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

Comparing trunk/mattDisertation/Introduction.tex (file contents):
Revision 1008 by mmeineke, Tue Feb 3 17:41:56 2004 UTC vs.
Revision 1063 by mmeineke, Mon Feb 23 19:16:22 2004 UTC

# Line 1 | Line 1
1  
2  
3 < \chapter{\label{chapt:intro}Introduction and Theoretical Background}
3 > \chapter{\label{chapt:intro}INTRODUCTION AND THEORETICAL BACKGROUND}
4  
5  
6   The techniques used in the course of this research fall under the two
7   main classes of molecular simulation: Molecular Dynamics and Monte
8 < Carlo. Molecular Dynamic simulations integrate the equations of motion
9 < for a given system of particles, allowing the researcher to gain
10 < insight into the time dependent evolution of a system. Diffusion
8 > Carlo. Molecular Dynamics simulations integrate the equations of
9 > motion for a given system of particles, allowing the researcher to
10 > gain insight into the time dependent evolution of a system. Diffusion
11   phenomena are readily studied with this simulation technique, making
12   Molecular Dynamics the main simulation technique used in this
13   research. Other aspects of the research fall under the Monte Carlo
14   class of simulations. In Monte Carlo, the configuration space
15 < available to the collection of particles is sampled stochastically,
16 < or randomly. Each configuration is chosen with a given probability
17 < based on the Maxwell Boltzmann distribution. These types of simulations
18 < are best used to probe properties of a system that are only dependent
19 < only on the state of the system. Structural information about a system
20 < is most readily obtained through these types of methods.
15 > available to the collection of particles is sampled stochastically, or
16 > randomly. Each configuration is chosen with a given probability based
17 > on the Maxwell Boltzmann distribution. These types of simulations are
18 > best used to probe properties of a system that are dependent only on
19 > the state of the system. Structural information about a system is most
20 > readily obtained through these types of methods.
21  
22   Although the two techniques employed seem dissimilar, they are both
23   linked by the overarching principles of Statistical
24 < Thermodynamics. Statistical Thermodynamics governs the behavior of
24 > Mechanics. Statistical Meachanics governs the behavior of
25   both classes of simulations and dictates what each method can and
26   cannot do. When investigating a system, one most first analyze what
27   thermodynamic properties of the system are being probed, then chose
# Line 31 | Line 31 | follows is a brief derivation of Boltzmann weighted st
31  
32   The following section serves as a brief introduction to some of the
33   Statistical Mechanics concepts present in this dissertation.  What
34 < follows is a brief derivation of Boltzmann weighted statistics, and an
34 > follows is a brief derivation of Boltzmann weighted statistics and an
35   explanation of how one can use the information to calculate an
36   observable for a system.  This section then concludes with a brief
37   discussion of the ergodic hypothesis and its relevance to this
# Line 39 | Line 39 | Consider a system, $\gamma$, with some total energy,,
39  
40   \subsection{\label{introSec:boltzman}Boltzmann weighted statistics}
41  
42 < Consider a system, $\gamma$, with some total energy,, $E_{\gamma}$.
43 < Let $\Omega(E_{\gamma})$ represent the number of degenerate ways
44 < $\boldsymbol{\Gamma}$, the collection of positions and conjugate
45 < momenta of system $\gamma$, can be configured to give
46 < $E_{\gamma}$. Further, if $\gamma$ is in contact with a bath system
47 < where energy is exchanged between the two systems, $\Omega(E)$, where
48 < $E$ is the total energy of both systems, can be represented as
42 > Consider a system, $\gamma$, with total energy $E_{\gamma}$.  Let
43 > $\Omega(E_{\gamma})$ represent the number of degenerate ways
44 > $\boldsymbol{\Gamma}\{r_1,r_2,\ldots r_n,p_1,p_2,\ldots p_n\}$, the
45 > collection of positions and conjugate momenta of system $\gamma$, can
46 > be configured to give $E_{\gamma}$. Further, if $\gamma$ is a subset
47 > of a larger system, $\boldsymbol{\Lambda}\{E_1,E_2,\ldots
48 > E_{\gamma},\ldots E_n\}$, the total degeneracy of the system can be
49 > expressed as,
50   \begin{equation}
51 + \Omega(\boldsymbol{\Lambda}) = \Omega(E_1) \times \Omega(E_2) \times \ldots
52 +        \Omega(E_{\gamma}) \times \ldots \Omega(E_n)
53 + \label{introEq:SM0.1}
54 + \end{equation}
55 + This multiplicative combination of degeneracies is illustrated in
56 + Fig.~\ref{introFig:degenProd}.
