, p^N, t) dt
325 < \label{eq:timeAverage2}
325 > \label{Ineq:timeAverage2}
326   \end{equation}
327   for an infinite time interval.
328  
# Line 311 | Line 332 | phase space. Mathematically, this leads to
332   phase space. Mathematically, this leads to
333   \begin{equation}
334   \langle F(q^N, p^N, t) \rangle = \langle F(q^N, p^N) \rangle_t.
335 < \label{eq:ergodicity}
335 > \label{Ineq:ergodicity}
336   \end{equation}
337 < Eq.~\ref{eq:ergodicity} validates the Monte Carlo method which we will
337 > Eq.~\ref{Ineq:ergodicity} validates the Monte Carlo method which we will
338   discuss in section~\ref{In:ssec:mc}. An ensemble average of a quantity
339   can be related to the time average measured in the experiments.
340  
# Line 332 | Line 353 | C(t) = \langle A(0)A(\tau) \rangle = \lim_{T \rightarr
353   \begin{equation}
354   C(t) = \langle A(0)A(\tau) \rangle = \lim_{T \rightarrow \infty}
355   \frac{1}{T} \int_{0}^{T} dt A(t) A(t + \tau).
356 < \label{eq:autocorrelationFunction}
356 > \label{Ineq:autocorrelationFunction}
357   \end{equation}
358 < Eq.~\ref{eq:autocorrelationFunction} is the correlation function of a
358 > Eq.~\ref{Ineq:autocorrelationFunction} is the correlation function of a
359   single variable, called {\it autocorrelation function}. The defination
360   of the correlation function for two different variables is similar to
361   that of autocorrelation function, which is
362   \begin{equation}
363   C(t) = \langle A(0)B(\tau) \rangle = \lim_{T \rightarrow \infty}
364   \frac{1}{T} \int_{0}^{T} dt A(t) B(t + \tau),
365 < \label{eq:crosscorrelationFunction}
365 > \label{Ineq:crosscorrelationFunction}
366   \end{equation}
367   and called {\it cross correlation function}.
368  
369 < In section~\ref{In:ssec:average} we know from Eq.~\ref{eq:ergodicity}
369 > In section~\ref{In:ssec:average} we know from Eq.~\ref{Ineq:ergodicity}
370   the relationship between time average and ensemble average. We can put
371   the correlation function in a classical mechanics form,
372   \begin{equation}
373   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)
374 < \label{eq:autocorrelationFunctionCM}
374 > \label{Ineq:autocorrelationFunctionCM}
375   \end{equation}
376   and
377   \begin{equation}
378   C(t) = \langle A(0)B(\tau) \rangle = \int d \vec q~^N \int d \vec p~^N A(t) B(t + \tau)
379   \rho(q, p)
380 < \label{eq:crosscorrelationFunctionCM}
380 > \label{Ineq:crosscorrelationFunctionCM}
381   \end{equation}
382   as autocorrelation function and cross correlation function
383   respectively. $\rho(q, p)$ is the density distribution at equillibrium
# Line 364 | Line 385 | C(t) \sim e^{-t / \tau_r},
385   single exponential
386   \begin{equation}
387   C(t) \sim e^{-t / \tau_r},
388 < \label{eq:relaxation}
388 > \label{Ineq:relaxation}
389   \end{equation}
390   where $\tau_r$ is known as relaxation time which discribes the rate of
391   the decay.
# Line 390 | Line 411 | temperature are constants. The average energy is given
411   applied to the canonical ensemble, a Boltzmann-weighted ensemble, in
412   which the $N$, the total number of particles, $V$, total volume, $T$,
413   temperature are constants. The average energy is given by substituding
414 < Eq.~\ref{eq:canonicalEnsemble} into Eq.~\ref{eq:statAverage2},
414 > Eq.~\ref{Ineq:canonicalEnsemble} into Eq.~\ref{Ineq:statAverage2},
415   \begin{equation}
416   \langle E \rangle = \frac{1}{Z_N} \int d \vec q~^N \int d \vec p~^N E e^{-H(q^N, p^N) / k_B T}.
