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 1087 by mmeineke, Fri Mar 5 22:16:34 2004 UTC vs.
Revision 1088 by mmeineke, Sun Mar 7 04:01:29 2004 UTC

# Line 5 | Line 5 | Carlo. Molecular Dynamics simulations integrate the eq
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 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, 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.
8 > Carlo. Molecular Dynamics simulations are carried out by integrating
9 > the equations of motion for a given system of particles, allowing the
10 > researcher to gain insight into the time dependent evolution of a
11 > system. Transport phenomena are readily studied with this simulation
12 > technique, making Molecular Dynamics the main simulation technique
13 > used in this research. Other aspects of the research fall in the
14 > Monte Carlo class of simulations. In Monte Carlo, the configuration
15 > space available to the collection of particles is sampled
16 > stochastically, or randomly. Each configuration is chosen with a given
17 > probability based on the Boltzmann distribution. These types
18 > of simulations are best used to probe properties of a system that are
19 > dependent only on the state of the system. Structural information
20 > about a system is most readily obtained through these types of
21 > methods.
22  
23   Although the two techniques employed seem dissimilar, they are both
24   linked by the overarching principles of Statistical
25 < Mechanics. Statistical Meachanics governs the behavior of
25 > Mechanics. Statistical Mechanics governs the behavior of
26   both classes of simulations and dictates what each method can and
27 < cannot do. When investigating a system, one most first analyze what
28 < thermodynamic properties of the system are being probed, then chose
27 > cannot do. When investigating a system, one must first analyze what
28 > thermodynamic properties of the system are being probed, then choose
29   which method best suits that objective.
30  
31   \section{\label{introSec:statThermo}Statistical Mechanics}
32  
33   The following section serves as a brief introduction to some of the
34 < Statistical Mechanics concepts present in this dissertation.  What
34 > Statistical Mechanics concepts presented in this dissertation.  What
35   follows is a brief derivation of Boltzmann weighted statistics and an
36   explanation of how one can use the information to calculate an
37   observable for a system.  This section then concludes with a brief
# Line 124 | Line 125 | Showing that the partitioning of energy between the tw
125   \beta( E_{\gamma} ) = \beta( E_{\text{bath}} )
126   \label{introEq:SM8}
127   \end{equation}
128 < Showing that the partitioning of energy between the two systems is
129 < actually a process of thermal equilibration.\cite{Frenkel1996}
128 > Eq.~\ref{introEq:SM8} shows that the partitioning of energy between
129 > the two systems is actually a process of thermal
130 > equilibration.\cite{Frenkel1996}
131  
132 < An application of these results is to formulate the form of an
133 < expectation value of an observable, $A$, in the canonical ensemble. In
134 < the canonical ensemble, the number of particles, $N$, the volume, $V$,
135 < and the temperature, $T$, are all held constant while the energy, $E$,
136 < is allowed to fluctuate. Returning to the previous example, the bath
132 > An application of these results is to formulate the expectation value
133 > of an observable, $A$, in the canonical ensemble. In the canonical
134 > ensemble, the number of particles, $N$, the volume, $V$, and the
135 > temperature, $T$, are all held constant while the energy, $E$, is
136 > allowed to fluctuate. Returning to the previous example, the bath
137   system is now an infinitely large thermal bath, whose exchange of
138   energy with the system $\gamma$ holds the temperature constant.  The
139 < partitioning of energy in the bath system is then related to the total
140 < energy of both systems and the fluctuations in $E_{\gamma}$:
139 > partitioning of energy between the bath and the system is then related
140 > to the total energy of both systems and the fluctuations in
141 > $E_{\gamma}$:
142   \begin{equation}
143   \Omega( E_{\gamma} ) = \Omega( E - E_{\gamma} )
144   \label{introEq:SM9}
# Line 148 | Line 151 | an integration over all accessible phase space, $P_{\g
151   \label{introEq:SM10}
152   \end{equation}
153   Where $\int\limits_{\boldsymbol{\Gamma}} d\boldsymbol{\Gamma}$ denotes
154 < an integration over all accessible phase space, $P_{\gamma}$ is the
155 < probability of being in a given phase state and
156 < $A(\boldsymbol{\Gamma})$ is some observable that is a function of the
154 > an integration over all accessible points in phase space, $P_{\gamma}$
155 > is the probability of being in a given phase state and
156 > $A(\boldsymbol{\Gamma})$ is an observable that is a function of the
157   phase state.