57 +
58 + \begin{figure}
59 + \centering
60 + \includegraphics[width=\linewidth]{omegaFig.eps}
61 + \caption[An explanation of the combination of degeneracies]{Systems A and B both have three energy levels and two indistinguishable particles. When the total energy is 2, there are two ways for each system to disperse the energy. However, for system C, the superset of A and B, the total degeneracy is the product of the degeneracy of each system. In this case $\Omega(\text{C})$ is 4.}
62 + \label{introFig:degenProd}
63 + \end{figure}
64 +
65 + Next, consider the specific case of $\gamma$ in contact with a
66 + bath. Exchange of energy is allowed between the bath and the system,
67 + subject to the constraint that the total energy
68 + ($E_{\text{bath}}+E_{\gamma}$) remain constant. $\Omega(E)$, where $E$
69 + is the total energy of both systems, can be represented as
70 + \begin{equation}
71   \Omega(E) = \Omega(E_{\gamma}) \times \Omega(E - E_{\gamma})
72   \label{introEq:SM1}
73   \end{equation}
# Line 57 | Line 78 | degenerative configurations in $E$. \cite{Frenkel1996}
78   \end{equation}
79  
80   The solution to Eq.~\ref{introEq:SM2} maximizes the number of
81 < degenerative configurations in $E$. \cite{Frenkel1996}
81 > degenerate configurations in $E$. \cite{Frenkel1996}
82   This gives
83   \begin{equation}
84   \frac{\partial \ln \Omega(E)}{\partial E_{\gamma}} = 0 =
# Line 176 | Line 197 | One last important consideration is that of ergodicity
197  
198   \subsection{\label{introSec:ergodic}The Ergodic Hypothesis}
199  
200 < One last important consideration is that of ergodicity. Ergodicity is
201 < the assumption that given an infinite amount of time, a system will
202 < visit every available point in phase space.\cite{Frenkel1996} For most
203 < systems, this is a valid assumption, except in cases where the system
204 < may be trapped in a local feature (\emph{e.g.}~glasses). When valid,
205 < ergodicity allows the unification of a time averaged observation and
206 < an ensemble averaged one. If an observation is averaged over a
207 < sufficiently long time, the system is assumed to visit all
208 < appropriately available points in phase space, giving a properly
209 < weighted statistical average. This allows the researcher freedom of
210 < choice when deciding how best to measure a given observable.  When an
211 < ensemble averaged approach seems most logical, the Monte Carlo
212 < techniques described in Sec.~\ref{introSec:monteCarlo} can be utilized.
213 < Conversely, if a problem lends itself to a time averaging approach,
214 < the Molecular Dynamics techniques in Sec.~\ref{introSec:MD} can be
215 < employed.
200 > In the case of a Molecular Dynamics simulation, rather than
201 > calculating an ensemble average integral over phase space as in
202 > Eq.~\ref{introEq:SM15}, it becomes easier to caclulate the time
203 > average of an observable. Namely,
204 > \begin{equation}
205 > \langle A \rangle_t = \frac{1}{\tau}
206 >        \int_0^{\tau} A[\boldsymbol{\Gamma}(t)]\,dt
207 > \label{introEq:SM16}
208 > \end{equation}
209 > Where the value of an observable is averaged over the length of time
210 > that the simulation is run. This type of measurement mirrors the
211 > experimental measurement of an observable. In an experiment, the
212 > instrument analyzing the system must average its observation of the
213 > finite time of the measurement. What is required then, is a principle
214 > to relate the time average to the ensemble average. This is the
215 > ergodic hypothesis.
216  
217 + Ergodicity is the assumption that given an infinite amount of time, a
218 + system will visit every available point in phase
219 + space.\cite{Frenkel1996} For most systems, this is a valid assumption,
220 + except in cases where the system may be trapped in a local feature
221 + (\emph{e.g.}~glasses). When valid, ergodicity allows the unification
222 + of a time averaged observation and an ensemble averaged one. If an
223 + observation is averaged over a sufficiently long time, the system is
224 + assumed to visit all appropriately available points in phase space,
225 + giving a properly weighted statistical average. This allows the
226 + researcher freedom of choice when deciding how best to measure a given
227 + observable.  When an ensemble averaged approach seems most logical,
228 + the Monte Carlo techniques described in Sec.~\ref{introSec:monteCarlo}
229 + can be utilized.  Conversely, if a problem lends itself to a time
230 + averaging approach, the Molecular Dynamics techniques in
231 + Sec.~\ref{introSec:MD} can be employed.
232 +
233   \section{\label{introSec:monteCarlo}Monte Carlo Simulations}
234  
235   The Monte Carlo method was developed by Metropolis and Ulam for their
# Line 210 | Line 247 | I = (b-a)\langle f(x) \rangle
247   \end{equation}
248   The equation can be recast as:
249   \begin{equation}
250 < I = (b-a)\langle f(x) \rangle
250 > I = \int^b_a \frac{f(x)}{\rho(x)} \rho(x) dx
251 > \label{introEq:Importance1}
252 > \end{equation}
253 > Where $\rho(x)$ is an arbitrary probability distribution in $x$.  If
254 > one conducts $\tau$ trials selecting a random number, $\zeta_\tau$,
255 > from the distribution $\rho(x)$ on the interval $[a,b]$, then
256 > Eq.~\ref{introEq:Importance1} becomes
257 > \begin{equation}
258 > I= \lim_{\tau \rightarrow \infty}\biggl \langle \frac{f(x)}{\rho(x)} \biggr \rangle_{\text{trials}[a,b]}
259 > \label{introEq:Importance2}
260 > \end{equation}
261 > If $\rho(x)$ is uniformly distributed over the interval $[a,b]$,
262 > \begin{equation}
263 > \rho(x) = \frac{1}{b-a}
264 > \label{introEq:importance2b}
265 > \end{equation}
266 > then the integral can be rewritten as
267 > \begin{equation}
268 > I = (b-a)\lim_{\tau \rightarrow \infty}
269 >        \langle f(x) \rangle_{\text{trials}[a,b]}
270   \label{eq:MCex2}
271   \end{equation}
216 Where $\langle f(x) \rangle$ is the unweighted average over the interval
217 $[a,b]$. The calculation of the integral could then be solved by
218 randomly choosing points along the interval $[a,b]$ and calculating
219 the value of $f(x)$ at each point. The accumulated average would then
220 approach $I$ in the limit where the number of trials is infinitely
221 large.
