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 3365 by xsun, Mon Mar 10 21:52:50 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.
55   \begin{figure}
56   \centering
57 < \includegraphics[width=\linewidth]{./figures/frustration.pdf}
58 < \caption{Sketch to illustrate the frustration on triangular
59 < lattice. Spins are represented by arrows, no matter which direction
60 < the spin on the top of triangle points to, the Hamiltonian of the
61 < system is the same, hence there are infinite possibilities for the
49 < packing of the spins.}
50 < \label{fig:frustration}
57 > \includegraphics[width=\linewidth]{./figures/inFrustration.pdf}
58 > \caption{Frustration on a triangular lattice, the spins are
59 > represented by arrows. No matter which direction the spin on the top
60 > of triangle points to, the Hamiltonain of the system goes up.}
61 > \label{Infig:frustration}
62   \end{figure}
63 < Figure~\ref{fig:frustration} shows an illustration of the frustration
64 < on a triangular lattice. The direction of the spin on top of the
65 < triangle has no effects on the Hamiltonian of the system, therefore
66 < infinite possibilities for the packing of spins induce the frustration
67 < of the lattice.
63 > Figure~\ref{Infig:frustration} shows an illustration of the
64 > frustration on a triangular lattice. When $J < 0$, the spins want to
65 > be anti-aligned, The direction of the spin on top of the triangle will
66 > make the energy go up no matter which direction it picks, therefore
67 > infinite possibilities for the packing of spins induce what is known
68 > as ``complete regular frustration'' which leads to disordered low
69 > temperature phases.
70  
71   The lack of translational degree of freedom in lattice models prevents
72 < their utilization on investigating the emergence of the surface
73 < buckling which is the imposition of the ripple formation. In this
74 < dissertation, a modified lattice model is introduced to this specific
62 < situation in Ch.~\ref{chap:mc}.
72 > their utilization in models for surface buckling which would
73 > correspond to ripple formation. In chapter~\ref{chap:mc}, a modified
74 > lattice model is introduced to tackle this specific situation.
75  
76   \section{Overview of Classical Statistical Mechanics\label{In:sec:SM}}
77   Statistical mechanics provides a way to calculate the macroscopic
# Line 95 | Line 107 | normalize $\rho$ to unity,
107   normalize $\rho$ to unity,
108   \begin{equation}
109   1 = \int d \vec q~^N \int d \vec p~^N \rho,
110 < \label{normalized}
110 > \label{Ineq:normalized}
111   \end{equation}
112   then the value of $\rho$ gives the probability of finding the system
113   in a unit volume in the phase space.
# Line 104 | Line 116 | space at any instant can be written as:
116   time. The number of representive points at a given volume in the phase
117   space at any instant can be written as:
118   \begin{equation}
119 < \label{eq:deltaN}
119 > \label{Ineq:deltaN}
120   \delta N = \rho~\delta q_1 \delta q_2 \ldots \delta q_N \delta p_1 \delta p_2 \ldots \delta p_N.
121   \end{equation}
122   To calculate the change in the number of representive points in this
# Line 112 | Line 124 | representive points entering the volume at $q_1$ per u
124   of representive points in $q_1$ axis. The rate of the number of the
125   representive points entering the volume at $q_1$ per unit time is:
126   \begin{equation}
127 < \label{eq:deltaNatq1}
127 > \label{Ineq:deltaNatq1}
128   \rho~\dot q_1 \delta q_2 \ldots \delta q_N \delta p_1 \delta p_2 \ldots \delta p_N,
129   \end{equation}
130   and the rate of the number of representive points leaving the volume
131   at another position $q_1 + \delta q_1$ is:
132   \begin{equation}
133 < \label{eq:deltaNatq1plusdeltaq1}
133 > \label{Ineq:deltaNatq1plusdeltaq1}
134   \left( \rho + \frac{\partial \rho}{\partial q_1} \delta q_1 \right)\left(\dot q_1 +
135   \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.
