1 |
< |
\chapter{\label{chap:intro}INTRODUCTION AND BACKGROUND} |
1 |
> |
\chapter{\label{chap:intro}INTRODUCTION AND BACKGROUND} |
2 |
|
|
3 |
+ |
The following dissertation presents the primary aspects of the |
4 |
+ |
research I have performed and been involved with over the last several |
5 |
+ |
years. Rather than presenting the topics in a chronological fashion, |
6 |
+ |
they are arranged to form a series where the later topics apply and |
7 |
+ |
extend the findings of the former topics. This layout does lead to |
8 |
+ |
occasional situations where knowledge gleaned from earlier chapters |
9 |
+ |
(particularly chapter \ref{chap:electrostatics}) is not applied |
10 |
+ |
outright in the later work; however, I feel that this organization is |
11 |
+ |
more instructive and provides a more cohesive progression of research |
12 |
+ |
efforts. |
13 |
+ |
|
14 |
+ |
This introductory chapter gives a general overview of molecular |
15 |
+ |
simulations, with particular emphasis on considerations that need to |
16 |
+ |
be made in order to apply the technique properly. This leads quite |
17 |
+ |
naturally into chapter \ref{chap:electrostatics}, where we investigate |
18 |
+ |
correction techniques for proper handling of long-ranged electrostatic |
19 |
+ |
interactions in molecular simulations. In particular we develop and |
20 |
+ |
evaluate some new simple pairwise methods. These techniques make an |
21 |
+ |
appearance in the later chapters, as they are applied to specific |
22 |
+ |
systems of interest, showing how it they can improve the quality of |
23 |
+ |
various molecular simulations. |
24 |
+ |
|
25 |
+ |
In chapter \ref{chap:water}, we focus on simple water models, |
26 |
+ |
specifically the single-point soft sticky dipole (SSD) family of water |
27 |
+ |
models. These single-point models are more efficient than the common |
28 |
+ |
multi-point partial charge models and, in many cases, better capture |
29 |
+ |
the dynamic properties of water. We discuss improvements to these |
30 |
+ |
models in regards to long-range electrostatic corrections and show |
31 |
+ |
that these models can work well with the techniques discussed in |
32 |
+ |
chapter \ref{chap:electrostatics}. By investigating and improving |
33 |
+ |
simple water models, we are extending the limits of computational |
34 |
+ |
efficiency for systems that depend heavily on water calculations. |
35 |
+ |
|
36 |
+ |
In chapter \ref{chap:ice}, we study a unique polymorph of ice that we |
37 |
+ |
discovered while performing water simulations with the fast simple |
38 |
+ |
water models discussed in the previous chapter. This form of ice, |
39 |
+ |
which we call ``imaginary ice'' (Ice-$i$), has a low-density structure |
40 |
+ |
which is different from any known polymorph characterized in either |
41 |
+ |
experiment or other simulations. In this work, we perform a free |
42 |
+ |
energy analysis and see that this structure is in fact the |
43 |
+ |
thermodynamically preferred form of ice for both the single-point and |
44 |
+ |
commonly used multi-point water models under the chosen simulation |
45 |
+ |
conditions. We also consider electrostatic corrections, again |
46 |
+ |
including the techniques discussed in chapter |
47 |
+ |
\ref{chap:electrostatics}, to see how they influence the free energy |
48 |
+ |
results. This work, to some degree, addresses the appropriateness of |
49 |
+ |
using these simplistic water models outside of the conditions for |
50 |
+ |
which they were developed. |
51 |
+ |
|
52 |
+ |
Finally, in chapter \ref{chap:conclusion}, we summarize the work |
53 |
+ |
presented in the previous chapters and connect ideas together with |
54 |
+ |
some final comments. The supporting information follows in the |
55 |
+ |
appendix, and it gives a more detailed look at systems discussed in |
56 |
+ |
chapter \ref{chap:electrostatics}. |
57 |
+ |
|
58 |
+ |
\section{On Molecular Simulation} |
59 |
+ |
|
60 |
|
In order to advance our understanding of natural chemical and physical |
61 |
|
processes, researchers develop explanations for events observed |
62 |
|
experimentally. These hypotheses, supported by a body of corroborating |
63 |
|
observations, can develop into accepted theories for how these |
64 |
|
processes occur. This validation process, as well as testing the |
65 |
|
limits of these theories, is essential in developing a firm foundation |
66 |
< |
for our knowledge. Theories involving molecular scale systems cause |
67 |
< |
particular difficulties in this process because their size and often |
68 |
< |
fast motions make them difficult to observe experimentally. One useful |
69 |
< |
tool for addressing these difficulties is computer simulation. |
66 |
> |
for our knowledge. Developing and validating theories involving |
67 |
> |
molecular scale systems is particularly difficult because their size |
68 |
> |
and often fast motions make them difficult to observe |
69 |
> |
experimentally. One useful tool for addressing these difficulties is |
70 |
> |
computer simulation. |
71 |
|
|
72 |
|
In computer simulations, we can develop models from either the theory |
73 |
< |