272  
273   However, in Statistical Mechanics, one is typically interested in
274   integrals of the form:
# Line 240 | Line 290 | Importance Sampling is a method where one selects a di
290   average. This is where importance sampling comes into
291   play.\cite{allen87:csl}
292  
293 < Importance Sampling is a method where one selects a distribution from
294 < which the random configurations are chosen in order to more
295 < efficiently calculate the integral.\cite{Frenkel1996} Consider again
296 < Eq.~\ref{eq:MCex1} rewritten to be:
247 < \begin{equation}
248 < I = \int^b_a \frac{f(x)}{\rho(x)} \rho(x) dx
249 < \label{introEq:Importance1}
250 < \end{equation}
251 < Where $\rho(x)$ is an arbitrary probability distribution in $x$.  If
252 < one conducts $\tau$ trials selecting a random number, $\zeta_\tau$,
253 < from the distribution $\rho(x)$ on the interval $[a,b]$, then
254 < Eq.~\ref{introEq:Importance1} becomes
255 < \begin{equation}
256 < I= \biggl \langle \frac{f(x)}{\rho(x)} \biggr \rangle_{\text{trials}}
257 < \label{introEq:Importance2}
258 < \end{equation}
259 < Looking at Eq.~\ref{eq:mcEnsAvg}, and realizing
293 > Importance sampling is a method where the distribution, from which the
294 > random configurations are chosen, is selected in such a way as to
295 > efficiently sample the integral in question.  Looking at
296 > Eq.~\ref{eq:mcEnsAvg}, and realizing
297   \begin {equation}
298   \rho_{kT}(\mathbf{r}^N) =
299          \frac{e^{-\beta V(\mathbf{r}^N)}}
# Line 280 | Line 317 | Eq.~\ref{introEq:Importance4} becomes
317   By selecting $\rho(\mathbf{r}^N)$ to be $\rho_{kT}(\mathbf{r}^N)$
318   Eq.~\ref{introEq:Importance4} becomes
319   \begin{equation}
320 < \langle A \rangle = \langle A(\mathbf{r}^N) \rangle_{\text{trials}}
320 > \langle A \rangle = \langle A(\mathbf{r}^N) \rangle_{kT}
321   \label{introEq:Importance5}
322   \end{equation}
323   The difficulty is selecting points $\mathbf{r}^N$ such that they are
324   sampled from the distribution $\rho_{kT}(\mathbf{r}^N)$.  A solution
325 < was proposed by Metropolis et al.\cite{metropolis:1953} which involved
325 > was proposed by Metropolis \emph{et al}.\cite{metropolis:1953} which involved
326   the use of a Markov chain whose limiting distribution was
327   $\rho_{kT}(\mathbf{r}^N)$.
328  
# Line 398 | Line 435 | diffusion constants, velocity auto correlations, foldi
435   integrated in order to obtain information about both the positions and
436   momentum of a system, allowing the calculation of not only
437   configurational observables, but momenta dependent ones as well:
438 < diffusion constants, velocity auto correlations, folding/unfolding
439 < events, etc.  Due to the principle of ergodicity,
438 > diffusion constants, relaxation events, folding/unfolding
439 > events, etc. With the time dependent information gained from a
440 > Molecular Dynamics simulation, one can also calculate time correlation
441 > functions of the form\cite{Hansen86}
442 > \begin{equation}
443 > \langle A(t)\,A(0)\rangle = \lim_{\tau\rightarrow\infty} \frac{1}{\tau}
444 >        \int_0^{\tau} A(t+t^{\prime})\,A(t^{\prime})\,dt^{\prime}
445 > \label{introEq:timeCorr}
446 > \end{equation}
447 > These correlations can be used to measure fundamental time constants
448 > of a system, such as diffusion constants from the velocity
449 > autocorrelation or dipole relaxation times from the dipole
450 > autocorrelation.  Due to the principle of ergodicity,
451   Sec.~\ref{introSec:ergodic}, the average of these observables over the
452   time period of the simulation are taken to be the ensemble averages
453   for the system.
454  
455   The choice of when to use molecular dynamics over Monte Carlo
456   techniques, is normally decided by the observables in which the
457 < researcher is interested.  If the observables depend on momenta in
457 > researcher is interested.  If the observables depend on time in
458   any fashion, then the only choice is molecular dynamics in some form.
459   However, when the observable is dependent only on the configuration,
460 < then most of the time Monte Carlo techniques will be more efficient.
460 > then for most small systems, Monte Carlo techniques will be more efficient.