136   \end{equation}
137   Here the higher order differentials are neglected. So the change of
138   the number of the representive points is the difference of
139 < eq.~\ref{eq:deltaNatq1} and eq.~\ref{eq:deltaNatq1plusdeltaq1}, which
139 > eq.~\ref{Ineq:deltaNatq1} and eq.~\ref{Ineq:deltaNatq1plusdeltaq1}, which
140   gives us:
141   \begin{equation}
142 < \label{eq:deltaNatq1axis}
142 > \label{Ineq:deltaNatq1axis}
143   -\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,
144   \end{equation}
145   where, higher order differetials are neglected. If we sum over all the
146   axes in the phase space, we can get the change of the number of
147   representive points in a given volume with time:
148   \begin{equation}
149 < \label{eq:deltaNatGivenVolume}
149 > \label{Ineq:deltaNatGivenVolume}
150   \frac{d(\delta N)}{dt} = -\sum_{i=1}^N \left[\rho \left(\frac{\partial
151   {\dot q_i}}{\partial q_i} + \frac{\partial
152   {\dot p_i}}{\partial p_i}\right) + \left( \frac{\partial {\rho}}{\partial
# Line 144 | Line 156 | From Hamilton's equation of motion,
156   \begin{equation}
157   \frac{\partial {\dot q_i}}{\partial q_i} = - \frac{\partial
158   {\dot p_i}}{\partial p_i},
159 < \label{eq:canonicalFormOfEquationOfMotion}
159 > \label{Ineq:canonicalFormOfEquationOfMotion}
160   \end{equation}
161   this cancels out the first term on the right side of
162 < eq.~\ref{eq:deltaNatGivenVolume}. If both sides of
163 < eq.~\ref{eq:deltaNatGivenVolume} are divided by $\delta q_1 \delta q_2
162 > eq.~\ref{Ineq:deltaNatGivenVolume}. If both sides of
163 > eq.~\ref{Ineq:deltaNatGivenVolume} are divided by $\delta q_1 \delta q_2
164   \ldots \delta q_N \delta p_1 \delta p_2 \ldots \delta p_N$, then we
165   can derive Liouville's theorem:
166   \begin{equation}
167   \left( \frac{\partial \rho}{\partial t} \right)_{q, p} = -\sum_{i} \left(
168   \frac{\partial {\rho}}{\partial
169   q_i} \dot q_i + \frac{\partial {\rho}}{\partial p_i} \dot p_i \right).
170 < \label{eq:simpleFormofLiouville}
170 > \label{Ineq:simpleFormofLiouville}
171   \end{equation}
172   This is the basis of statistical mechanics. If we move the right
173 < side of equation~\ref{eq:simpleFormofLiouville} to the left, we
173 > side of equation~\ref{Ineq:simpleFormofLiouville} to the left, we
174   will obtain
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   = 0.
180 < \label{eq:anotherFormofLiouville}
180 > \label{Ineq:anotherFormofLiouville}
181   \end{equation}
182   It is easy to note that the left side of
183 < equation~\ref{eq:anotherFormofLiouville} is the total derivative of
183 > equation~\ref{Ineq:anotherFormofLiouville} is the total derivative of
184   $\rho$ with respect of $t$, which means
185   \begin{equation}
186   \frac{d \rho}{dt} = 0,
187 < \label{eq:conservationofRho}
187 > \label{Ineq:conservationofRho}
188   \end{equation}
189   and the rate of density change is zero in the neighborhood of any
190   selected moving representive points in the phase space.
# Line 184 | Line 196 | phase space is zero,
196   phase space is zero,
197   \begin{equation}
198   \left( \frac{\partial \rho}{\partial t} \right)_{q, p} = 0.
199 < \label{eq:statEquilibrium}
199 > \label{Ineq:statEquilibrium}
200   \end{equation}
201   We may conclude the ensemble is in {\it statistical equilibrium}. An
202   ensemble in statistical equilibrium often means the system is also in
# Line 195 | Line 207 | q_i} \dot q_i + \frac{\partial {\rho}}{\partial p_i} \
207   \frac{\partial {\rho}}{\partial
208   q_i} \dot q_i + \frac{\partial {\rho}}{\partial p_i} \dot p_i \right)
209   = 0.