or experimental knowledge and then test them in a controlled |
74 |
< |
environment. Work done in this manner allows us to further refine |
73 |
> |
or our experimental knowledge and then test them in controlled |
74 |
> |
environments. Work done in this manner allows us to further refine |
75 |
|
theories, more accurately represent what is happening in experimental |
76 |
|
observations, and even make predictions about what one will see in |
77 |
< |
experiments. Thus, computer simulations of molecular systems act as a |
78 |
< |
bridge between theory and experiment. |
77 |
> |
future experiments. Thus, computer simulations of molecular systems |
78 |
> |
act as a bridge between theory and experiment. |
79 |
|
|
80 |
|
Depending on the system of interest, there are a variety of different |
81 |
|
computational techniques that can used to test and gather information |
93 |
|
molecular dynamics can be used to investigate dynamical |
94 |
|
quantities. The research presented in the following chapters utilized |
95 |
|
molecular dynamics near exclusively, so we will present a general |
96 |
< |
introduction to molecular dynamics and not Monte Carlo. There are |
97 |
< |
several resources available for those desiring a more in-depth |
98 |
< |
presentation of either of these |
41 |
< |
techniques.\cite{Allen87,Frenkel02,Leach01} |
96 |
> |
introduction to molecular dynamics. There are several resources |
97 |
> |
available for those desiring a more in-depth presentation of either of |
98 |
> |
these techniques.\cite{Allen87,Frenkel02,Leach01} |
99 |
|
|
100 |
|
\section{\label{sec:MolecularDynamics}Molecular Dynamics} |
101 |
|
|
105 |
|
configuration to evolve in a manner that mimics real molecular |
106 |
|
systems. To do this, we start with clarifying what we know about a |
107 |
|
given configuration of particles at time $t_1$, basically that each |
108 |
< |
particle in the configuration has a position ($q$) and velocity |
109 |
< |
($\dot{q}$). We now want to know what the configuration will be at |
108 |
> |
particle in the configuration has a position ($\mathbf{q}$) and velocity |
109 |
> |
($\dot{\mathbf{q}}$). We now want to know what the configuration will be at |
110 |
|
time $t_2$. To find out, we need the classical equations of |
111 |
|
motion, and one useful formulation of them is the Lagrangian form. |
112 |
|
|
113 |
< |
The Lagrangian ($L$) is a function of the positions and velocites that |
113 |
> |
The Lagrangian ($L$) is a function of the positions and velocities that |
114 |
|
takes the form, |
115 |
|
\begin{equation} |
116 |
|
L = K - V, |
122 |
|
motion, to say that the variation of the integral of the Lagrangian |
123 |
|
over time is zero,\cite{Tolman38} |
124 |
|
\begin{equation} |
125 |
< |
\delta\int_{t_1}^{t_2}L(q,\dot{q})dt = 0. |
125 |
> |
\delta\int_{t_1}^{t_2}L(\mathbf{q},\dot{\mathbf{q}})dt = 0. |
126 |
|
\end{equation} |
127 |
< |
This can be written as a summation of integrals to give |
127 |
> |
The variation can be transferred to the variables that make up the |
128 |
> |
Lagrangian, |
129 |
|
\begin{equation} |
130 |
< |
\int_{t_1}^{t_2}\sum_{i=1}^{3N}\left(\frac{\partial L}{\partial q_i}\delta q_i |
131 |
< |
+ \frac{\partial L}{\partial \dot{q}_i}\delta \dot{q}_i\right)dt = 0. |
130 |
> |
\int_{t_1}^{t_2}\sum_{i=1}^{3N}\left( |
131 |
> |
\frac{\partial L}{\partial \mathbf{q}_i}\delta \mathbf{q}_i |
132 |
> |
+ \frac{\partial L}{\partial \dot{\mathbf{q}}_i}\delta |
133 |
> |
\dot{\mathbf{q}}_i\right)dt = 0. |
134 |
|
\end{equation} |
135 |
< |
Using fact that $\dot q$ is the derivative with respect to time of $q$ |
136 |
< |
and integrating the second partial derivative in the parenthesis by |
137 |
< |
parts, this equation simplifies to |
135 |
> |
Using the fact that $\dot{\mathbf{q}}$ is the derivative of |
136 |
> |
$\mathbf{q}$ with respect to time and integrating the second partial |
137 |
> |
derivative in the parenthesis by parts, this equation simplifies to |
138 |
|
\begin{equation} |
139 |
|
\int_{t_1}^{t_2}\sum_{i=1}^{3N}\left( |
140 |
< |
\frac{d}{dt}\frac{\partial L}{\partial \dot{q}_i} |
141 |
< |
- \frac{\partial L}{\partial q_i}\right)\delta {q}_i dt = 0, |
140 |
> |
\frac{d}{dt}\frac{\partial L}{\partial \dot{\mathbf{q}}_i} |
141 |
> |
- \frac{\partial L}{\partial \mathbf{q}_i}\right) |
142 |
> |
\delta {\mathbf{q}}_i dt = 0, |
143 |
|
\end{equation} |
144 |
< |
and the above equation is only valid for |
144 |
> |
and since each variable is independent, we can separate the |
145 |
> |
contribution from each of the variables: |
146 |
|
\begin{equation} |
147 |
< |
\frac{d}{dt}\frac{\partial L}{\partial \dot{q}_i} |
148 |
< |
- \frac{\partial L}{\partial q_i} = 0\quad\quad(i=1,2,\dots,3N). |
147 |
> |
\frac{d}{dt}\frac{\partial L}{\partial \dot{\mathbf{q}}_i} |
148 |
> |
- \frac{\partial L}{\partial \mathbf{q}_i} = 0 |
149 |
> |
\quad\quad(i=1,2,\dots,N). |
150 |
|
\label{eq:formulation} |
151 |
|
\end{equation} |
152 |
|
To obtain the classical equations of motion for the particles, we can |
179 |
|
\begin{equation} |
180 |
|
m\frac{d^2\mathbf{r}_i}{dt^2} = \sum_{j\ne i}\mathbf{f}(r_{ij}), |
181 |
|
\end{equation} |
182 |
< |
with the following simple algorithm: |