461  
462   The focus of research in the second half of this dissertation is
463   centered around the dynamic properties of phospholipid bilayers,
# Line 481 | Line 529 | addition, this sets that any given particle pair has a
529   simulation box on an infinite lattice in Cartesian space.  Any given
530   particle leaving the simulation box on one side will have an image of
531   itself enter on the opposite side (see Fig.~\ref{introFig:pbc}).  In
532 < addition, this sets that any given particle pair has an image, real or
533 < periodic, within $fix$ of each other.  A discussion of the method used
534 < to calculate the periodic image can be found in
532 > addition, this sets that any two particles have an image, real or
533 > periodic, within $\text{box}/2$ of each other.  A discussion of the
534 > method used to calculate the periodic image can be found in
535   Sec.\ref{oopseSec:pbc}.
536  
537   \begin{figure}
# Line 499 | Line 547 | $r_{\text{cut}}$ has a maximum value of $\text{box}/2$
547   efficiency of the force evaluation, as particles farther than a
548   predetermined distance, $r_{\text{cut}}$, are not included in the
549   calculation.\cite{Frenkel1996} In a simulation with periodic images,
550 < $r_{\text{cut}}$ has a maximum value of $\text{box}/2$.
551 < Fig.~\ref{introFig:rMax} illustrates how when using an
552 < $r_{\text{cut}}$ larger than this value, or in the extreme limit of no
553 < $r_{\text{cut}}$ at all, the corners of the simulation box are
554 < unequally weighted due to the lack of particle images in the $x$, $y$,
555 < or $z$ directions past a distance of $\text{box} / 2$.
550 > there are two methods to choose from, both with their own cutoff
551 > limits. In the minimum image convention, $r_{\text{cut}}$ has a
552 > maximum value of $\text{box}/2$. This is because each atom has only
553 > one image that is seen by another atom, and further the image used is
554 > the one that minimizes the distance between the two atoms. A system of
555 > wrapped images about a central atom therefore has a maximum length
556 > scale of box on a side (Fig.~\ref{introFig:rMaxMin}). The second
557 > convention, multiple image convention, has a maximum $r_{\text{cut}}$
558 > of box. Here multiple images of each atom are replicated in the
559 > periodic cells surrounding the central atom, this causes the atom to
560 > see multiple copies of several atoms. If the cutoff radius is larger
561 > than box, however, then the atom will see an image of itself, and
562 > attempt to calculate an unphysical self-self force interaction
563 > (Fig.~\ref{introFig:rMaxMult}). Due to the increased complexity and
564 > commputaional ineffeciency of the multiple image method, the minimum
565 > image method is the periodic method used throughout this research.
566  
567   \begin{figure}
568   \centering
569   \includegraphics[width=\linewidth]{rCutMaxFig.eps}
570 < \caption[An explanation of $r_{\text{cut}}$]{The yellow atom has all other images wrapped to itself as the center. If $r_{\text{cut}}=\text{box}/2$, then the distribution is uniform (blue atoms). However, when $r_{\text{cut}}>\text{box}/2$ the corners are disproportionately weighted (green atoms) vs the axial directions (shaded regions).}
571 < \label{introFig:rMax}
570 > \caption[An explanation of minimum image convention]{The yellow atom has all other images wrapped to itself as the center. If $r_{\text{cut}}=\text{box}/2$, then the distribution is uniform (blue atoms). However, when $r_{\text{cut}}>\text{box}/2$ the corners are disproportionately weighted (green atoms) vs the axial directions (shaded regions).}
571 > \label{introFig:rMaxMin}
572   \end{figure}
573  
574 + \begin{figure}
575 + \centering
576 + \includegraphics[width=\linewidth]{rCutMaxMultFig.eps}
577 + \caption[An explanation of multiple image convention]{The yellow atom is the central wrapping point. The blue atoms are the minimum images of the system about the central atom. The boxes with the green atoms are multiple images of the central box. If $r_{\text{cut}} \geq \{text{box}$ then the central atom sees multiple images of itself (red atom), creating a self-self force evaluation.}
578 + \label{introFig:rMaxMult}
579 + \end{figure}
580 +
581   With the use of an $r_{\text{cut}}$, however, comes a discontinuity in
582   the potential energy curve (Fig.~\ref{introFig:shiftPot}). To fix this
583   discontinuity, one calculates the potential energy at the
# Line 557 | Line 622 | Adding together Eq.~\ref{introEq:verletForward} and
622          \mathcal{O}(\Delta t^4)
623   \label{introEq:verletBack}
624   \end{equation}
625 < Adding together Eq.~\ref{introEq:verletForward} and
625 > Where $m$ is the mass of the particle, $q(t)$ is the position at time
626 > $t$, $v(t)$ the velocity, and $F(t)$ the force acting on the
627 > particle. Adding together Eq.~\ref{introEq:verletForward} and
628   Eq.~\ref{introEq:verletBack} results in,
629   \begin{equation}
630 < eq here
630 > q(t+\Delta t)+q(t-\Delta t) =
631 >        2q(t) + \frac{F(t)}{m}\Delta t^2 + \mathcal{O}(\Delta t^4)
632   \label{introEq:verletSum}
633   \end{equation}
634   Or equivalently,
635   \begin{equation}
636 < eq here
636 > q(t+\Delta t) \approx
637 >        2q(t) - q(t-\Delta t) + \frac{F(t)}{m}\Delta t^2
638   \label{introEq:verletFinal}
639   \end{equation}
640   Which contains an error in the estimate of the new positions on the
# Line 573 | Line 642 | with a velocity reformulation of the Verlet method.\ci
642  
643   In practice, however, the simulations in this research were integrated
644   with a velocity reformulation of the Verlet method.\cite{allen87:csl}
645 < \begin{equation}
646 < eq here
647 < \label{introEq:MDvelVerletPos}
648 < \end{equation}
649 < \begin{equation}
581 < eq here
645 > \begin{align}
646 > q(t+\Delta t) &= q(t) + v(t)\Delta t + \frac{F(t)}{2m}\Delta t^2 %
647 > \label{introEq:MDvelVerletPos} \\%
648 > %
649 > v(t+\Delta t) &= v(t) + \frac{\Delta t}{2m}[F(t) + F(t+\Delta t)] %
650   \label{introEq:MDvelVerletVel}
651 < \end{equation}
651 > \end{align}
652   The original Verlet algorithm can be regained by substituting the
653   velocity back into Eq.~\ref{introEq:MDvelVerletPos}.  The Verlet
654   formulations are chosen in this research because the algorithms have
# Line 602 | Line 670 | hypothesis (Sec.~\ref{introSec:StatThermo}), it is kno
670   reversible.  The fact that it shadows the true Hamiltonian in phase
671   space is acceptable in actual simulations as one is interested in the
672   ensemble average of the observable being measured.  From the ergodic
673 < hypothesis (Sec.~\ref{introSec:StatThermo}), it is known that the time
673 > hypothesis (Sec.~\ref{introSec:statThermo}), it is known that the time
674   average will match the ensemble average, therefore two similar
675   trajectories in phase space should give matching statistical averages.