210 < \label{constantofMotion}
210 > \label{Ineq:constantofMotion}
211   \end{equation}
212   If $\rho$ is a function only of some constant of the motion, $\rho$ is
213   independent of time. For a conservative system, the energy of the
# Line 204 | Line 216 | phase space,
216   phase space,
217   \begin{equation}
218   \rho = \mathrm{const.}
219 < \label{eq:uniformEnsemble}
219 > \label{Ineq:uniformEnsemble}
220   \end{equation}
221   the ensemble is called {\it uniform ensemble}.  Another useful
222   ensemble is called {\it microcanonical ensemble}, for which:
223   \begin{equation}
224   \rho = \delta \left( H(q^N, p^N) - E \right) \frac{1}{\Sigma (N, V, E)}
225 < \label{eq:microcanonicalEnsemble}
225 > \label{Ineq:microcanonicalEnsemble}
226   \end{equation}
227   where $\Sigma(N, V, E)$ is a normalization constant parameterized by
228   $N$, the total number of particles, $V$, the total physical volume and
# Line 220 | Line 232 | S = - k_B \int d \vec q~^N \int d \vec p~^N \rho \ln [
232   Hamiltonian of the system. The Gibbs entropy is defined as
233   \begin{equation}
234   S = - k_B \int d \vec q~^N \int d \vec p~^N \rho \ln [C^N \rho],
235 < \label{eq:gibbsEntropy}
235 > \label{Ineq:gibbsEntropy}
236   \end{equation}
237   where $k_B$ is the Boltzmann constant and $C^N$ is a number which
238   makes the argument of $\ln$ dimensionless, in this case, it is the
# Line 228 | Line 240 | S = k_B \ln \left(\frac{\Sigma(N, V, E)}{C^N}\right).
240   ensemble is given by
241   \begin{equation}
242   S = k_B \ln \left(\frac{\Sigma(N, V, E)}{C^N}\right).
243 < \label{eq:entropy}
243 > \label{Ineq:entropy}
244   \end{equation}
245   If the density distribution $\rho$ is given by
246   \begin{equation}
247   \rho = \frac{1}{Z_N}e^{-H(q^N, p^N) / k_B T},
248 < \label{eq:canonicalEnsemble}
248 > \label{Ineq:canonicalEnsemble}
249   \end{equation}
250   the ensemble is known as the {\it canonical ensemble}. Here,
251   \begin{equation}
252   Z_N = \int d \vec q~^N \int_\Gamma d \vec p~^N  e^{-H(q^N, p^N) / k_B T},
253 < \label{eq:partitionFunction}
253 > \label{Ineq:partitionFunction}
254   \end{equation}
255   which is also known as {\it partition function}. $\Gamma$ indicates
256   that the integral is over all the phase space. In the canonical
# Line 250 | Line 262 | the thermodynamics maximizes the entropy $S$,
262   \begin{array}{ccc}
263   \delta S = 0 & \mathrm{and} & \delta^2 S < 0.
264   \end{array}
265 < \label{eq:maximumPrinciple}
265 > \label{Ineq:maximumPrinciple}
266   \end{equation}
267 < From Eq.~\ref{eq:maximumPrinciple} and two constrains of the canonical
267 > From Eq.~\ref{Ineq:maximumPrinciple} and two constrains of the canonical
268   ensemble, {\it i.e.}, total probability and average energy conserved,
269   the partition function is calculated as
270   \begin{equation}
271   Z_N = e^{-A/k_B T},
272 < \label{eq:partitionFunctionWithFreeEnergy}
272 > \label{Ineq:partitionFunctionWithFreeEnergy}
273   \end{equation}
274   where $A$ is the Helmholtz free energy. The significance of
275 < Eq.~\ref{eq:entropy} and~\ref{eq:partitionFunctionWithFreeEnergy} is
275 > Eq.~\ref{Ineq:entropy} and~\ref{Ineq:partitionFunctionWithFreeEnergy} is
276   that they serve as a connection between macroscopic properties of the
277   system and the distribution of the microscopic states.