182 |
> |
with the following algorithm: |
183 |
|
\begin{equation} |
184 |
|
\mathbf{r}_i(t+\delta t) = -\mathbf{r}_i(t-\delta t) + 2\mathbf{r}_i(t) |
185 |
|
+ \sum_{j\ne i}\mathbf{f}(r_{ij}(t))\delta t^2, |
189 |
|
interesting to note that equation (\ref{eq:verlet}) does not include |
190 |
|
velocities, and this makes sense since they are not present in the |
191 |
|
differential equation. The velocities are necessary for the |
192 |
< |
calculation of the kinetic energy, and they can be calculated at each |
193 |
< |
time step with the following equation: |
192 |
> |
calculation of the kinetic energy and can be calculated at each time |
193 |
> |
step with the equation: |
194 |
|
\begin{equation} |
195 |
|
\mathbf{v}_i(t) = \frac{\mathbf{r}_i(t+\delta t)-\mathbf{r}_i(t-\delta t)} |
196 |
|
{2\delta t}. |
204 |
|
higher order information that is discarded after integrating |
205 |
|
steps. Another interesting property of this algorithm is that it is |
206 |
|
symplectic, meaning that it preserves area in phase-space. Symplectic |
207 |
< |
algorithms system stays in the region of phase-space dictated by the |
208 |
< |
ensemble and enjoy greater long-time energy |
209 |
< |
conservations.\cite{Frenkel02} |
207 |
> |
algorithms keep the system evolving in the region of phase-space |
208 |
> |
dictated by the ensemble and enjoy a greater degree of energy |
209 |
> |
conservation.\cite{Frenkel02} |
210 |
|
|
211 |
|
While the error in the positions calculated using the Verlet algorithm |
212 |
|
is small ($\mathcal{O}(\delta t^4)$), the error in the velocities is |
213 |
< |
quite a bit larger ($\mathcal{O}(\delta t^2)$).\cite{Allen87} Swope |
214 |
< |
{\it et al.} developed a corrected for of this algorithm, the |
213 |
> |
substantially larger ($\mathcal{O}(\delta t^2)$).\cite{Allen87} Swope |
214 |
> |
{\it et al.} developed a corrected version of this algorithm, the |
215 |
|
`velocity Verlet' algorithm, which improves the error of the velocity |
216 |
|
calculation and thus the energy conservation.\cite{Swope82} This |
217 |
|
algorithm involves a full step of the positions, |
224 |
|
\mathbf{v}\left(t+\frac{1}{2}\delta t\right) = \mathbf{v}(t) |
225 |
|
+ \frac{1}{2}\delta t\mathbf{a}(t). |
226 |
|
\end{equation} |
227 |
< |
After calculating new forces, the velocities can be updated to a full |
228 |
< |
step, |
227 |
> |
After forces are calculated at the new positions, the velocities can |
228 |
> |
be updated to a full step, |
229 |
|
\begin{equation} |
230 |
|
\mathbf{v}(t+\delta t) = \mathbf{v}\left(t+\frac{1}{2}\delta t\right) |
231 |
|
+ \frac{1}{2}\delta t\mathbf{a}(t+\delta t). |
240 |
|
\subsection{\label{sec:IntroIntegrate}Rigid Body Motion} |
241 |
|
|
242 |
|
Rigid bodies are non-spherical particles or collections of particles |
243 |
< |
(e.g. $\mbox{C}_{60}$) that have a constant internal potential and |
244 |
< |
move collectively.\cite{Goldstein01} Discounting iterative constraint |
245 |
< |
procedures like {\sc shake} and {\sc rattle} for approximating rigid |
246 |
< |
bodies, they are not included in most simulation packages because of |
247 |
< |
the algorithmic complexity involved in propagating orientational |
243 |
> |
that have a constant internal potential and move |
244 |
> |
collectively.\cite{Goldstein01} To move these particles, the |
245 |
> |
translational and rotational motion can be integrated |
246 |
> |
independently. Discounting iterative constraint procedures like {\sc |
247 |
> |
shake} and {\sc rattle} for approximating rigid body dynamics, rigid |
248 |
> |
bodies are not included in most simulation packages because of the |
249 |
> |
algorithmic complexity involved in propagating the orientational |
250 |
|
degrees of freedom.\cite{Ryckaert77,Andersen83,Krautler01} Integrators |
251 |
|
which propagate orientational motion with an acceptable level of |
252 |
|
energy conservation for molecular dynamics are relatively new |
270 |
|
where $\boldsymbol{\tau}_i$ and $\mathbf{r}_i$ are the torque on and |
271 |
|
position of the center of mass respectively, while $\mathbf{f}_{ia}$, |
272 |
|
$\mathbf{r}_{ia}$, and $\boldsymbol{\tau}_{ia}$ are the force on, |
273 |
< |
position of, and torque on the component particles. |
273 |
> |
position of, and torque on the component particles of the rigid body. |
274 |
|
|
275 |
|
The summation of the total torque is done in the body fixed axis. In |
276 |
|
order to move between the space fixed and body fixed coordinate axes, |
295 |
|
representation of the orientation of a rigid |
296 |
|
body.\cite{Evans77,Evans77b,Allen87} Thus, the elements of |
297 |
|
$\mathsf{A}$ can be expressed as arithmetic operations involving the |
298 |
< |
four quaternions ($q_w, q_x, q_y,$ and $q_z$), |
298 |
> |
four quaternions ($q_0, q_1, q_2,$ and $q_3$), |
299 |
|
\begin{equation} |
300 |
|
\mathsf{A} = \left( \begin{array}{l@{\quad}l@{\quad}l} |
301 |
|
q_0^2+q_1^2-q_2^2-q_3^2 & 2(q_1q_2+q_0q_3) & 2(q_1q_3-q_0q_2) \\ |
306 |
|
Integration of the equations of motion involves a series of arithmetic |
307 |
|
operations involving the quaternions and angular momenta and leads to |
308 |
|