158  
159   Because entropy seeks to maximize the number of degenerate states at a
# Line 206 | Line 209 | Where the value of an observable is averaged over the
209          \int_0^{\tau} A[\boldsymbol{\Gamma}(t)]\,dt
210   \label{introEq:SM16}
211   \end{equation}
212 < Where the value of an observable is averaged over the length of time
213 < that the simulation is run. This type of measurement mirrors the
214 < experimental measurement of an observable. In an experiment, the
212 > Where the value of an observable is averaged over the length of time,
213 > $\tau$, that the simulation is run. This type of measurement mirrors
214 > the experimental measurement of an observable. In an experiment, the
215   instrument analyzing the system must average its observation over the
216   finite time of the measurement. What is required then, is a principle
217   to relate the time average to the ensemble average. This is the
# Line 221 | Line 224 | assumed to visit all appropriately available points in
224   (\emph{e.g.}~glasses). When valid, ergodicity allows the unification
225   of a time averaged observation and an ensemble averaged one. If an
226   observation is averaged over a sufficiently long time, the system is
227 < assumed to visit all appropriately available points in phase space,
227 > assumed to visit all energetically accessible points in phase space,
228   giving a properly weighted statistical average. This allows the
229   researcher freedom of choice when deciding how best to measure a given
230   observable.  When an ensemble averaged approach seems most logical,
# Line 234 | Line 237 | named, because it heavily uses random numbers in its
237  
238   The Monte Carlo method was developed by Metropolis and Ulam for their
239   work in fissionable material.\cite{metropolis:1949} The method is so
240 < named, because it heavily uses random numbers in its
241 < solution.\cite{allen87:csl} The Monte Carlo method allows for the
242 < solution of integrals through the stochastic sampling of the values
243 < within the integral. In the simplest case, the evaluation of an
244 < integral would follow a brute force method of
245 < sampling.\cite{Frenkel1996} Consider the following single dimensional
243 < integral:
240 > named, because it uses random numbers extensively.\cite{allen87:csl}
241 > The Monte Carlo method allows for the solution of integrals through
242 > the stochastic sampling of the values within the integral. In the
243 > simplest case, the evaluation of an integral would follow a brute
244 > force method of sampling.\cite{Frenkel1996} Consider the following
245 > single dimensional integral:
246   \begin{equation}
247 < I = f(x)dx
247 > I = \int_a^b f(x)dx
248   \label{eq:MCex1}
249   \end{equation}
250   The equation can be recast as:
# Line 371 | Line 373 | will only yield the limiting distribution again.
373   distribution, and successive applications of the transition matrix
374   will only yield the limiting distribution again.
375   \begin{equation}
376 < \boldsymbol{\rho}_{\text{limit}} = \boldsymbol{\rho}_{\text{initial}}
376 > \boldsymbol{\rho}_{\text{limit}} = \boldsymbol{\rho}_{\text{limit}}
377          \boldsymbol{\Pi}
378   \label{introEq:MCmarkovEquil}
379   \end{equation}
# Line 383 | Line 385 | distribution.  Meaning, that at equilibrium the probab
385   $\boldsymbol{\rho}_{\text{limit}}$ matches the Boltzmann distribution
386   of states.  The method accomplishes this by imposing the strong
387   condition of microscopic reversibility on the equilibrium
388 < distribution.  Meaning, that at equilibrium the probability of going
388 > distribution.  This means that at equilibrium, the probability of going
389   from $m$ to $n$ is the same as going from $n$ to $m$.
390   \begin{equation}
391   \rho_m\pi_{mn} = \rho_n\pi_{nm}
# Line 418 | Line 420 | Metropolis method proceeds as follows
420   \begin{enumerate}
421   \item Generate an initial configuration $\mathbf{r}^N$ which has some finite probability in $\rho_{kT}$.
422   \item Modify $\mathbf{r}^N$, to generate configuration $\mathbf{r^{\prime}}^N$.
423 < \item If the new configuration lowers the energy of the system, accept the move with unity ($\mathbf{r}^N$ becomes $\mathbf{r^{\prime}}^N$).  Otherwise accept with probability $e^{-\beta \Delta \mathcal{U}}$.
423 > \item If the new configuration lowers the energy of the system, accept the move ($\mathbf{r}^N$ becomes $\mathbf{r^{\prime}}^N$).  Otherwise accept with probability $e^{-\beta \Delta \mathcal{U}}$.
424   \item Accumulate the average for the configurational observable of interest.
425   \item Repeat from step 2 until the average converges.