278  
# Line 276 | Line 288 | can be calculated based on the definition shown by
288   of any quantity ($F(q^N, p^N$)) which depends on the coordinates
289   ($q^N$) and the momenta ($p^N$) for all the systems in the ensemble
290   can be calculated based on the definition shown by
291 < Eq.~\ref{eq:statAverage1}
291 > Eq.~\ref{Ineq:statAverage1}
292   \begin{equation}
293   \langle F(q^N, p^N, t) \rangle = \frac{\int d \vec q~^N \int d \vec p~^N
294   F(q^N, p^N, t) \rho}{\int d \vec q~^N \int d \vec p~^N \rho}.
295 < \label{eq:statAverage1}
295 > \label{Ineq:statAverage1}
296   \end{equation}
297   Since the density distribution $\rho$ is normalized to unity, the mean
298   value of $F(q^N, p^N)$ is simplified to
299   \begin{equation}
300   \langle F(q^N, p^N, t) \rangle = \int d \vec q~^N \int d \vec p~^N F(q^N,
301   p^N, t) \rho,
302 < \label{eq:statAverage2}
302 > \label{Ineq:statAverage2}
303   \end{equation}
304   called {\it ensemble average}. However, the quantity is often averaged
305   for a finite time in real experiments,
306   \begin{equation}
307   \langle F(q^N, p^N) \rangle_t = \lim_{T \rightarrow \infty}
308   \frac{1}{T} \int_{t_0}^{t_0+T} F(q^N, p^N, t) dt.
309 < \label{eq:timeAverage1}
309 > \label{Ineq:timeAverage1}
310   \end{equation}
311   Usually this time average is independent of $t_0$ in statistical
312 < mechanics, so Eq.~\ref{eq:timeAverage1} becomes
312 > mechanics, so Eq.~\ref{Ineq:timeAverage1} becomes
313   \begin{equation}
314   \langle F(q^N, p^N) \rangle_t = \lim_{T \rightarrow \infty}
315   \frac{1}{T} \int_{0}^{T} F(q^N, p^N, t) dt
316 < \label{eq:timeAverage2}
316 > \label{Ineq:timeAverage2}
317   \end{equation}
318   for an infinite time interval.
319  
# Line 311 | Line 323 | phase space. Mathematically, this leads to
323   phase space. Mathematically, this leads to
324   \begin{equation}
325   \langle F(q^N, p^N, t) \rangle = \langle F(q^N, p^N) \rangle_t.
326 < \label{eq:ergodicity}
326 > \label{Ineq:ergodicity}
327   \end{equation}
328 < Eq.~\ref{eq:ergodicity} validates the Monte Carlo method which we will
328 > Eq.~\ref{Ineq:ergodicity} validates the Monte Carlo method which we will
329   discuss in section~\ref{In:ssec:mc}. An ensemble average of a quantity
330   can be related to the time average measured in the experiments.
331  
# Line 332 | Line 344 | C(t) = \langle A(0)A(\tau) \rangle = \lim_{T \rightarr
344   \begin{equation}
345   C(t) = \langle A(0)A(\tau) \rangle = \lim_{T \rightarrow \infty}
346   \frac{1}{T} \int_{0}^{T} dt A(t) A(t + \tau).
347 < \label{eq:autocorrelationFunction}
347 > \label{Ineq:autocorrelationFunction}
348   \end{equation}
349 < Eq.~\ref{eq:autocorrelationFunction} is the correlation function of a
349 > Eq.~\ref{Ineq:autocorrelationFunction} is the correlation function of a
350   single variable, called {\it autocorrelation function}. The defination
351   of the correlation function for two different variables is similar to
352   that of autocorrelation function, which is
353   \begin{equation}
354   C(t) = \langle A(0)B(\tau) \rangle = \lim_{T \rightarrow \infty}
355   \frac{1}{T} \int_{0}^{T} dt A(t) B(t + \tau),
356 < \label{eq:crosscorrelationFunction}
356 > \label{Ineq:crosscorrelationFunction}
357   \end{equation}
358   and called {\it cross correlation function}.