performance enhancements over Euler angles, particularly for very |
309 |
< |
small systems.\cite{Evans77} This integration methods works well for |
309 |
> |
small systems.\cite{Evans77} This integration method works well for |
310 |
|
propagating orientational motion in the canonical ensemble ($NVT$); |
311 |
|
however, energy conservation concerns arise when using the simple |
312 |
|
quaternion technique under the microcanonical ($NVE$) ensemble. An |
313 |
< |
earlier implementation of {\sc oopse} utilized quaternions for |
313 |
> |
earlier implementation of our simulation code utilized quaternions for |
314 |
|
propagation of rotational motion; however, a detailed investigation |
315 |
|
showed that they resulted in a steady drift in the total energy, |
316 |
< |
something that has been observed by Kol {\it et al.} (also see |
317 |
< |
section~\ref{sec:waterSimMethods}).\cite{Kol97} |
316 |
> |
something that had also been observed by Kol {\it et al.} (see |
317 |
> |
section~\ref{sec:waterSimMethods} for information on this |
318 |
> |
investigation).\cite{Kol97} |
319 |
|
|
320 |
< |
Because of the outlined issues involving integration of the |
321 |
< |
orientational motion using both Euler angles and quaternions, we |
322 |
< |
decided to focus on a relatively new scheme that propagates the entire |
323 |
< |
nine parameter rotation matrix. This techniques is a velocity-Verlet |
324 |
< |
version of the symplectic splitting method proposed by Dullweber, |
325 |
< |
Leimkuhler and McLachlan ({\sc dlm}).\cite{Dullweber97} When there are |
326 |
< |
no directional atoms or rigid bodies present in the simulation, this |
327 |
< |
integrator becomes the standard velocity-Verlet integrator which is |
328 |
< |
known to effectively sample the microcanonical ($NVE$) |
320 |
> |
Because of these issues involving integration of the orientational |
321 |
> |
motion using both Euler angles and quaternions, we decided to focus on |
322 |
> |
a relatively new scheme that propagates the entire nine parameter |
323 |
> |
rotation matrix. This techniques is a velocity-Verlet version of the |
324 |
> |
symplectic splitting method proposed by Dullweber, Leimkuhler and |
325 |
> |
McLachlan ({\sc dlm}).\cite{Dullweber97} When there are no directional |
326 |
> |
atoms or rigid bodies present in the simulation, this {\sc dlm} |
327 |
> |
integrator reduces to the standard velocity-Verlet integrator, which |
328 |
> |
is known to effectively sample the constant energy $NVE$ |
329 |
|
ensemble.\cite{Frenkel02} |
330 |
|
|
331 |
|
The key aspect of the integration method proposed by Dullweber |
338 |
|
substantial benefits in energy conservation. |
339 |
|
|
340 |
|
The integration of the equations of motion is carried out in a |
341 |
< |
velocity-Verlet style two-part algorithm.\cite{Swope82} The first-part |
341 |
> |
velocity-Verlet style two-part algorithm.\cite{Swope82} The first part |
342 |
|
({\tt moveA}) consists of a half-step ($t + \delta t/2$) of the linear |
343 |
|
velocity (${\bf v}$) and angular momenta ({\bf j}) and a full-step ($t |
344 |
|
+ \delta t$) of the positions ({\bf r}) and rotation matrix, |
400 |
|
{\it symplectic}), |
401 |
|
\item the integrator is time-{\it reversible}, and |
402 |
|
\item the error for a single time step is of order |
403 |
< |
$\mathcal{O}\left(\delta t^4\right)$ for time steps of length $\delta t$. |
403 |
> |
$\mathcal{O}\left(\delta t^3\right)$ for time steps of length $\delta t$. |
404 |
|
\end{enumerate} |
405 |
|
|
406 |
|
After the initial half-step ({\tt moveA}), the forces and torques are |
487 |
|
0.001~kcal~mol$^{-1}$ per particle is desired, a nanosecond of |
488 |
|
simulation time will require ~19 hours of CPU time with the {\sc dlm} |
489 |
|
integrator, while the quaternion scheme will require ~154 hours of CPU |
490 |
< |
time. This demonstrates the computational advantage of the integration |
491 |
< |
scheme utilized in {\sc oopse}. |
490 |
> |
time. This demonstrates the computational advantage of the {\sc dlm} |
491 |
> |
integration scheme. |
492 |
|
|
493 |
|
\section{Accumulating Interactions} |
494 |
|
|
497 |
|
potential and forces (and torques if the particle is a rigid body or |
498 |
|
multipole) on each particle from their surroundings. This can quickly |
499 |
|
become a cumbersome task for large systems since the number of pair |
500 |
< |
interactions increases by $\mathcal{O}(N(N-1)/2)$ if you accumulate |
501 |
< |
interactions between all particles in the system, note the utilization |
502 |
< |
of Newton's third law to reduce the interaction number from |
503 |
< |
$\mathcal{O}(N^2)$. The case of periodic boundary conditions further |
504 |
< |
complicates matters by turning the finite system into an infinitely |
505 |
< |
repeating one. Fortunately, we can reduce the scale of this problem by |
506 |
< |
using spherical cutoff methods. |
500 |
> |
interactions increases by $\mathcal{O}(N(N-1)/2)$ when accumulating |
501 |
> |
interactions between all particles in the system. (Note the |
502 |
> |
utilization of Newton's third law to reduce the interaction number |
503 |
> |
from $\mathcal{O}(N^2)$.) Using periodic boundary conditions (section |