426   \end{enumerate}
# Line 431 | Line 433 | Molecular Dynamics is when the equations of motion for
433   \section{\label{introSec:MD}Molecular Dynamics Simulations}
434  
435   The main simulation tool used in this research is Molecular Dynamics.
436 < Molecular Dynamics is when the equations of motion for a system are
437 < integrated in order to obtain information about both the positions and
438 < momentum of a system, allowing the calculation of not only
439 < configurational observables, but momenta dependent ones as well:
440 < diffusion constants, relaxation events, folding/unfolding
441 < events, etc. With the time dependent information gained from a
442 < Molecular Dynamics simulation, one can also calculate time correlation
443 < functions of the form\cite{Hansen86}
436 > Molecular Dynamics is the numerical integration of the equations of
437 > motion for a system in order to obtain information about the dynamic
438 > changes in the positions and momentum of the constituent particles.
439 > This allows the calculation of not only configurational observables,
440 > but momentum dependent ones as well: diffusion constants, relaxation
441 > events, folding/unfolding events, etc. With the time dependent
442 > information gained from a Molecular Dynamics simulation, one can also
443 > calculate time correlation functions of the form\cite{Hansen86}
444   \begin{equation}
445   \langle A(t)\,A(0)\rangle = \lim_{\tau\rightarrow\infty} \frac{1}{\tau}
446          \int_0^{\tau} A(t+t^{\prime})\,A(t^{\prime})\,dt^{\prime}
# Line 448 | Line 450 | Sec.~\ref{introSec:ergodic}, the average of these obse
450   of a system, such as diffusion constants from the velocity
451   autocorrelation or dipole relaxation times from the dipole
452   autocorrelation.  Due to the principle of ergodicity,
453 < Sec.~\ref{introSec:ergodic}, the average of these observables over the
454 < time period of the simulation are taken to be the ensemble averages
455 < for the system.
453 > Sec.~\ref{introSec:ergodic}, the average of static observables over
454 > the length of the simulation are taken to be the ensemble averages for
455 > the system.
456  
457   The choice of when to use molecular dynamics over Monte Carlo
458   techniques, is normally decided by the observables in which the
459 < researcher is interested.  If the observables depend on time in
460 < any fashion, then the only choice is molecular dynamics in some form.
459 > researcher is interested.  If the observables depend on time in any
460 > fashion, then the only choice is molecular dynamics in some form.
461   However, when the observable is dependent only on the configuration,
462 < then for most small systems, Monte Carlo techniques will be more efficient.
462 > then for most small systems, Monte Carlo techniques will be more
463 > efficient.  The focus of research in the second half of this
464 > dissertation is centered on the dynamic properties of phospholipid
465 > bilayers, making molecular dynamics key in the simulation of those
466 > properties.
467  
468 < The focus of research in the second half of this dissertation is
463 < centered around the dynamic properties of phospholipid bilayers,
464 < making molecular dynamics key in the simulation of those properties.
468 > \subsection{\label{introSec:mdAlgorithm}Molecular Dynamics Algorithms}
469  
466 \subsection{\label{introSec:mdAlgorithm}The Molecular Dynamics Algorithm}
467
470   To illustrate how the molecular dynamics technique is applied, the
471   following sections will describe the sequence involved in a
472   simulation.  Sec.~\ref{introSec:mdInit} deals with the initialization
# Line 480 | Line 482 | water, and in other cases structured the lipids into p
482   Ch.~\ref{chapt:lipid} deals with the formation and equilibrium dynamics of
483   phospholipid membranes.  Therefore in these simulations initial
484   positions were selected that in some cases dispersed the lipids in
485 < water, and in other cases structured the lipids into performed
485 > water, and in other cases structured the lipids into preformed
486   bilayers.  Important considerations at this stage of the simulation are:
487   \begin{itemize}
488 < \item There are no major overlaps of molecular or atomic orbitals
489 < \item Velocities are chosen in such a way as to not give the system a non=zero total momentum or angular momentum.
490 < \item It is also sometimes desirable to select the velocities to correctly sample the target temperature.
488 > \item There are no overlapping molecules or atoms.
489 > \item Velocities are chosen in such a way as to give the system zero total momentum and angular momentum.
490 > \item It is also sometimes desirable to select the velocities to correctly sample the ensemble at a particular target temperature.
491   \end{itemize}
492  
493 < The first point is important due to the amount of potential energy
494 < generated by having two particles too close together.  If overlap
495 < occurs, the first evaluation of forces will return numbers so large as
496 < to render the numerical integration of the motion meaningless.  The
497 < second consideration keeps the system from drifting or rotating as a
498 < whole.  This arises from the fact that most simulations are of systems
499 < in equilibrium in the absence of outside forces.  Therefore any net
493 > The first point is important due to the forces generated when two
494 > particles are too close together.  If overlap occurs, the first
495 > evaluation of forces will return numbers so large as to render the
496 > numerical integration of the equations motion meaningless.  The second
497 > consideration keeps the system from drifting or rotating as a whole.