676  
677   \subsection{\label{introSec:MDfurther}Further Considerations}
678 +
679   In the simulations presented in this research, a few additional
680   parameters are needed to describe the motions.  The simulations
681 < involving water and phospholipids in Ch.~\ref{chaptLipids} are
681 > involving water and phospholipids in Ch.~\ref{chapt:lipid} are
682   required to integrate the equations of motions for dipoles on atoms.
683   This involves an additional three parameters be specified for each
684   dipole atom: $\phi$, $\theta$, and $\psi$.  These three angles are
685   taken to be the Euler angles, where $\phi$ is a rotation about the
686   $z$-axis, and $\theta$ is a rotation about the new $x$-axis, and
687   $\psi$ is a final rotation about the new $z$-axis (see
688 < Fig.~\ref{introFig:euleerAngles}).  This sequence of rotations can be
689 < accumulated into a single $3 \times 3$ matrix $\mathbf{A}$
688 > Fig.~\ref{introFig:eulerAngles}).  This sequence of rotations can be
689 > accumulated into a single $3 \times 3$ matrix, $\mathbf{A}$,
690   defined as follows:
691   \begin{equation}
692 < eq here
692 > \mathbf{A} =
693 > \begin{bmatrix}
694 >        \cos\phi\cos\psi-\sin\phi\cos\theta\sin\psi &%
695 >        \sin\phi\cos\psi+\cos\phi\cos\theta\sin\psi &%
696 >        \sin\theta\sin\psi \\%
697 >        %
698 >        -\cos\phi\sin\psi-\sin\phi\cos\theta\cos\psi &%
699 >        -\sin\phi\sin\psi+\cos\phi\cos\theta\cos\psi &%
700 >        \sin\theta\cos\psi \\%
701 >        %
702 >        \sin\phi\sin\theta &%
703 >        -\cos\phi\sin\theta &%
704 >        \cos\theta
705 > \end{bmatrix}
706   \label{introEq:EulerRotMat}
707   \end{equation}
708  
709 < The equations of motion for Euler angles can be written down as
710 < \cite{allen87:csl}
711 < \begin{equation}
712 < eq here
713 < \label{introEq:MDeuleeerPsi}
714 < \end{equation}
709 > \begin{figure}
710 > \centering
711 > \includegraphics[width=\linewidth]{eulerRotFig.eps}
712 > \caption[Euler rotation of Cartesian coordinates]{The rotation scheme for Euler angles. First is a rotation of $\phi$ about the $z$ axis (blue rotation). Next is a rotation of $\theta$ about the new $x$ axis (green rotation). Lastly is a final rotation of $\psi$ about the new $z$ axis (red rotation).}
713 > \label{introFig:eulerAngles}
714 > \end{figure}
715 >
716 > The equations of motion for Euler angles can be written down
717 > as\cite{allen87:csl}
718 > \begin{align}
719 > \dot{\phi} &= -\omega^s_x \frac{\sin\phi\cos\theta}{\sin\theta} +
720 >        \omega^s_y \frac{\cos\phi\cos\theta}{\sin\theta} +
721 >        \omega^s_z
722 > \label{introEq:MDeulerPhi} \\%
723 > %
724 > \dot{\theta} &= \omega^s_x \cos\phi + \omega^s_y \sin\phi
725 > \label{introEq:MDeulerTheta} \\%
726 > %
727 > \dot{\psi} &= \omega^s_x \frac{\sin\phi}{\sin\theta} -
728 >        \omega^s_y \frac{\cos\phi}{\sin\theta}
729 > \label{introEq:MDeulerPsi}
730 > \end{align}
731   Where $\omega^s_i$ is the angular velocity in the lab space frame
732   along Cartesian coordinate $i$.  However, a difficulty arises when
733   attempting to integrate Eq.~\ref{introEq:MDeulerPhi} and
734   Eq.~\ref{introEq:MDeulerPsi}. The $\frac{1}{\sin \theta}$ present in
735   both equations means there is a non-physical instability present when
736 < $\theta$ is 0 or $\pi$.
737 <
738 < To correct for this, the simulations integrate the rotation matrix,
739 < $\mathbf{A}$, directly, thus avoiding the instability.