359  
360 < In section~\ref{In:ssec:average} we know from Eq.~\ref{eq:ergodicity}
360 > In section~\ref{In:ssec:average} we know from Eq.~\ref{Ineq:ergodicity}
361   the relationship between time average and ensemble average. We can put
362   the correlation function in a classical mechanics form,
363   \begin{equation}
364   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)
365 < \label{eq:autocorrelationFunctionCM}
365 > \label{Ineq:autocorrelationFunctionCM}
366   \end{equation}
367   and
368   \begin{equation}
369   C(t) = \langle A(0)B(\tau) \rangle = \int d \vec q~^N \int d \vec p~^N A(t) B(t + \tau)
370   \rho(q, p)
371 < \label{eq:crosscorrelationFunctionCM}
371 > \label{Ineq:crosscorrelationFunctionCM}
372   \end{equation}
373   as autocorrelation function and cross correlation function
374   respectively. $\rho(q, p)$ is the density distribution at equillibrium
# Line 364 | Line 376 | C(t) \sim e^{-t / \tau_r},
376   single exponential
377   \begin{equation}
378   C(t) \sim e^{-t / \tau_r},
379 < \label{eq:relaxation}
379 > \label{Ineq:relaxation}
380   \end{equation}
381   where $\tau_r$ is known as relaxation time which discribes the rate of
382   the decay.
# Line 390 | Line 402 | temperature are constants. The average energy is given
402   applied to the canonical ensemble, a Boltzmann-weighted ensemble, in
403   which the $N$, the total number of particles, $V$, total volume, $T$,
404   temperature are constants. The average energy is given by substituding
405 < Eq.~\ref{eq:canonicalEnsemble} into Eq.~\ref{eq:statAverage2},
405 > Eq.~\ref{Ineq:canonicalEnsemble} into Eq.~\ref{Ineq:statAverage2},
406   \begin{equation}
407   \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}.
408 < \label{eq:energyofCanonicalEnsemble}
408 > \label{Ineq:energyofCanonicalEnsemble}
409   \end{equation}
410   So are the other properties of the system. The Hamiltonian is the
411   summation of Kinetic energy $K(p^N)$ as a function of momenta and
412   Potential energy $U(q^N)$ as a function of positions,
413   \begin{equation}
414   H(q^N, p^N) = K(p^N) + U(q^N).
415 < \label{eq:hamiltonian}
415 > \label{Ineq:hamiltonian}
416   \end{equation}
417   If the property $A$ is only a function of position ($ A = A(q^N)$),
418   the mean value of $A$ is reduced to
419   \begin{equation}
420   \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}},
421 < \label{eq:configurationIntegral}
421 > \label{Ineq:configurationIntegral}
422   \end{equation}
423   The kinetic energy $K(p^N)$ is factored out in
424 < Eq.~\ref{eq:configurationIntegral}. $\langle A
424 > Eq.~\ref{Ineq:configurationIntegral}. $\langle A
425   \rangle$ is a configuration integral now, and the
426 < Eq.~\ref{eq:configurationIntegral} is equivalent to
426 > Eq.~\ref{Ineq:configurationIntegral} is equivalent to
427   \begin{equation}
428   \langle A \rangle = \int d \vec q~^N A \rho(q^N).
429 < \label{eq:configurationAve}
429 > \label{Ineq:configurationAve}
430   \end{equation}
431  
432   In a Monte Carlo simulation of canonical ensemble, the probability of
# Line 422 | Line 434 | probability with time is given by
434   probability with time is given by
435   \begin{equation}
436   \frac{d \rho_s}{dt} = \sum_{s'} [ -w_{ss'}\rho_s + w_{s's}\rho_{s'} ],
437 < \label{eq:timeChangeofProb}
437 > \label{Ineq:timeChangeofProb}
438   \end{equation}
439   where $w_{ss'}$ is the tansition probability of going from state $s$
440   to state $s'$. Since $\rho_s$ is independent of time at equilibrium,
441   \begin{equation}
442   \frac{d \rho_{s}^{equilibrium}}{dt} = 0,
443 < \label{eq:equiProb}
443 > \label{Ineq:equiProb}
444   \end{equation}
445   which means $\sum_{s'} [ -w_{ss'}\rho_s + w_{s's}\rho_{s'} ]$ is $0$
446   for all $s'$. So
447   \begin{equation}
448   \frac{\rho_s^{equilibrium}}{\rho_{s'}^{equilibrium}} = \frac{w_{s's}}{w_{ss'}}.