498 > This arises from the fact that most simulations are of systems in
499 > equilibrium in the absence of outside forces.  Therefore any net
500   movement would be unphysical and an artifact of the simulation method
501   used.  The final point addresses the selection of the magnitude of the
502 < initial velocities.  For many simulations it is convenient to use
503 < this opportunity to scale the amount of kinetic energy to reflect the
502 > initial velocities.  For many simulations it is convenient to use this
503 > opportunity to scale the amount of kinetic energy to reflect the
504   desired thermal distribution of the system.  However, it must be noted
505 < that most systems will require further velocity rescaling after the
506 < first few initial simulation steps due to either loss or gain of
507 < kinetic energy from energy stored in potential degrees of freedom.
505 > that due to equipartition, most systems will require further velocity
506 > rescaling after the first few initial simulation steps due to either
507 > loss or gain of kinetic energy from energy stored as potential energy.
508  
509   \subsection{\label{introSec:mdForce}Force Evaluation}
510  
511   The evaluation of forces is the most computationally expensive portion
512 < of a given molecular dynamics simulation.  This is due entirely to the
513 < evaluation of long range forces in a simulation, typically pair-wise.
514 < These forces are most commonly the Van der Waals force, and sometimes
515 < Coulombic forces as well.  For a pair-wise force, there are $N(N-1)/ 2$
516 < pairs to be evaluated, where $N$ is the number of particles in the
517 < system.  This leads to the calculations scaling as $N^2$, making large
518 < simulations prohibitive in the absence of any computation saving
519 < techniques.
512 > of any molecular dynamics simulation.  This is due entirely to the
513 > evaluation of long range forces in a simulation, which are typically
514 > pair potentials.  These forces are most commonly the van der Waals
515 > force, and sometimes Coulombic forces as well.  For a pair-wise
516 > interaction, there are $N(N-1)/ 2$ pairs to be evaluated, where $N$ is
517 > the number of particles in the system.  This leads to calculations which
518 > scale as $N^2$, making large simulations prohibitive in the absence
519 > of any computation saving techniques.
520  
521 < Another consideration one must resolve, is that in a given simulation
521 > Another consideration one must resolve, is that in a given simulation,
522   a disproportionate number of the particles will feel the effects of
523   the surface.\cite{allen87:csl} For a cubic system of 1000 particles
524   arranged in a $10 \times 10 \times 10$ cube, 488 particles will be
525 < exposed to the surface.  Unless one is simulating an isolated particle
526 < group in a vacuum, the behavior of the system will be far from the
527 < desired bulk characteristics.  To offset this, simulations employ the
528 < use of periodic boundary images.\cite{born:1912}
525 > exposed to the surface.  Unless one is simulating an isolated cluster
526 > in a vacuum, the behavior of the system will be far from the desired
527 > bulk characteristics.  To offset this, simulations employ the use of
528 > periodic images.\cite{born:1912}
529  
530   The technique involves the use of an algorithm that replicates the
531 < simulation box on an infinite lattice in Cartesian space.  Any given
531 > simulation box on an infinite lattice in Cartesian space.  Any
532   particle leaving the simulation box on one side will have an image of
533 < itself enter on the opposite side (see Fig.~\ref{introFig:pbc}).  In
534 < addition, this sets that any two particles have an image, real or
535 < periodic, within $\text{box}/2$ of each other.  A discussion of the
536 < method used to calculate the periodic image can be found in
537 < Sec.\ref{oopseSec:pbc}.
533 > itself enter on the opposite side (see
534 > Fig.~\ref{introFig:pbc}). Therefore, a pair of particles have an
535 > image of each other, real or periodic, within $\text{box}/2$.  A
536 > discussion of the method used to calculate the periodic image can be
537 > found in Sec.\ref{oopseSec:pbc}.