642 < This method was proposed by Dullwebber
643 < \emph{et. al.}\cite{Dullwebber:1997}, and is presented in
736 > $\theta$ is 0 or $\pi$. To correct for this, the simulations integrate
737 > the rotation matrix, $\mathbf{A}$, directly, thus avoiding the
738 > instability.  This method was proposed by Dullweber
739 > \emph{et. al.}\cite{Dullweber1997}, and is presented in
740   Sec.~\ref{introSec:MDsymplecticRot}.
741  
742 < \subsubsection{\label{introSec:MDliouville}Liouville Propagator}
742 > \subsection{\label{introSec:MDliouville}Liouville Propagator}
743  
744   Before discussing the integration of the rotation matrix, it is
745   necessary to understand the construction of a ``good'' integration
746   scheme.  It has been previously
747 < discussed(Sec.~\ref{introSec:MDintegrate}) how it is desirable for an
747 > discussed(Sec.~\ref{introSec:mdIntegrate}) how it is desirable for an
748   integrator to be symplectic, or time reversible.  The following is an
749   outline of the Trotter factorization of the Liouville Propagator as a
750 < scheme for generating symplectic integrators. \cite{Tuckerman:1992}
750 > scheme for generating symplectic integrators.\cite{Tuckerman92}
751  
752   For a system with $f$ degrees of freedom the Liouville operator can be
753   defined as,
754   \begin{equation}
755 < eq here
755 > iL=\sum^f_{j=1} \biggl [\dot{q}_j\frac{\partial}{\partial q_j} +
756 >        F_j\frac{\partial}{\partial p_j} \biggr ]
757   \label{introEq:LiouvilleOperator}
758   \end{equation}
759 < Here, $r_j$ and $p_j$ are the position and conjugate momenta of a
760 < degree of freedom, and $f_j$ is the force on that degree of freedom.
759 > Here, $q_j$ and $p_j$ are the position and conjugate momenta of a
760 > degree of freedom, and $F_j$ is the force on that degree of freedom.
761   $\Gamma$ is defined as the set of all positions and conjugate momenta,
762 < $\{r_j,p_j\}$, and the propagator, $U(t)$, is defined
762 > $\{q_j,p_j\}$, and the propagator, $U(t)$, is defined
763   \begin {equation}
764 < eq here
764 > U(t) = e^{iLt}
765   \label{introEq:Lpropagator}
766   \end{equation}
767   This allows the specification of $\Gamma$ at any time $t$ as
768   \begin{equation}
769 < eq here
769 > \Gamma(t) = U(t)\Gamma(0)
770   \label{introEq:Lp2}
771   \end{equation}
772   It is important to note, $U(t)$ is a unitary operator meaning
# Line 680 | Line 777 | Trotter theorem to yield
777  
778   Decomposing $L$ into two parts, $iL_1$ and $iL_2$, one can use the
779   Trotter theorem to yield
780 < \begin{equation}
781 < eq here
782 < \label{introEq:Lp4}
783 < \end{equation}
784 < Where $\Delta t = \frac{t}{P}$.
780 > \begin{align}
781 > e^{iLt} &= e^{i(L_1 + L_2)t} \notag \\%
782 > %
783 >        &= \biggl [ e^{i(L_1 +L_2)\frac{t}{P}} \biggr]^P \notag \\%
784 > %
785 >        &= \biggl [ e^{iL_1\frac{\Delta t}{2}}\, e^{iL_2\Delta t}\,
786 >        e^{iL_1\frac{\Delta t}{2}} \biggr ]^P +
787 >        \mathcal{O}\biggl (\frac{t^3}{P^2} \biggr ) \label{introEq:Lp4}
788 > \end{align}
789 > Where $\Delta t = t/P$.
790   With this, a discrete time operator $G(\Delta t)$ can be defined:
791 < \begin{equation}
792 < eq here
791 > \begin{align}
792 > G(\Delta t) &= e^{iL_1\frac{\Delta t}{2}}\, e^{iL_2\Delta t}\,
793 >        e^{iL_1\frac{\Delta t}{2}} \notag \\%
794 > %
795 >        &= U_1 \biggl ( \frac{\Delta t}{2} \biggr )\, U_2 ( \Delta t )\,
796 >        U_1 \biggl ( \frac{\Delta t}{2} \biggr )
797   \label{introEq:Lp5}
798 < \end{equation}
799 < Because $U_1(t)$ and $U_2(t)$ are unitary, $G|\Delta t)$ is also
798 > \end{align}
799 > Because $U_1(t)$ and $U_2(t)$ are unitary, $G(\Delta t)$ is also
800   unitary.  Meaning an integrator based on this factorization will be
801   reversible in time.