449 < \label{eq:timeChangeofProb}
449 > \label{Ineq:relationshipofRhoandW}
450   \end{equation}
451   If
452   \begin{equation}
453   \frac{w_{s's}}{w_{ss'}} = e^{-(U_s - U_{s'}) / k_B T},
454 < \label{eq:conditionforBoltzmannStatistics}
454 > \label{Ineq:conditionforBoltzmannStatistics}
455   \end{equation}
456   then
457   \begin{equation}
458   \frac{\rho_s^{equilibrium}}{\rho_{s'}^{equilibrium}} = e^{-(U_s - U_{s'}) / k_B T}.
459 < \label{eq:satisfyofBoltzmannStatistics}
459 > \label{Ineq:satisfyofBoltzmannStatistics}
460   \end{equation}
461 < Eq.~\ref{eq:satisfyofBoltzmannStatistics} implies that
461 > Eq.~\ref{Ineq:satisfyofBoltzmannStatistics} implies that
462   $\rho^{equilibrium}$ satisfies Boltzmann statistics. An algorithm,
463   shows how Monte Carlo simulation generates a transition probability
464 < governed by \ref{eq:conditionforBoltzmannStatistics}, is schemed as
464 > governed by \ref{Ineq:conditionforBoltzmannStatistics}, is schemed as
465   \begin{enumerate}
466 < \item\label{itm:oldEnergy} Choose an particle randomly, calculate the energy.
467 < \item\label{itm:newEnergy} Make a random displacement for particle,
466 > \item\label{Initm:oldEnergy} Choose an particle randomly, calculate the energy.
467 > \item\label{Initm:newEnergy} Make a random displacement for particle,
468   calculate the new energy.
469    \begin{itemize}
470 <     \item Keep the new configuration and return to step~\ref{itm:oldEnergy} if energy
470 >     \item Keep the new configuration and return to step~\ref{Initm:oldEnergy} if energy
471   goes down.
472       \item Pick a random number between $[0,1]$ if energy goes up.
473          \begin{itemize}
474             \item Keep the new configuration and return to
475 < step~\ref{itm:oldEnergy} if the random number smaller than
475 > step~\ref{Initm:oldEnergy} if the random number smaller than
476   $e^{-(U_{new} - U_{old})} / k_B T$.
477             \item Keep the old configuration and return to
478 < step~\ref{itm:oldEnergy} if the random number larger than
478 > step~\ref{Initm:oldEnergy} if the random number larger than
479   $e^{-(U_{new} - U_{old})} / k_B T$.
480          \end{itemize}
481    \end{itemize}
482 < \item\label{itm:accumulateAvg} Accumulate the average after it converges.
482 > \item\label{Initm:accumulateAvg} Accumulate the average after it converges.
483   \end{enumerate}
484   It is important to notice that the old configuration has to be sampled
485   again if it is kept.
# Line 489 | Line 501 | defined by
501   defined by
502   \begin{equation}
503   \frac{1}{2} k_B T = \langle \frac{1}{2} m v_\alpha \rangle,
504 < \label{eq:temperature}
504 > \label{Ineq:temperature}
505   \end{equation}
506   here $m$ is the mass of the particle and $v_\alpha$ is the $\alpha$
507   component of the velocity of the particle. The right side of
508 < Eq.~\ref{eq:temperature} is the average kinetic energy of the
508 > Eq.~\ref{Ineq:temperature} is the average kinetic energy of the
509   system. A simple Molecular Dynamics simulation scheme
510   is:~\cite{Frenkel1996}
511   \begin{enumerate}
512 < \item\label{itm:initialize} Assign the initial positions and momenta
512 > \item\label{Initm:initialize} Assign the initial positions and momenta
513   for the particles in the system.
514 < \item\label{itm:calcForce} Calculate the forces.
515 < \item\label{itm:equationofMotion} Integrate the equation of motion.
514 > \item\label{Initm:calcForce} Calculate the forces.
515 > \item\label{Initm:equationofMotion} Integrate the equation of motion.
516    \begin{itemize}
517 <     \item Return to step~\ref{itm:calcForce} if the equillibrium is
517 >     \item Return to step~\ref{Initm:calcForce} if the equillibrium is
518   not achieved.