504 |
> |
\ref{sec:periodicBoundaries}) further complicates matters by turning |
505 |
> |
the finite system into an infinitely repeating one. Fortunately, we |
506 |
> |
can reduce the scale of this problem by using spherical cutoff |
507 |
> |
methods. |
508 |
|
|
509 |
|
\begin{figure} |
510 |
|
\centering |
517 |
|
\end{figure} |
518 |
|
With spherical cutoffs, rather than accumulating the full set of |
519 |
|
interactions between all particles in the simulation, we only |
520 |
< |
explicitly consider interactions between local particles out to a |
521 |
< |
specified cutoff radius distance, $R_\textrm{c}$, (see figure |
520 |
> |
explicitly consider interactions between particles separated by less |
521 |
> |
than a specified cutoff radius distance, $R_\textrm{c}$, (see figure |
522 |
|
\ref{fig:sphereCut}). This reduces the scaling of the interaction to |
523 |
|
$\mathcal{O}(N\cdot\textrm{c})$, where `c' is a value that depends on |
524 |
|
the size of $R_\textrm{c}$ (c $\approx R_\textrm{c}^3$). Determination |
525 |
|
of which particles are within the cutoff is also an issue, because |
526 |
|
this process requires a full loop over all $N(N-1)/2$ pairs. To reduce |
527 |
|
the this expense, we can use neighbor lists.\cite{Verlet67,Thompson83} |
528 |
< |
With neighbor lists, we have a second list of particles built from a |
529 |
< |
list radius $R_\textrm{l}$, which is larger than $R_\textrm{c}$. Once |
530 |
< |
any of the particles within $R_\textrm{l}$ move a distance of |
531 |
< |
$R_\textrm{l}-R_\textrm{c}$ (the ``skin'' thickness), we rebuild the |
532 |
< |
list with the full $N(N-1)/2$ loop.\cite{Verlet67} With an appropriate |
533 |
< |
skin thickness, these updates are only performed every $\sim$20 |
534 |
< |
time steps, significantly reducing the time spent on pair-list |
535 |
< |
bookkeeping operations.\cite{Allen87} If these neighbor lists are |
536 |
< |
utilized, it is important that these list updates occur |
528 |
> |
|
529 |
> |
When using neighbor lists, we build a second list of particles built |
530 |
> |
from a list radius $R_\textrm{l}$, which is larger than |
531 |
> |
$R_\textrm{c}$. Once any particle within $R_\textrm{l}$ moves half the |
532 |
> |
distance of $R_\textrm{l}-R_\textrm{c}$ (the ``skin'' thickness), we |
533 |
> |
rebuild the list with the full $N(N-1)/2$ loop.\cite{Verlet67} With an |
534 |
> |
appropriate skin thickness, these updates are only performed every |
535 |
> |
$\sim$20 time steps, significantly reducing the time spent on |
536 |
> |
pair-list bookkeeping operations.\cite{Allen87} If these neighbor |
537 |
> |
lists are utilized, it is important that these list updates occur |
538 |
|
regularly. Incorrect application of this technique leads to |
539 |
|
non-physical dynamics, such as the ``flying block of ice'' behavior |
540 |
|
for which improper neighbor list handling was identified a one of the |
553 |
|
region in order to smooth out the discontinuity.} |
554 |
|
\label{fig:ljCutoff} |
555 |
|
\end{figure} |
556 |
< |
As particle move in and out of $R_\textrm{c}$, there will be sudden |
557 |
< |
discontinuous jumps in the potential (and forces) due to their |
558 |
< |
appearance and disappearance. In order to prevent heating and poor |
559 |
< |
energy conservation in the simulations, this discontinuity needs to be |
560 |
< |
smoothed out. There are several ways to modify the function so that it |
561 |
< |
crosses $R_\textrm{c}$ in a continuous fashion, and the easiest |
562 |
< |
methods is shifting the potential. To shift the potential, we simply |
563 |
< |
subtract out the value we calculate at $R_\textrm{c}$ from the whole |
564 |
< |
potential. For the shifted form of the Lennard-Jones potential is |
556 |
> |
As the distance between a pair of particles fluctuates around |
557 |
> |
$R_\textrm{c}$, there will be sudden discontinuous jumps in the |
558 |
> |
potential (and forces) due to their inclusion and exclusion from the |
559 |
> |
interaction loop. In order to prevent heating and poor energy |
560 |
> |
conservation in the simulations, this discontinuity needs to be |
561 |
> |
smoothed out. There are several ways to modify the potential function |
562 |
> |
to eliminate these discontinuities, and the easiest method is shifting |
563 |
> |
the potential. To shift the potential, we simply subtract out the |
564 |
> |
value of the function at $R_\textrm{c}$ from the whole potential. The |
565 |
> |
shifted form of the Lennard-Jones potential is |
566 |
|
\begin{equation} |
567 |
|
V_\textrm{SLJ} = \left\{\begin{array}{l@{\quad\quad}l} |
568 |
|
V_\textrm{LJ}(r_{ij}) - V_\textrm{LJ}(R_\textrm{c}) & r_{ij} < R_\textrm{c}, \\ |
571 |
|
\end{equation} |
572 |
|
where |
573 |
|
\begin{equation} |
574 |
< |
V_\textrm{LJ} = 4\epsilon\left[\left(\frac{\sigma}{r_{ij}}\right)^{12} - |
575 |
< |
\left(\frac{\sigma}{r_{ij}}\right)^6\right]. |
574 |
> |
V_\textrm{LJ}(r_{ij}) = |
575 |
> |
4\epsilon\left[\left(\frac{\sigma}{r_{ij}}\right)^{12} - |
576 |
> |
\left(\frac{\sigma}{r_{ij}}\right)^6\right]. |
577 |
|
\end{equation} |
578 |
|
In figure \ref{fig:ljCutoff}, the shifted form of the potential |
579 |
|