802  
803   As an example, consider the following decomposition of $L$:
804 + \begin{align}
805 + iL_1 &= \dot{q}\frac{\partial}{\partial q}%
806 + \label{introEq:Lp6a} \\%
807 + %
808 + iL_2 &= F(q)\frac{\partial}{\partial p}%
809 + \label{introEq:Lp6b}
810 + \end{align}
811 + This leads to propagator $G( \Delta t )$ as,
812   \begin{equation}
813 < eq here
814 < \label{introEq:Lp6}
813 > G(\Delta t) =  e^{\frac{\Delta t}{2} F(q)\frac{\partial}{\partial p}} \,
814 >        e^{\Delta t\,\dot{q}\frac{\partial}{\partial q}} \,
815 >        e^{\frac{\Delta t}{2} F(q)\frac{\partial}{\partial p}}
816 > \label{introEq:Lp7}
817   \end{equation}
818 < Operating $G(\Delta t)$ on $\Gamma)0)$, and utilizing the operator property
818 > Operating $G(\Delta t)$ on $\Gamma(0)$, and utilizing the operator property
819   \begin{equation}
820 < eq here
820 > e^{c\frac{\partial}{\partial x}}\, f(x) = f(x+c)
821   \label{introEq:Lp8}
822   \end{equation}
823 < Where $c$ is independent of $q$.  One obtains the following:  
824 < \begin{equation}
825 < eq here
826 < \label{introEq:Lp8}
827 < \end{equation}
823 > Where $c$ is independent of $x$.  One obtains the following:  
824 > \begin{align}
825 > \dot{q}\biggl (\frac{\Delta t}{2}\biggr ) &=
826 >        \dot{q}(0) + \frac{\Delta t}{2m}\, F[q(0)] \label{introEq:Lp9a}\\%
827 > %
828 > q(\Delta t) &= q(0) + \Delta t\, \dot{q}\biggl (\frac{\Delta t}{2}\biggr )%
829 >        \label{introEq:Lp9b}\\%
830 > %
831 > \dot{q}(\Delta t) &= \dot{q}\biggl (\frac{\Delta t}{2}\biggr ) +
832 >        \frac{\Delta t}{2m}\, F[q(0)] \label{introEq:Lp9c}
833 > \end{align}
834   Or written another way,
835 < \begin{equation}
836 < eq here
837 < \label{intorEq:Lp9}
838 < \end{equation}
835 > \begin{align}
836 > q(t+\Delta t) &= q(0) + \dot{q}(0)\Delta t +
837 >        \frac{F[q(0)]}{m}\frac{\Delta t^2}{2} %
838 > \label{introEq:Lp10a} \\%
839 > %
840 > \dot{q}(\Delta t) &= \dot{q}(0) + \frac{\Delta t}{2m}
841 >        \biggl [F[q(0)] + F[q(\Delta t)] \biggr] %
842 > \label{introEq:Lp10b}
843 > \end{align}
844   This is the velocity Verlet formulation presented in
845 < Sec.~\ref{introSec:MDintegrate}.  Because this integration scheme is
845 > Sec.~\ref{introSec:mdIntegrate}.  Because this integration scheme is
846   comprised of unitary propagators, it is symplectic, and therefore area
847   preserving in phase space.  From the preceding factorization, one can
848   see that the integration of the equations of motion would follow:
# Line 729 | Line 856 | see that the integration of the equations of motion wo
856   \item Repeat from step 1 with the new position, velocities, and forces assuming the roles of the initial values.
857   \end{enumerate}
858  
859 < \subsubsection{\label{introSec:MDsymplecticRot} Symplectic Propagation of the Rotation Matrix}
859 > \subsection{\label{introSec:MDsymplecticRot} Symplectic Propagation of the Rotation Matrix}
860  
861   Based on the factorization from the previous section,
862 < Dullweber\emph{et al.}\cite{Dullweber:1997}~ proposed a scheme for the
862 > Dullweber\emph{et al}.\cite{Dullweber1997}~ proposed a scheme for the
863   symplectic propagation of the rotation matrix, $\mathbf{A}$, as an
864   alternative method for the integration of orientational degrees of
865   freedom. The method starts with a straightforward splitting of the
866   Liouville operator:
867 < \begin{equation}
868 < eq here
869 < \label{introEq:SR1}
870 < \end{equation}
871 < Where $\boldsymbol{\tau}(\mathbf{A})$ are the torques of the system
872 < due to the configuration, and $\boldsymbol{/pi}$ are the conjugate
867 > \begin{align}
868 > iL_{\text{pos}} &= \dot{q}\frac{\partial}{\partial q} +
869 >        \mathbf{\dot{A}}\frac{\partial}{\partial \mathbf{A}}
870 > \label{introEq:SR1a} \\%
871 > %
872 > iL_F &= F(q)\frac{\partial}{\partial p}
873 > \label{introEq:SR1b} \\%
874 > iL_{\tau} &= \tau(\mathbf{A})\frac{\partial}{\partial \pi}
875 > \label{introEq:SR1b} \\%
876 > \end{align}
877 > Where $\tau(\mathbf{A})$ is the torque of the system
878 > due to the configuration, and $\pi$ is the conjugate
879   angular momenta of the system. The propagator, $G(\Delta t)$, becomes
880   \begin{equation}
881 < eq here
881 > G(\Delta t) = e^{\frac{\Delta t}{2} F(q)\frac{\partial}{\partial p}} \,
882 >        e^{\frac{\Delta t}{2} \tau(\mathbf{A})\frac{\partial}{\partial \pi}} \,
883 >        e^{\Delta t\,iL_{\text{pos}}} \,
884 >        e^{\frac{\Delta t}{2} \tau(\mathbf{A})\frac{\partial}{\partial \pi}} \,
885 >        e^{\frac{\Delta t}{2} F(q)\frac{\partial}{\partial p}}
886   \label{introEq:SR2}
887   \end{equation}
888   Propagation of the linear and angular momenta follows as in the Verlet
889   scheme.  The propagation of positions also follows the Verlet scheme
890   with the addition of a further symplectic splitting of the rotation
891 < matrix propagation, $\mathcal{G}_{\text{rot}}(\Delta t)$.