417 < \label{eq:energyofCanonicalEnsemble}
417 > \label{Ineq:energyofCanonicalEnsemble}
418   \end{equation}
419   So are the other properties of the system. The Hamiltonian is the
420   summation of Kinetic energy $K(p^N)$ as a function of momenta and
421   Potential energy $U(q^N)$ as a function of positions,
422   \begin{equation}
423   H(q^N, p^N) = K(p^N) + U(q^N).
424 < \label{eq:hamiltonian}
424 > \label{Ineq:hamiltonian}
425   \end{equation}
426   If the property $A$ is only a function of position ($ A = A(q^N)$),
427   the mean value of $A$ is reduced to
428   \begin{equation}
429   \langle A \rangle = \frac{\int d \vec q~^N \int d \vec p~^N A e^{-U(q^N) / k_B T}}{\int d \vec q~^N \int d \vec p~^N e^{-U(q^N) / k_B T}},
430 < \label{eq:configurationIntegral}
430 > \label{Ineq:configurationIntegral}
431   \end{equation}
432   The kinetic energy $K(p^N)$ is factored out in
433 < Eq.~\ref{eq:configurationIntegral}. $\langle A
433 > Eq.~\ref{Ineq:configurationIntegral}. $\langle A
434   \rangle$ is a configuration integral now, and the
435 < Eq.~\ref{eq:configurationIntegral} is equivalent to
435 > Eq.~\ref{Ineq:configurationIntegral} is equivalent to
436   \begin{equation}
437   \langle A \rangle = \int d \vec q~^N A \rho(q^N).
438 < \label{eq:configurationAve}
438 > \label{Ineq:configurationAve}
439   \end{equation}
440  
441   In a Monte Carlo simulation of canonical ensemble, the probability of
# Line 422 | Line 443 | probability with time is given by
443   probability with time is given by
444   \begin{equation}
445   \frac{d \rho_s}{dt} = \sum_{s'} [ -w_{ss'}\rho_s + w_{s's}\rho_{s'} ],
446 < \label{eq:timeChangeofProb}
446 > \label{Ineq:timeChangeofProb}
447   \end{equation}
448   where $w_{ss'}$ is the tansition probability of going from state $s$
449   to state $s'$. Since $\rho_s$ is independent of time at equilibrium,
450   \begin{equation}
451   \frac{d \rho_{s}^{equilibrium}}{dt} = 0,
452 < \label{eq:equiProb}
452 > \label{Ineq:equiProb}
453   \end{equation}
454   which means $\sum_{s'} [ -w_{ss'}\rho_s + w_{s's}\rho_{s'} ]$ is $0$
455   for all $s'$. So
456   \begin{equation}
457   \frac{\rho_s^{equilibrium}}{\rho_{s'}^{equilibrium}} = \frac{w_{s's}}{w_{ss'}}.
458 < \label{eq:timeChangeofProb}
458 > \label{Ineq:relationshipofRhoandW}
459   \end{equation}
460   If
461   \begin{equation}
462   \frac{w_{s's}}{w_{ss'}} = e^{-(U_s - U_{s'}) / k_B T},
463 < \label{eq:conditionforBoltzmannStatistics}
463 > \label{Ineq:conditionforBoltzmannStatistics}
464   \end{equation}
465   then
466   \begin{equation}
467   \frac{\rho_s^{equilibrium}}{\rho_{s'}^{equilibrium}} = e^{-(U_s - U_{s'}) / k_B T}.