reaches zero at the cutoff radius at the expense of the correct |
580 |
< |
magnitude below $R_\textrm{c}$. This correction method also does |
580 |
> |
magnitude inside $R_\textrm{c}$. This correction method also does |
581 |
|
nothing to correct the discontinuity in the forces. We can account for |
582 |
< |
this force discontinuity by shifting in the by applying the shifting |
583 |
< |
in the forces rather than just the potential via |
582 |
> |
this force discontinuity by applying the shifting in the forces as |
583 |
> |
well as in the potential via |
584 |
|
\begin{equation} |
585 |
|
V_\textrm{SFLJ} = \left\{\begin{array}{l@{\quad\quad}l} |
586 |
|
V_\textrm{LJ}({r_{ij}}) - V_\textrm{LJ}(R_\textrm{c}) - |
590 |
|
\end{array}\right. |
591 |
|
\end{equation} |
592 |
|
The forces are continuous with this potential; however, the inclusion |
593 |
< |
of the derivative term distorts the potential even further than the |
594 |
< |
simple shifting as shown in figure \ref{fig:ljCutoff}. The method for |
595 |
< |
correcting the discontinuity which results in the smallest |
596 |
< |
perturbation in both the potential and the forces is the use of a |
597 |
< |
switching function. The cubic switching function, |
593 |
> |
of the derivative term distorts the potential even further than |
594 |
> |
shifting only the potential as shown in figure \ref{fig:ljCutoff}. |
595 |
> |
|
596 |
> |
The method for correcting these discontinuities which results in the |
597 |
> |
smallest perturbation in both the potential and the forces is the use |
598 |
> |
of a switching function. The cubic switching function, |
599 |
|
\begin{equation} |
600 |
|
S(r) = \left\{\begin{array}{l@{\quad\quad}l} |
601 |
|
1 & r_{ij} \leqslant R_\textrm{sw}, \\ |
613 |
|
that the higher the order of the polynomial, the more abrupt the |
614 |
|
switching transition. |
615 |
|
|
616 |
< |
\subsection{Long-Range Interaction Corrections} |
616 |
> |
\subsection{\label{sec:LJCorrections}Long-Range Interaction Corrections} |
617 |
|
|
618 |
< |
While a good approximation, accumulating interaction only from the |
619 |
< |
nearby particles can lead to some issues, because the further away |
620 |
< |
surrounding particles do still have a small effect. For instance, |
618 |
> |
While a good approximation, accumulating interactions only from nearby |
619 |
> |
particles can lead to some issues, because particles at distances |
620 |
> |
greater than $R_\textrm{c}$ still have a small effect. For instance, |
621 |
|
while the strength of the Lennard-Jones interaction is quite weak at |
622 |
|
$R_\textrm{c}$ in figure \ref{fig:ljCutoff}, we are discarding all of |
623 |
< |
the attractive interaction that extends out to extremely long |
624 |
< |
distances. Thus, the potential is a little too high and the pressure |
625 |
< |
on the central particle from the surroundings is a little too low. For |
626 |
< |
homogeneous Lennard-Jones systems, we can correct for this neglect by |
627 |
< |
assuming a uniform density and integrating the missing part, |
623 |
> |
the attractive interactions that extend out to extremely long |
624 |
> |
distances. Thus, the potential will be a little too high and the |
625 |
> |
pressure on the central particle from the surroundings will be a |
626 |
> |
little too low. For homogeneous Lennard-Jones systems, we can correct |
627 |
> |
for this effect by assuming a uniform density ($\rho$) and integrating |
628 |
> |
the missing part, |
629 |
|
\begin{equation} |
630 |
|
V_\textrm{full}(r_{ij}) \approx V_\textrm{c} |
631 |
|
+ 2\pi N\rho\int^\infty_{R_\textrm{c}}r^2V_\textrm{LJ}(r)dr, |
632 |
|
\end{equation} |
633 |
|
where $V_\textrm{c}$ is the truncated Lennard-Jones |
634 |
< |
potential.\cite{Allen87} Like the potential, other Lennard-Jones |
635 |
< |
properties can be corrected by integration over the relevant |
636 |
< |
function. Note that with heterogeneous systems, this correction begins |
637 |
< |
to break down because the density is no longer uniform. |
634 |
> |
potential.\cite{Allen87} Like the potential, other properties can be |
635 |
> |
corrected by integration over the relevant function. Note that with |
636 |
> |
heterogeneous systems, this correction breaks down because the density |
637 |
> |
is no longer uniform. |
638 |
|
|
639 |
|
Correcting long-range electrostatic interactions is a topic of great |
640 |
|
importance in the field of molecular simulations. There have been |
641 |
|
several techniques developed to address this issue, like the Ewald |
642 |
|
summation and the reaction field technique. An in-depth analysis of |
643 |
< |
this problem, as well as useful corrective techniques, is presented in |
643 |
> |
this problem, as well as useful correction techniques, is presented in |
644 |
|
chapter \ref{chap:electrostatics}. |
645 |
|
|
646 |
< |
\subsection{Periodic Boundary Conditions} |
646 |
> |
\subsection{\label{sec:periodicBoundaries}Periodic Boundary Conditions} |
647 |
|
|
648 |
|
In typical molecular dynamics simulations there are no restrictions |
649 |
< |
placed on the motion of particles outside of what the interparticle |
649 |
> |
placed on the motion of particles outside of what the inter-particle |