538  
539   \begin{figure}
540   \centering
# Line 541 | Line 543 | Returning to the topic of the computational scale of t
543   \label{introFig:pbc}
544   \end{figure}
545  
546 < Returning to the topic of the computational scale of the force
546 > Returning to the topic of the computational scaling of the force
547   evaluation, the use of periodic boundary conditions requires that a
548   cutoff radius be employed.  Using a cutoff radius improves the
549   efficiency of the force evaluation, as particles farther than a
# Line 550 | Line 552 | one image that is seen by another atom, and further th
552   there are two methods to choose from, both with their own cutoff
553   limits. In the minimum image convention, $r_{\text{cut}}$ has a
554   maximum value of $\text{box}/2$. This is because each atom has only
555 < one image that is seen by another atom, and further the image used is
556 < the one that minimizes the distance between the two atoms. A system of
557 < wrapped images about a central atom therefore has a maximum length
558 < scale of box on a side (Fig.~\ref{introFig:rMaxMin}). The second
559 < convention, multiple image convention, has a maximum $r_{\text{cut}}$
560 < of box. Here multiple images of each atom are replicated in the
555 > one image that is seen by other atoms. The image used is the one that
556 > minimizes the distance between the two atoms. A system of wrapped
557 > images about a central atom therefore has a maximum length scale of
558 > box on a side (Fig.~\ref{introFig:rMaxMin}). The second convention,
559 > multiple image convention, has a maximum $r_{\text{cut}}$ of the box
560 > length itself. Here multiple images of each atom are replicated in the
561   periodic cells surrounding the central atom, this causes the atom to
562   see multiple copies of several atoms. If the cutoff radius is larger
563   than box, however, then the atom will see an image of itself, and
564   attempt to calculate an unphysical self-self force interaction
565   (Fig.~\ref{introFig:rMaxMult}). Due to the increased complexity and
566   commputaional ineffeciency of the multiple image method, the minimum
567 < image method is the periodic method used throughout this research.
567 > image method is the used throughout this research.
568  
569   \begin{figure}
570   \centering
# Line 574 | Line 576 | image method is the periodic method used throughout th
576   \begin{figure}
577   \centering
578   \includegraphics[width=\linewidth]{rCutMaxMultFig.eps}
579 < \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.}
579 > \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.}
580   \label{introFig:rMaxMult}
581   \end{figure}
582  
583 < With the use of an $r_{\text{cut}}$, however, comes a discontinuity in
583 > With the use of a cutoff radius, however, comes a discontinuity in
584   the potential energy curve (Fig.~\ref{introFig:shiftPot}). To fix this
585   discontinuity, one calculates the potential energy at the
586   $r_{\text{cut}}$, and adds that value to the potential, causing
# Line 594 | Line 596 | neighbor list. \cite{allen87:csl} In the Verlet method
596   \end{figure}
597  
598   The second main simplification used in this research is the Verlet
599 < neighbor list. \cite{allen87:csl} In the Verlet method, one generates
599 > neighbor list.\cite{allen87:csl} In the Verlet method, one generates
600   a list of all neighbor atoms, $j$, surrounding atom $i$ within some
601   cutoff $r_{\text{list}}$, where $r_{\text{list}}>r_{\text{cut}}$.
602   This list is created the first time forces are evaluated, then on
603   subsequent force evaluations, pair calculations are only calculated
604 < from the neighbor lists.  The lists are updated if any given particle
604 > from the neighbor lists.  The lists are updated if any particle
605   in the system moves farther than $r_{\text{list}}-r_{\text{cut}}$,
606 < giving rise to the possibility that a particle has left or joined a
606 > which indeicates the possibility that a particle has left or joined the
607   neighbor list.
608  
609 < \subsection{\label{introSec:mdIntegrate} Integration of the equations of motion}
609 > \subsection{\label{introSec:mdIntegrate} Integration of the Equations of Motion}
610  
611   A starting point for the discussion of molecular dynamics integrators
612   is the Verlet algorithm.\cite{Frenkel1996} It begins with a Taylor
613   expansion of position in time:
614   \begin{equation}
615   q(t+\Delta t)= q(t) + v(t)\Delta t + \frac{F(t)}{2m}\Delta t^2 +
616 <        \frac{\Delta t^3}{3!}\frac{\partial q(t)}{\partial t} +
616 >        \frac{\Delta t^3}{3!}\frac{\partial^3 q(t)}{\partial t^3} +
617          \mathcal{O}(\Delta t^4)
618   \label{introEq:verletForward}
619   \end{equation}
620   As well as,
621   \begin{equation}
622   q(t-\Delta t)= q(t) - v(t)\Delta t + \frac{F(t)}{2m}\Delta t^2 -
623 <        \frac{\Delta t^3}{3!}\frac{\partial q(t)}{\partial t} +
623 >        \frac{\Delta t^3}{3!}\frac{\partial^3 q(t)}{\partial t^3} +
624          \mathcal{O}(\Delta t^4)
625   \label{introEq:verletBack}
626   \end{equation}
# Line 641 | Line 643 | with a velocity reformulation of the Verlet method.\ci
643   order of $\Delta t^4$.