468 < \label{eq:satisfyofBoltzmannStatistics}
468 > \label{Ineq:satisfyofBoltzmannStatistics}
469   \end{equation}
470 < Eq.~\ref{eq:satisfyofBoltzmannStatistics} implies that
470 > Eq.~\ref{Ineq:satisfyofBoltzmannStatistics} implies that
471   $\rho^{equilibrium}$ satisfies Boltzmann statistics. An algorithm,
472   shows how Monte Carlo simulation generates a transition probability
473 < governed by \ref{eq:conditionforBoltzmannStatistics}, is schemed as
473 > governed by \ref{Ineq:conditionforBoltzmannStatistics}, is schemed as
474   \begin{enumerate}
475 < \item\label{itm:oldEnergy} Choose an particle randomly, calculate the energy.
476 < \item\label{itm:newEnergy} Make a random displacement for particle,
475 > \item\label{Initm:oldEnergy} Choose an particle randomly, calculate the energy.
476 > \item\label{Initm:newEnergy} Make a random displacement for particle,
477   calculate the new energy.
478    \begin{itemize}
479 <     \item Keep the new configuration and return to step~\ref{itm:oldEnergy} if energy
479 >     \item Keep the new configuration and return to step~\ref{Initm:oldEnergy} if energy
480   goes down.
481       \item Pick a random number between $[0,1]$ if energy goes up.
482          \begin{itemize}
483             \item Keep the new configuration and return to
484 < step~\ref{itm:oldEnergy} if the random number smaller than
484 > step~\ref{Initm:oldEnergy} if the random number smaller than
485   $e^{-(U_{new} - U_{old})} / k_B T$.
486             \item Keep the old configuration and return to
487 < step~\ref{itm:oldEnergy} if the random number larger than
487 > step~\ref{Initm:oldEnergy} if the random number larger than
488   $e^{-(U_{new} - U_{old})} / k_B T$.
489          \end{itemize}
490    \end{itemize}
491 < \item\label{itm:accumulateAvg} Accumulate the average after it converges.
491 > \item\label{Initm:accumulateAvg} Accumulate the average after it converges.
492   \end{enumerate}
493   It is important to notice that the old configuration has to be sampled
494   again if it is kept.
# Line 489 | Line 510 | defined by
510   defined by
511   \begin{equation}
512   \frac{1}{2} k_B T = \langle \frac{1}{2} m v_\alpha \rangle,
513 < \label{eq:temperature}
513 > \label{Ineq:temperature}
514   \end{equation}
515   here $m$ is the mass of the particle and $v_\alpha$ is the $\alpha$
516   component of the velocity of the particle. The right side of
517 < Eq.~\ref{eq:temperature} is the average kinetic energy of the
517 > Eq.~\ref{Ineq:temperature} is the average kinetic energy of the
518   system. A simple Molecular Dynamics simulation scheme
519   is:~\cite{Frenkel1996}
520   \begin{enumerate}
521 < \item\label{itm:initialize} Assign the initial positions and momenta
521 > \item\label{Initm:initialize} Assign the initial positions and momenta
522   for the particles in the system.
523 < \item\label{itm:calcForce} Calculate the forces.
524 < \item\label{itm:equationofMotion} Integrate the equation of motion.
523 > \item\label{Initm:calcForce} Calculate the forces.
524 > \item\label{Initm:equationofMotion} Integrate the equation of motion.
525    \begin{itemize}
526 <     \item Return to step~\ref{itm:calcForce} if the equillibrium is
526 >     \item Return to step~\ref{Initm:calcForce} if the equillibrium is
527   not achieved.
528 <     \item Go to step~\ref{itm:calcAvg} if the equillibrium is
528 >     \item Go to step~\ref{Initm:calcAvg} if the equillibrium is
529   achieved.
530    \end{itemize}
531 < \item\label{itm:calcAvg} Compute the quantities we are interested in.
531 > \item\label{Initm:calcAvg} Compute the quantities we are interested in.
532   \end{enumerate}
533   The initial positions of the particles are chosen as that there is no
534   overlap for the particles. The initial velocities at first are
535   distributed randomly to the particles, and then shifted to make the
536   momentum of the system $0$, at last scaled to the desired temperature
537 < of the simulation according Eq.~\ref{eq:temperature}.