650 |
|
interactions dictate. This means that if a particle collides with |
651 |
|
other particles, it is free to move away from the site of the |
652 |
|
collision. If we consider the entire system as a collection of |
679 |
|
region in space. To avoid this ``soft'' confinement, it is common |
680 |
|
practice to allow the particle coordinates to expand in an |
681 |
|
unrestricted fashion while calculating interactions using a wrapped |
682 |
< |
set of ``minimum image'' coordinates. These minimum image coordinates |
683 |
< |
need not be stored because they are easily calculated on the fly when |
684 |
< |
determining particle distances. |
682 |
> |
set of ``minimum image'' coordinates. These coordinates need not be |
683 |
> |
stored because they are easily calculated while determining particle |
684 |
> |
distances. |
685 |
|
|
686 |
|
\section{Calculating Properties} |
687 |
|
|
688 |
< |
In order to use simulations to model experimental process and evaluate |
689 |
< |
theories, we need to be able to extract properties from the |
688 |
> |
In order to use simulations to model experimental processes and |
689 |
> |
evaluate theories, we need to be able to extract properties from the |
690 |
|
results. In experiments, we can measure thermodynamic properties such |
691 |
|
as the pressure and free energy. In computer simulations, we can |
692 |
|
calculate properties from the motion and configuration of particles in |
695 |
|
|
696 |
|
The work presented in the later chapters use the canonical ($NVT$), |
697 |
|
isobaric-isothermal ($NPT$), and microcanonical ($NVE$) statistical |
698 |
< |
mechanics ensembles. The different ensembles lend themselves to more |
698 |
> |
mechanical ensembles. The different ensembles lend themselves to more |
699 |
|
effectively calculating specific properties. For instance, if we |
700 |
|
concerned ourselves with the calculation of dynamic properties, which |
701 |
|
are dependent upon the motion of the particles, it is better to choose |
702 |
< |
an ensemble that does not add system motions to keep properties like |
703 |
< |
the temperature or pressure constant. In this case, the $NVE$ ensemble |
704 |
< |
would be the most appropriate choice. In chapter \ref{chap:ice}, we |
705 |
< |
discuss calculating free energies, which are non-mechanical |
706 |
< |
thermodynamic properties, and these calculations also depend on the |
707 |
< |
chosen ensemble.\cite{Allen87} The Helmholtz free energy ($A$) depends |
708 |
< |
on the $NVT$ partition function ($Q_{NVT}$), |
702 |
> |
an ensemble that does not add artificial motions to keep properties |
703 |
> |
like the temperature or pressure constant. In this case, the $NVE$ |
704 |
> |
ensemble would be the most appropriate choice. In chapter |
705 |
> |
\ref{chap:ice}, we discuss calculating free energies, which are |
706 |
> |
non-mechanical thermodynamic properties, and these calculations also |
707 |
> |
depend on the chosen ensemble.\cite{Allen87} The Helmholtz free energy |
708 |
> |
($A$) depends on the $NVT$ partition function ($Q_{NVT}$), |
709 |
|
\begin{equation} |
710 |
|
A = -k_\textrm{B}T\ln Q_{NVT}, |
711 |
|
\end{equation} |
724 |
|
chosen property like the density in the $NPT$ ensemble, where the |
725 |
|
volume is allowed to fluctuate. The density for the simulation is |
726 |
|
calculated from an average over the densities for the individual |
727 |
< |
configurations. Being that the configurations from the integration |
728 |
< |
process are related to one another by the time evolution of the |
729 |
< |
interactions, this average is technically a time average. In |
730 |
< |
calculating thermodynamic properties, we would actually prefer an |
731 |
< |
ensemble average that is representative of all accessible states of |
732 |
< |
the system. We can calculate thermodynamic properties from the time |
733 |
< |
average by taking advantage of the ergodic hypothesis, which states |
734 |
< |
that over a long period of time, the time average and the ensemble |
735 |
< |
average are the same. |
727 |
> |
configurations. Since the configurations from the integration process |
728 |
> |
are related to one another by the time evolution of the interactions, |
729 |
> |
this average is technically a time average. In calculating |
730 |
> |
thermodynamic properties, we would actually prefer an ensemble average |
731 |
> |
that is representative of all accessible states of the system. We can |
732 |
> |
calculate thermodynamic properties from the time average by taking |
733 |
> |
advantage of the ergodic hypothesis, which states that for a |
734 |
> |
sufficiently chaotic system, and over a long enough period of time, |
735 |
> |
the time and ensemble averages are the same. |
736 |
|
|
737 |
< |
In addition to the average, the fluctuations of that particular |
738 |
< |
property can be determined via the standard deviation. Fluctuations |
739 |
< |
are useful for measuring various thermodynamic properties in computer |
737 |
> |
In addition to the average, the fluctuations of a particular property |
738 |
> |
can be determined via the standard deviation. Not only are |
739 |
> |