644  
645   In practice, however, the simulations in this research were integrated
646 < with a velocity reformulation of the Verlet method.\cite{allen87:csl}
646 > with the velocity reformulation of the Verlet method.\cite{allen87:csl}
647   \begin{align}
648   q(t+\Delta t) &= q(t) + v(t)\Delta t + \frac{F(t)}{2m}\Delta t^2 %
649   \label{introEq:MDvelVerletPos} \\%
# Line 652 | Line 654 | very little long term drift in energy conservation.  E
654   The original Verlet algorithm can be regained by substituting the
655   velocity back into Eq.~\ref{introEq:MDvelVerletPos}.  The Verlet
656   formulations are chosen in this research because the algorithms have
657 < very little long term drift in energy conservation.  Energy
658 < conservation in a molecular dynamics simulation is of extreme
659 < importance, as it is a measure of how closely one is following the
660 < ``true'' trajectory with the finite integration scheme.  An exact
661 < solution to the integration will conserve area in phase space, as well
657 > very little long time drift in energy.  Energy conservation in a
658 > molecular dynamics simulation is of extreme importance, as it is a
659 > measure of how closely one is following the ``true'' trajectory with
660 > the finite integration scheme.  An exact solution to the integration
661 > will conserve area in phase space (i.e.~will be symplectic), as well
662   as be reversible in time, that is, the trajectory integrated forward
663   or backwards will exactly match itself.  Having a finite algorithm
664   that both conserves area in phase space and is time reversible,
665 < therefore increases, but does not guarantee the ``correctness'' or the
666 < integrated trajectory.
665 > therefore increases the reliability, but does not guarantee the
666 > ``correctness'' of the integrated trajectory.
667  
668   It can be shown,\cite{Frenkel1996} that although the Verlet algorithm
669   does not rigorously preserve the actual Hamiltonian, it does preserve
# Line 677 | Line 679 | parameters are needed to describe the motions.  The si
679   \subsection{\label{introSec:MDfurther}Further Considerations}
680  
681   In the simulations presented in this research, a few additional
682 < parameters are needed to describe the motions.  The simulations
683 < involving water and phospholipids in Ch.~\ref{chapt:lipid} are
682 > parameters are needed to describe the motions.  In the simulations
683 > involving water and phospholipids in Ch.~\ref{chapt:lipid}, we are
684   required to integrate the equations of motions for dipoles on atoms.
685   This involves an additional three parameters be specified for each
686   dipole atom: $\phi$, $\theta$, and $\psi$.  These three angles are
# Line 686 | Line 688 | accumulated into a single $3 \times 3$ matrix, $\mathb
688   $z$-axis, and $\theta$ is a rotation about the new $x$-axis, and
689   $\psi$ is a final rotation about the new $z$-axis (see
690   Fig.~\ref{introFig:eulerAngles}).  This sequence of rotations can be
691 < accumulated into a single $3 \times 3$ matrix, $\mathbf{A}$,
691 > accumulated into a single $3 \times 3$ matrix, $\mathsf{A}$,
692   defined as follows:
693   \begin{equation}
694 < \mathbf{A} =
694 > \mathsf{A} =
695   \begin{bmatrix}
696          \cos\phi\cos\psi-\sin\phi\cos\theta\sin\psi &%
697          \sin\phi\cos\psi+\cos\phi\cos\theta\sin\psi &%
# Line 728 | Line 730 | Where $\omega^s_i$ is the angular velocity in the lab
730          \omega^s_y \frac{\cos\phi}{\sin\theta}
731   \label{introEq:MDeulerPsi}
732   \end{align}
733 < Where $\omega^s_i$ is the angular velocity in the lab space frame
734 < along Cartesian coordinate $i$.  However, a difficulty arises when
733 > Where $\omega^s_{\alpha}$ is the angular velocity in the lab space frame
734 > along Cartesian coordinate $\alpha$.  However, a difficulty arises when
735   attempting to integrate Eq.~\ref{introEq:MDeulerPhi} and
736   Eq.~\ref{introEq:MDeulerPsi}. The $\frac{1}{\sin \theta}$ present in
737   both equations means there is a non-physical instability present when
738   $\theta$ is 0 or $\pi$. To correct for this, the simulations integrate
739 < the rotation matrix, $\mathbf{A}$, directly, thus avoiding the
739 > the rotation matrix, $\mathsf{A}$, directly, thus avoiding the
740   instability.  This method was proposed by Dullweber
741   \emph{et. al.}\cite{Dullweber1997}, and is presented in
742   Sec.~\ref{introSec:MDsymplecticRot}.
# Line 743 | Line 745 | scheme.  It has been previously
745  
746   Before discussing the integration of the rotation matrix, it is
747   necessary to understand the construction of a ``good'' integration
748 < scheme.  It has been previously
749 < discussed(Sec.~\ref{introSec:mdIntegrate}) how it is desirable for an
750 < integrator to be symplectic, or time reversible.  The following is an
748 > scheme.  It has been previously discussed
749 > (Sec.~\ref{introSec:mdIntegrate}) how it is desirable for an
750 > integrator to be symplectic and time reversible.  The following is an
751   outline of the Trotter factorization of the Liouville Propagator as a
752 < scheme for generating symplectic integrators.\cite{Tuckerman92}
752 > scheme for generating symplectic, time-reversible
753 > integrators.\cite{Tuckerman92}
754  
755   For a system with $f$ degrees of freedom the Liouville operator can be
756   defined as,
# Line 797 | Line 800 | unitary.  Meaning an integrator based on this factoriz
800   \label{introEq:Lp5}
801   \end{align}
802   Because $U_1(t)$ and $U_2(t)$ are unitary, $G(\Delta t)$ is also
803 < unitary.  Meaning an integrator based on this factorization will be
803 > unitary.  This means that an integrator based on this factorization will be
804   reversible in time.
805  
806   As an example, consider the following decomposition of $L$:
# Line 860 | Line 863 | symplectic propagation of the rotation matrix, $\mathb
863  
864   Based on the factorization from the previous section,
865   Dullweber\emph{et al}.\cite{Dullweber1997}~ proposed a scheme for the
866 < symplectic propagation of the rotation matrix, $\mathbf{A}$, as an
866 > symplectic propagation of the rotation matrix, $\mathsf{A}$, as an
867   alternative method for the integration of orientational degrees of
868   freedom. The method starts with a straightforward splitting of the
869   Liouville operator:
870   \begin{align}
871   iL_{\text{pos}} &= \dot{q}\frac{\partial}{\partial q} +
872 <        \mathbf{\dot{A}}\frac{\partial}{\partial \mathbf{A}}
872 >        \mathsf{\dot{A}}\frac{\partial}{\partial \mathsf{A}}
873   \label{introEq:SR1a} \\%
874   %
875   iL_F &= F(q)\frac{\partial}{\partial p}
876   \label{introEq:SR1b} \\%
877 < iL_{\tau} &= \tau(\mathbf{A})\frac{\partial}{\partial \pi}
877 > iL_{\tau} &= \tau(\mathsf{A})\frac{\partial}{\partial j}
878   \label{introEq:SR1b} \\%
879   \end{align}
880 < Where $\tau(\mathbf{A})$ is the torque of the system
881 < due to the configuration, and $\pi$ is the conjugate
880 > Where $\tau(\mathsf{A})$ is the torque of the system
881 > due to the configuration, and $j$ is the conjugate
882   angular momenta of the system. The propagator, $G(\Delta t)$, becomes
883   \begin{equation}
884   G(\Delta t) = e^{\frac{\Delta t}{2} F(q)\frac{\partial}{\partial p}} \,
885 <        e^{\frac{\Delta t}{2} \tau(\mathbf{A})\frac{\partial}{\partial \pi}} \,
885 >        e^{\frac{\Delta t}{2} \tau(\mathsf{A})\frac{\partial}{\partial j}} \,
886          e^{\Delta t\,iL_{\text{pos}}} \,
887 <        e^{\frac{\Delta t}{2} \tau(\mathbf{A})\frac{\partial}{\partial \pi}} \,
887 >        e^{\frac{\Delta t}{2} \tau(\mathsf{A})\frac{\partial}{\partial j}} \,
888          e^{\frac{\Delta t}{2} F(q)\frac{\partial}{\partial p}}
889   \label{introEq:SR2}
890   \end{equation}
# Line 899 | Line 902 | Where $\mathcal{U}_j$ is a unitary rotation of $\mathb
902          \mathcal{U}_x \biggl(\frac{\Delta t}{2}\biggr)\,
903   \label{introEq:SR3}
904   \end{equation}
905 < Where $\mathcal{U}_j$ is a unitary rotation of $\mathbf{A}$ and
906 < $\pi$ about each axis $j$.  As all propagations are now
905 > Where $\mathcal{U}_{\alpha}$ is a unitary rotation of $\mathsf{A}$ and
906 > $j$ about each axis $\alpha$.  As all propagations are now
907   unitary and symplectic, the entire integration scheme is also
908   symplectic and time reversible.