519 <     \item Go to step~\ref{itm:calcAvg} if the equillibrium is
519 >     \item Go to step~\ref{Initm:calcAvg} if the equillibrium is
520   achieved.
521    \end{itemize}
522 < \item\label{itm:calcAvg} Compute the quantities we are interested in.
522 > \item\label{Initm:calcAvg} Compute the quantities we are interested in.
523   \end{enumerate}
524   The initial positions of the particles are chosen as that there is no
525   overlap for the particles. The initial velocities at first are
526   distributed randomly to the particles, and then shifted to make the
527   momentum of the system $0$, at last scaled to the desired temperature
528 < of the simulation according Eq.~\ref{eq:temperature}.
528 > of the simulation according Eq.~\ref{Ineq:temperature}.
529  
530 < The core of Molecular Dynamics simulations is step~\ref{itm:calcForce}
531 < and~\ref{itm:equationofMotion}. The calculation of the forces are
530 > The core of Molecular Dynamics simulations is step~\ref{Initm:calcForce}
531 > and~\ref{Initm:equationofMotion}. The calculation of the forces are
532   often involved numerous effort, this is the most time consuming step
533   in the Molecular Dynamics scheme. The evaluation of the forces is
534   followed by
535   \begin{equation}
536   f(q) = - \frac{\partial U(q)}{\partial q},
537 < \label{eq:force}
537 > \label{Ineq:force}
538   \end{equation}
539   $U(q)$ is the potential of the system. Once the forces computed, are
540   the positions and velocities updated by integrating Newton's equation
541   of motion,
542   \begin{equation}
543   f(q) = \frac{dp}{dt} = \frac{m dv}{dt}.
544 < \label{eq:newton}
544 > \label{Ineq:newton}
545   \end{equation}
546   Here is an example of integrating algorithms, Verlet algorithm, which
547   is one of the best algorithms to integrate Newton's equation of
# Line 538 | Line 550 | q(t+\Delta t)= q(t) + v(t) \Delta t + \frac{f(t)}{2m}\
550   q(t+\Delta t)= q(t) + v(t) \Delta t + \frac{f(t)}{2m}\Delta t^2 +
551          \frac{\Delta t^3}{3!}\frac{\partial^3 q(t)}{\partial t^3} +
552          \mathcal{O}(\Delta t^4)
553 < \label{eq:verletFuture}
553 > \label{Ineq:verletFuture}
554   \end{equation}
555   for a later time $t+\Delta t$, and
556   \begin{equation}
557   q(t-\Delta t)= q(t) - v(t) \Delta t + \frac{f(t)}{2m}\Delta t^2 -
558          \frac{\Delta t^3}{3!}\frac{\partial^3 q(t)}{\partial t^3} +
559          \mathcal{O}(\Delta t^4) ,
560 < \label{eq:verletPrevious}
560 > \label{Ineq:verletPrevious}
561   \end{equation}
562   for a previous time $t-\Delta t$. The summation of the
563 < Eq.~\ref{eq:verletFuture} and~\ref{eq:verletPrevious} gives
563 > Eq.~\ref{Ineq:verletFuture} and~\ref{Ineq:verletPrevious} gives
564   \begin{equation}
565   q(t+\Delta t)+q(t-\Delta t) =
566          2q(t) + \frac{f(t)}{m}\Delta t^2 + \mathcal{O}(\Delta t^4),
567 < \label{eq:verletSum}
567 > \label{Ineq:verletSum}
568   \end{equation}
569   so, the new position can be expressed as
570   \begin{equation}
571   q(t+\Delta t) \approx
572          2q(t) - q(t-\Delta t) + \frac{f(t)}{m}\Delta t^2.