891 > matrix propagation, $\mathcal{U}_{\text{rot}}(\Delta t)$, within
892 > $U_{\text{pos}}(\Delta t)$.
893   \begin{equation}
894 < eq here
894 > \mathcal{U}_{\text{rot}}(\Delta t) =
895 >        \mathcal{U}_x \biggl(\frac{\Delta t}{2}\biggr)\,
896 >        \mathcal{U}_y \biggl(\frac{\Delta t}{2}\biggr)\,
897 >        \mathcal{U}_z (\Delta t)\,
898 >        \mathcal{U}_y \biggl(\frac{\Delta t}{2}\biggr)\,
899 >        \mathcal{U}_x \biggl(\frac{\Delta t}{2}\biggr)\,
900   \label{introEq:SR3}
901   \end{equation}
902 < Where $\mathcal{G}_j$ is a unitary rotation of $\mathbf{A}$ and
903 < $\boldsymbol{\pi}$ about each axis $j$.  As all propagations are now
902 > Where $\mathcal{U}_j$ is a unitary rotation of $\mathbf{A}$ and
903 > $\pi$ about each axis $j$.  As all propagations are now
904   unitary and symplectic, the entire integration scheme is also
905   symplectic and time reversible.
906  
907   \section{\label{introSec:layout}Dissertation Layout}
908  
909 < This dissertation is divided as follows:Chapt.~\ref{chapt:RSA}
909 > This dissertation is divided as follows:Ch.~\ref{chapt:RSA}
910   presents the random sequential adsorption simulations of related
911   pthalocyanines on a gold (111) surface. Ch.~\ref{chapt:OOPSE}
912   is about the writing of the molecular dynamics simulation package
913 < {\sc oopse}, Ch.~\ref{chapt:lipid} regards the simulations of
914 < phospholipid bilayers using a mesoscale model, and lastly,
913 > {\sc oopse}. Ch.~\ref{chapt:lipid} regards the simulations of
914 > phospholipid bilayers using a mesoscale model. And lastly,
915   Ch.~\ref{chapt:conclusion} concludes this dissertation with a
916   summary of all results. The chapters are arranged in chronological
917   order, and reflect the progression of techniques I employed during my
918   research.  
919  
920 < The chapter concerning random sequential adsorption
921 < simulations is a study in applying the principles of theoretical
922 < research in order to obtain a simple model capable of explaining the
923 < results.  My advisor, Dr. Gezelter, and I were approached by a
924 < colleague, Dr. Lieberman, about possible explanations for partial
925 < coverage of a gold surface by a particular compound of hers. We
926 < suggested it might be due to the statistical packing fraction of disks
927 < on a plane, and set about to simulate this system.  As the events in
928 < our model were not dynamic in nature, a Monte Carlo method was
929 < employed.  Here, if a molecule landed on the surface without
930 < overlapping another, then its landing was accepted.  However, if there
931 < was overlap, the landing we rejected and a new random landing location
932 < was chosen.  This defined our acceptance rules and allowed us to
933 < construct a Markov chain whose limiting distribution was the surface
934 < coverage in which we were interested.
920 > The chapter concerning random sequential adsorption simulations is a
921 > study in applying Statistical Mechanics simulation techniques in order
922 > to obtain a simple model capable of explaining the results.  My
923 > advisor, Dr. Gezelter, and I were approached by a colleague,
924 > Dr. Lieberman, about possible explanations for the  partial coverage of a
925 > gold surface by a particular compound of hers. We suggested it might
926 > be due to the statistical packing fraction of disks on a plane, and
927 > set about to simulate this system.  As the events in our model were
928 > not dynamic in nature, a Monte Carlo method was employed.  Here, if a
929 > molecule landed on the surface without overlapping another, then its
930 > landing was accepted.  However, if there was overlap, the landing we
931 > rejected and a new random landing location was chosen.  This defined
932 > our acceptance rules and allowed us to construct a Markov chain whose
933 > limiting distribution was the surface coverage in which we were
934 > interested.
935  
936   The following chapter, about the simulation package {\sc oopse},
937   describes in detail the large body of scientific code that had to be
938 < written in order to study phospholipid bilayer.  Although there are
938 > written in order to study phospholipid bilayers.  Although there are
939   pre-existing molecular dynamic simulation packages available, none
940   were capable of implementing the models we were developing.{\sc oopse}
941   is a unique package capable of not only integrating the equations of
# Line 804 | Line 947 | This model retains information about solvent ordering
947  
948   Bringing us to Ch.~\ref{chapt:lipid}. Using {\sc oopse}, I have been
949   able to parameterize a mesoscale model for phospholipid simulations.
950 < This model retains information about solvent ordering about the
950 > This model retains information about solvent ordering around the
951   bilayer, as well as information regarding the interaction of the
952 < phospholipid head groups' dipole with each other and the surrounding
952 > phospholipid head groups' dipoles with each other and the surrounding
953   solvent.  These simulations give us insight into the dynamic events
954   that lead to the formation of phospholipid bilayers, as well as
955   provide the foundation for future exploration of bilayer phase

Diff Legend

Removed lines
+ Added lines
< Changed lines
> Changed lines