537 > of the simulation according Eq.~\ref{Ineq:temperature}.
538  
539 < The core of Molecular Dynamics simulations is step~\ref{itm:calcForce}
540 < and~\ref{itm:equationofMotion}. The calculation of the forces are
539 > The core of Molecular Dynamics simulations is step~\ref{Initm:calcForce}
540 > and~\ref{Initm:equationofMotion}. The calculation of the forces are
541   often involved numerous effort, this is the most time consuming step
542   in the Molecular Dynamics scheme. The evaluation of the forces is
543   followed by
544   \begin{equation}
545   f(q) = - \frac{\partial U(q)}{\partial q},
546 < \label{eq:force}
546 > \label{Ineq:force}
547   \end{equation}
548   $U(q)$ is the potential of the system. Once the forces computed, are
549   the positions and velocities updated by integrating Newton's equation
550   of motion,
551   \begin{equation}
552   f(q) = \frac{dp}{dt} = \frac{m dv}{dt}.
553 < \label{eq:newton}
553 > \label{Ineq:newton}
554   \end{equation}
555   Here is an example of integrating algorithms, Verlet algorithm, which
556   is one of the best algorithms to integrate Newton's equation of
# Line 538 | Line 559 | q(t+\Delta t)= q(t) + v(t) \Delta t + \frac{f(t)}{2m}\
559   q(t+\Delta t)= q(t) + v(t) \Delta t + \frac{f(t)}{2m}\Delta t^2 +
560          \frac{\Delta t^3}{3!}\frac{\partial^3 q(t)}{\partial t^3} +
561          \mathcal{O}(\Delta t^4)
562 < \label{eq:verletFuture}
562 > \label{Ineq:verletFuture}
563   \end{equation}
564   for a later time $t+\Delta t$, and
565   \begin{equation}
566   q(t-\Delta t)= q(t) - v(t) \Delta t + \frac{f(t)}{2m}\Delta t^2 -
567          \frac{\Delta t^3}{3!}\frac{\partial^3 q(t)}{\partial t^3} +
568          \mathcal{O}(\Delta t^4) ,
569 < \label{eq:verletPrevious}
569 > \label{Ineq:verletPrevious}
570   \end{equation}
571   for a previous time $t-\Delta t$. The summation of the
572 < Eq.~\ref{eq:verletFuture} and~\ref{eq:verletPrevious} gives
572 > Eq.~\ref{Ineq:verletFuture} and~\ref{Ineq:verletPrevious} gives
573   \begin{equation}
574   q(t+\Delta t)+q(t-\Delta t) =
575          2q(t) + \frac{f(t)}{m}\Delta t^2 + \mathcal{O}(\Delta t^4),
576 < \label{eq:verletSum}
576 > \label{Ineq:verletSum}
577   \end{equation}
578   so, the new position can be expressed as
579   \begin{equation}
580   q(t+\Delta t) \approx
581          2q(t) - q(t-\Delta t) + \frac{f(t)}{m}\Delta t^2.