573 < \label{eq:newPosition}
573 > \label{Ineq:newPosition}
574   \end{equation}
575   The higher order of the $\Delta t$ is omitted.
576  
# Line 586 | Line 598 | H = \frac{p^2}{2m} + U(q) + H_B + \Delta U(q),
598   Hamiltonian of such a system is written as
599   \begin{equation}
600   H = \frac{p^2}{2m} + U(q) + H_B + \Delta U(q),
601 < \label{eq:hamiltonianofCoupling}
601 > \label{Ineq:hamiltonianofCoupling}
602   \end{equation}
603   where $H_B$ is the Hamiltonian of the bath which equals to
604   \begin{equation}
605   H_B = \sum_{\alpha = 1}^{N} \left\{ \frac{p_\alpha^2}{2m_\alpha} +
606   \frac{1}{2} m_\alpha \omega_\alpha^2 q_\alpha^2\right\},
607 < \label{eq:hamiltonianofBath}
607 > \label{Ineq:hamiltonianofBath}
608   \end{equation}
609   $\alpha$ is all the degree of freedoms of the bath, $\omega$ is the
610   bath frequency, and $\Delta U(q)$ is the bilinear coupling given by
611   \begin{equation}
612   \Delta U = -\sum_{\alpha = 1}^{N} g_\alpha q_\alpha q,
613 < \label{eq:systemBathCoupling}
613 > \label{Ineq:systemBathCoupling}
614   \end{equation}
615   where $g$ is the coupling constant. By solving the Hamilton's equation
616   of motion, the {\it Generalized Langevin Equation} for this system is
617   derived to
618   \begin{equation}
619   m \ddot q = -\frac{\partial W(q)}{\partial q} - \int_0^t \xi(t) \dot q(t-t')dt' + R(t),
620 < \label{eq:gle}
620 > \label{Ineq:gle}
621   \end{equation}
622   with mean force,
623   \begin{equation}
624   W(q) = U(q) - \sum_{\alpha = 1}^N \frac{g_\alpha^2}{2m_\alpha
625   \omega_\alpha^2}q^2,
626 < \label{eq:meanForce}
626 > \label{Ineq:meanForce}
627   \end{equation}
628   being only a dependence of coordinates of the solute particles, {\it
629   friction kernel},
630   \begin{equation}
631   \xi(t) = \sum_{\alpha = 1}^N \frac{-g_\alpha}{m_\alpha
632   \omega_\alpha} \cos(\omega_\alpha t),
633 < \label{eq:xiforGLE}
633 > \label{Ineq:xiforGLE}
634   \end{equation}
635   and the random force,
636   \begin{equation}
637   R(t) = \sum_{\alpha = 1}^N \left( g_\alpha q_\alpha(0)-\frac{g_\alpha}{m_\alpha
638   \omega_\alpha^2}q(0)\right) \cos(\omega_\alpha t) + \frac{\dot
639   q_\alpha(0)}{\omega_\alpha} \sin(\omega_\alpha t),
640 < \label{eq:randomForceforGLE}
640 > \label{Ineq:randomForceforGLE}
641   \end{equation}
642   as only a dependence of the initial conditions. The relationship of
643   friction kernel $\xi(t)$ and random force $R(t)$ is given by
644   \begin{equation}
645   \xi(t) = \frac{1}{k_B T} \langle R(t)R(0) \rangle
646 < \label{eq:relationshipofXiandR}
646 > \label{Ineq:relationshipofXiandR}
647   \end{equation}
648   from their definitions. In Langevin limit, the friction is treated
649   static, which means
650   \begin{equation}
651   \xi(t) = 2 \xi_0 \delta(t).
652 < \label{eq:xiofStaticFriction}
652 > \label{Ineq:xiofStaticFriction}
653   \end{equation}
654 < After substitude $\xi(t)$ into Eq.~\ref{eq:gle} with
655 < Eq.~\ref{eq:xiofStaticFriction}, {\it Langevin Equation} is extracted
654 > After substitude $\xi(t)$ into Eq.~\ref{Ineq:gle} with
655 > Eq.~\ref{Ineq:xiofStaticFriction}, {\it Langevin Equation} is extracted
656   to
657   \begin{equation}
658   m \ddot q = -\frac{\partial U(q)}{\partial q} - \xi \dot q(t) + R(t).
659 < \label{eq:langevinEquation}
659 > \label{Ineq:langevinEquation}
660   \end{equation}
661   The applying of Langevin Equation to dynamic simulations is discussed
662   in Ch.~\ref{chap:ld}.

Diff Legend

Removed lines
+ Added lines
< Changed lines
> Changed lines