fluctuations useful for determining the spread of values around the |
740 |
> |
average and the error in the calculation of the value, they are also |
741 |
> |
useful for measuring various thermodynamic properties in computer |
742 |
|
simulations. In section \ref{sec:t5peThermo}, we use fluctuations in |
743 |
< |
propeties like the enthalpy and volume to calculate various |
743 |
> |
properties like the enthalpy and volume to calculate other |
744 |
|
thermodynamic properties, such as the constant pressure heat capacity |
745 |
|
and the isothermal compressibility. |
746 |
|
|
747 |
< |
\section{Application of the Techniques} |
747 |
> |
\section{OOPSE} |
748 |
|
|
749 |
|
In the following chapters, the above techniques are all utilized in |
750 |
|
the study of molecular systems. There are a number of excellent |
752 |
|
incorporate many of these |
753 |
|
methods.\cite{Brooks83,MacKerell98,Pearlman95,Berendsen95,Lindahl01,Smith96,Ponder87} |
754 |
|
Because of our interest in rigid body dynamics, point multipoles, and |
755 |
< |
systems where the orientational degrees cannot be handled using the |
756 |
< |
common constraint procedures (like {\sc shake}), the majority of the |
757 |
< |
following work was performed using {\sc oopse}, the object-oriented |
758 |
< |
parallel simulation engine.\cite{Meineke05} The {\sc oopse} package |
759 |
< |
started out as a collection of separate programs written within our |
760 |
< |
group, and has developed into one of the few parallel molecular |
761 |
< |
dynamics packages capable of accurately integrating rigid bodies and |
762 |
< |
point multipoles. |
755 |
> |
systems where the orientational degrees of freedom cannot be handled |
756 |
> |
using the common constraint procedures (like {\sc shake}), the |
757 |
> |
majority of the following work was performed using {\sc oopse}, the |
758 |
> |
object-oriented parallel simulation engine.\cite{Meineke05} The {\sc |
759 |
> |
oopse} package started out as a collection of separate programs |
760 |
> |
written within our group, and has developed into one of the few |
761 |
> |
parallel molecular dynamics packages capable of accurately integrating |
762 |
> |
rigid bodies and point multipoles. This simulation package is |
763 |
> |
open-source software, available to anyone interested in performing |
764 |
> |
molecular dynamics simulations. More information about {\sc oopse} can |
765 |
> |
be found in reference \cite{Meineke05} or at the {\tt |
766 |
> |
http://oopse.org} website. |
767 |
|
|
690 |
– |
In chapter \ref{chap:electrostatics}, we investigate correction |
691 |
– |
techniques for proper handling of long-ranged electrostatic |
692 |
– |
interactions. In particular we develop and evaluate some new pairwise |
693 |
– |
methods which we have incorporated into {\sc oopse}. These techniques |
694 |
– |
make an appearance in the later chapters, as they are applied to |
695 |
– |
specific systems of interest, showing how it they can improve the |
696 |
– |
quality of various molecular simulations. |
768 |
|
|
698 |
– |
In chapter \ref{chap:water}, we focus on simple water models, |
699 |
– |
specifically the single-point soft sticky dipole (SSD) family of water |
700 |
– |
models. These single-point models are more efficient than the common |
701 |
– |
multi-point partial charge models and, in many cases, better capture |
702 |
– |
the dynamic properties of water. We discuss improvements to these |
703 |
– |
models in regards to long-range electrostatic corrections and show |
704 |
– |
that these models can work well with the techniques discussed in |
705 |
– |
chapter \ref{chap:electrostatics}. By investigating and improving |
706 |
– |
simple water models, we are extending the limits of computational |
707 |
– |
efficiency for systems that heavily depend on water calculations. |
708 |
– |
|
709 |
– |
In chapter \ref{chap:ice}, we study a unique polymorph of ice that we |
710 |
– |
discovered while performing water simulations with the fast simple |
711 |
– |
water models discussed in the previous chapter. This form of ice, |
712 |
– |
which we called ``imaginary ice'' (Ice-$i$), has a low-density |
713 |
– |
structure which is different from any known polymorph from either |
714 |
– |
experiment or other simulations. In this study, we perform a free |
715 |
– |
energy analysis and see that this structure is in fact the |
716 |
– |
thermodynamically preferred form of ice for both the single-point and |
717 |
– |
commonly used multi-point water models under the chosen simulation |
718 |
– |
conditions. We also consider electrostatic corrections, again |
719 |
– |
including the techniques discussed in chapter |
720 |
– |
\ref{chap:electrostatics}, to see how they influence the free energy |
721 |
– |
results. This work, to some degree, addresses the appropriateness of |
722 |
– |
using these simplistic water models outside of the conditions for |
723 |
– |
which they were developed. |
724 |
– |
|
725 |
– |
In the final chapter we summarize the work presented previously. We |
726 |
– |
also close with some final comments before the supporting information |
727 |
– |
presented in the appendices. |