909  
910   \section{\label{introSec:layout}Dissertation Layout}
911  
912 < This dissertation is divided as follows:Ch.~\ref{chapt:RSA}
912 > This dissertation is divided as follows: Ch.~\ref{chapt:RSA}
913   presents the random sequential adsorption simulations of related
914   pthalocyanines on a gold (111) surface. Ch.~\ref{chapt:oopse}
915 < is about the writing of the molecular dynamics simulation package
915 > is about our molecular dynamics simulation package
916   {\sc oopse}. Ch.~\ref{chapt:lipid} regards the simulations of
917   phospholipid bilayers using a mesoscale model. And lastly,
918   Ch.~\ref{chapt:conclusion} concludes this dissertation with a
# Line 921 | Line 924 | Dr. Lieberman, about possible explanations for the  pa
924   study in applying Statistical Mechanics simulation techniques in order
925   to obtain a simple model capable of explaining the results.  My
926   advisor, Dr. Gezelter, and I were approached by a colleague,
927 < Dr. Lieberman, about possible explanations for the  partial coverage of a
928 < gold surface by a particular compound of hers. We suggested it might
929 < be due to the statistical packing fraction of disks on a plane, and
930 < set about to simulate this system.  As the events in our model were
931 < not dynamic in nature, a Monte Carlo method was employed.  Here, if a
932 < molecule landed on the surface without overlapping another, then its
933 < landing was accepted.  However, if there was overlap, the landing we
934 < rejected and a new random landing location was chosen.  This defined
935 < our acceptance rules and allowed us to construct a Markov chain whose
936 < limiting distribution was the surface coverage in which we were
937 < interested.
927 > Dr. Lieberman, about possible explanations for the partial coverage of
928 > a gold surface by a particular compound synthesized in her group. We
929 > suggested it might be due to the statistical packing fraction of disks
930 > on a plane, and set about simulating this system.  As the events in
931 > our model were not dynamic in nature, a Monte Carlo method was
932 > employed.  Here, if a molecule landed on the surface without
933 > overlapping with another, then its landing was accepted.  However, if there
934 > was overlap, the landing we rejected and a new random landing location
935 > was chosen.  This defined our acceptance rules and allowed us to
936 > construct a Markov chain whose limiting distribution was the surface
937 > coverage in which we were interested.
938  
939   The following chapter, about the simulation package {\sc oopse},
940   describes in detail the large body of scientific code that had to be
941   written in order to study phospholipid bilayers.  Although there are
942   pre-existing molecular dynamic simulation packages available, none
943 < were capable of implementing the models we were developing.{\sc oopse}
944 < is a unique package capable of not only integrating the equations of
945 < motion in Cartesian space, but is also able to integrate the
946 < rotational motion of rigid bodies and dipoles.  Add to this the
947 < ability to perform calculations across parallel processors and a
948 < flexible script syntax for creating systems, and {\sc oopse} becomes a
949 < very powerful scientific instrument for the exploration of our model.
943 > were capable of implementing the models we were developing. {\sc
944 > oopse} is a unique package capable of not only integrating the
945 > equations of motion in Cartesian space, but is also able to integrate
946 > the rotational motion of rigid bodies and dipoles.  It also has the
947 > ability to perform calculations across distributed parallel processors
948 > and contains a flexible script syntax for creating systems.  {\sc
949 > oopse} has become a very powerful scientific instrument for the
950 > exploration of our model.
951  
952 < Bringing us to Ch.~\ref{chapt:lipid}. Using {\sc oopse}, I have been
952 > In Ch.~\ref{chapt:lipid}, utilizing {\sc oopse}, I have been
953   able to parameterize a mesoscale model for phospholipid simulations.
954   This model retains information about solvent ordering around the
955   bilayer, as well as information regarding the interaction of the
# Line 955 | Line 959 | Which leads into the last chapter, where I discuss fut
959   provide the foundation for future exploration of bilayer phase
960   behavior with this model.  
961  
962 < Which leads into the last chapter, where I discuss future directions
963 < for both{\sc oopse} and this mesoscale model.  Additionally, I will
964 < give a summary of results for this dissertation.
962 > In the last chapter, I discuss future directions
963 > for both {\sc oopse} and this mesoscale model.  Additionally, I will
964 > give a summary of the results found in this dissertation.
965  
966  

Diff Legend

Removed lines
+ Added lines
< Changed lines
> Changed lines