582 < \label{eq:newPosition}
582 > \label{Ineq:newPosition}
583   \end{equation}
584   The higher order of the $\Delta t$ is omitted.
585  
# Line 586 | Line 607 | H = \frac{p^2}{2m} + U(q) + H_B + \Delta U(q),
607   Hamiltonian of such a system is written as
608   \begin{equation}
609   H = \frac{p^2}{2m} + U(q) + H_B + \Delta U(q),
610 < \label{eq:hamiltonianofCoupling}
610 > \label{Ineq:hamiltonianofCoupling}
611   \end{equation}
612   where $H_B$ is the Hamiltonian of the bath which equals to
613   \begin{equation}
614   H_B = \sum_{\alpha = 1}^{N} \left\{ \frac{p_\alpha^2}{2m_\alpha} +
615   \frac{1}{2} m_\alpha \omega_\alpha^2 q_\alpha^2\right\},
616 < \label{eq:hamiltonianofBath}
616 > \label{Ineq:hamiltonianofBath}
617   \end{equation}
618   $\alpha$ is all the degree of freedoms of the bath, $\omega$ is the
619   bath frequency, and $\Delta U(q)$ is the bilinear coupling given by
620   \begin{equation}
621   \Delta U = -\sum_{\alpha = 1}^{N} g_\alpha q_\alpha q,
622 < \label{eq:systemBathCoupling}
622 > \label{Ineq:systemBathCoupling}
623   \end{equation}
624   where $g$ is the coupling constant. By solving the Hamilton's equation
625   of motion, the {\it Generalized Langevin Equation} for this system is
626   derived to
627   \begin{equation}
628   m \ddot q = -\frac{\partial W(q)}{\partial q} - \int_0^t \xi(t) \dot q(t-t')dt' + R(t),
629 < \label{eq:gle}
629 > \label{Ineq:gle}
630   \end{equation}
631   with mean force,
632   \begin{equation}
633   W(q) = U(q) - \sum_{\alpha = 1}^N \frac{g_\alpha^2}{2m_\alpha
634   \omega_\alpha^2}q^2,
635 < \label{eq:meanForce}
635 > \label{Ineq:meanForce}
636   \end{equation}
637   being only a dependence of coordinates of the solute particles, {\it
638   friction kernel},
639   \begin{equation}
640   \xi(t) = \sum_{\alpha = 1}^N \frac{-g_\alpha}{m_\alpha
641   \omega_\alpha} \cos(\omega_\alpha t),
642 < \label{eq:xiforGLE}
642 > \label{Ineq:xiforGLE}
643   \end{equation}
644   and the random force,
645   \begin{equation}
646   R(t) = \sum_{\alpha = 1}^N \left( g_\alpha q_\alpha(0)-\frac{g_\alpha}{m_\alpha
647   \omega_\alpha^2}q(0)\right) \cos(\omega_\alpha t) + \frac{\dot
648   q_\alpha(0)}{\omega_\alpha} \sin(\omega_\alpha t),
649 < \label{eq:randomForceforGLE}
649 > \label{Ineq:randomForceforGLE}
650   \end{equation}
651   as only a dependence of the initial conditions. The relationship of
652   friction kernel $\xi(t)$ and random force $R(t)$ is given by
653   \begin{equation}
654   \xi(t) = \frac{1}{k_B T} \langle R(t)R(0) \rangle
655 < \label{eq:relationshipofXiandR}
655 > \label{Ineq:relationshipofXiandR}
656   \end{equation}
657   from their definitions. In Langevin limit, the friction is treated
658   static, which means
659   \begin{equation}
660   \xi(t) = 2 \xi_0 \delta(t).
661 < \label{eq:xiofStaticFriction}
661 > \label{Ineq:xiofStaticFriction}
662   \end{equation}
663 < After substitude $\xi(t)$ into Eq.~\ref{eq:gle} with
664 < Eq.~\ref{eq:xiofStaticFriction}, {\it Langevin Equation} is extracted
663 > After substitude $\xi(t)$ into Eq.~\ref{Ineq:gle} with
664 > Eq.~\ref{Ineq:xiofStaticFriction}, {\it Langevin Equation} is extracted
665   to
666   \begin{equation}
667   m \ddot q = -\frac{\partial U(q)}{\partial q} - \xi \dot q(t) + R(t).
668 < \label{eq:langevinEquation}
668 > \label{Ineq:langevinEquation}
669   \end{equation}
670   The applying of Langevin Equation to dynamic simulations is discussed
671   in Ch.~\ref{chap:ld}.

Diff Legend

Removed lines
+ Added lines
< Changed lines
> Changed lines