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

Comparing trunk/chrisDissertation/Introduction.tex (file contents):
Revision 2987 by chrisfen, Wed Aug 30 22:36:06 2006 UTC vs.
Revision 3025 by chrisfen, Tue Sep 26 16:02:25 2006 UTC

# Line 1 | Line 1
1 < \chapter{\label{chap:intro}INTRODUCTION AND BACKGROUND}
1 > \chapter{\label{chap:intro}INTRODUCTION AND BACKGROUND}
2  
3 < \section{\label{sec:IntroIntegrate}Rigid Body Motion}
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. 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 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 > 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
82 > from the developed models. In the study of classical systems, the two
83 > most commonly used techniques are Monte Carlo and molecular
84 > dynamics. Both of these methods operate by calculating interactions
85 > between particles of our model systems; however, the progression of
86 > the simulation under the different techniques is vastly
87 > different. Monte Carlo operates through random configuration changes
88 > that that follow rules adhering to a specific statistical mechanics
89 > ensemble, while molecular dynamics is chiefly concerned with solving
90 > the classic equations of motion to move between configurations within
91 > an ensemble. Thermodynamic properties can be calculated from both
92 > techniques; but because of the random nature of Monte Carlo, only
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. 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 >
102 > As stated above, in molecular dynamics we focus on evolving
103 > configurations of molecules over time. In order to use this as a tool
104 > for understanding experiments and testing theories, we want the
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 ($\mathbf{q}$) 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 velocities that
114 > takes the form,
115 > \begin{equation}
116 > L = K - V,
117 > \label{eq:lagrangian}
118 > \end{equation}
119 > where $K$ is the kinetic energy and $V$ is the potential energy. We
120 > can use Hamilton's principle, which states that the integral of the
121 > Lagrangian over time has a stationary value for the correct path of
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(\mathbf{q},\dot{\mathbf{q}})dt = 0.
126 > \end{equation}
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(
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 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{\mathbf{q}}_i}
141 >        - \frac{\partial L}{\partial \mathbf{q}_i}\right)
142 >                \delta {\mathbf{q}}_i dt = 0,
143 > \end{equation}
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{\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
153 > substitute equation (\ref{eq:lagrangian}) into the above equation with
154 > $m\dot{\mathbf{r}}^2/2$ for the kinetic energy, giving
155 > \begin{equation}
156 > \frac{d}{dt}(m\dot{\mathbf{r}})+\frac{dV}{d\mathbf{r}}=0,
157 > \end{equation}
158 > or more recognizably,
159 > \begin{equation}
160 > \mathbf{f} = m\mathbf{a},
161 > \end{equation}
162 > where $\mathbf{f} = -dV/d\mathbf{r}$ and $\mathbf{a} =
163 > d^2\mathbf{r}/dt^2$. This Lagrangian formulation shown in equation
164 > (\ref{eq:formulation}) is generalized, and it can be used to determine
165 > equations of motion in forms outside of the typical Cartesian case
166 > shown here.
167 >
168 > \subsection{\label{sec:Verlet}Verlet Integration}
169 >
170 > In order to perform molecular dynamics, we need an algorithm that
171 > integrates the equations of motion described above. Ideal algorithms
172 > are both simple in implementation and highly accurate. There have been
173 > a large number of algorithms developed for this purpose; however, for
174 > reasons discussed below, we are going to focus on the Verlet class of
175 > integrators.\cite{Gear66,Beeman76,Berendsen86,Allen87,Verlet67,Swope82}
176 >
177 > In Verlet's original study of computer ``experiments'', he directly
178 > integrated the Newtonian second order differential equation of motion,
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 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,
186 > \label{eq:verlet}
187 > \end{equation}
188 > where $\delta t$ is the time step of integration.\cite{Verlet67} It is
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 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}.
197 > \end{equation}
198 >
199 > Like the equation of motion it solves, the Verlet algorithm has the
200 > beneficial property of being time-reversible, meaning that you can
201 > integrate the configuration forward and then backward and end up at
202 > the original configuration. Some other methods for integration, like
203 > predictor-corrector methods, lack this property in that they require
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 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 > 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,
218 > \begin{equation}
219 > \mathbf{r}(t+\delta t) = \mathbf{r}(t) + \delta t\mathbf{v}(t)
220 >                                + \frac{1}{2}\delta t^2\mathbf{a}(t),
221 > \end{equation}
222 > and a half step of the velocities,
223 > \begin{equation}
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 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).
232 > \end{equation}
233 > By integrating in this manner, the error in the velocities reduces to
234 > $\mathcal{O}(\delta t^3)$. It should be noted that the error in the
235 > positions increases to $\mathcal{O}(\delta t^3)$, but the resulting
236 > improvement in the energies coupled with the maintained simplicity,
237 > time-reversibility, and symplectic nature make it an improvement over
238 > the original form.
239 >
240 > \subsection{\label{sec:IntroIntegrate}Rigid Body Motion}
241 >
242 > Rigid bodies are non-spherical particles or collections of particles
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
253 > inventions.
254 >
255 > Moving a rigid body involves determination of both the force and
256 > torque applied by the surroundings, which directly affect the
257 > translational and rotational motion in turn. In order to accumulate
258 > the total force on a rigid body, the external forces and torques must
259 > first be calculated for all the internal particles. The total force on
260 > the rigid body is simply the sum of these external forces.
261 > Accumulation of the total torque on the rigid body is more complex
262 > than the force because the torque is applied to the center of mass of
263 > the rigid body. The space-fixed torque on rigid body $i$ is
264 > \begin{equation}
265 > \boldsymbol{\tau}_i=
266 >        \sum_{a}\biggl[(\mathbf{r}_{ia}-\mathbf{r}_i)\times \mathbf{f}_{ia}
267 >        + \boldsymbol{\tau}_{ia}\biggr],
268 > \label{eq:torqueAccumulate}
269 > \end{equation}
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 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,
277 > parameters describing the orientation must be maintained for each
278 > rigid body. At a minimum, the rotation matrix ($\mathsf{A}$) can be
279 > described by the three Euler angles ($\phi, \theta,$ and $\psi$),
280 > where the elements of $\mathsf{A}$ are composed of trigonometric
281 > operations involving $\phi, \theta,$ and $\psi$.\cite{Goldstein01}
282 > Direct propagation of the Euler angles has a known $1/\sin\theta$
283 > divergence in the equations of motion for $\phi$ and $\psi$, leading
284 > to numerical instabilities any time one of the directional atoms or
285 > rigid bodies has an orientation near $\theta=0$ or
286 > $\theta=\pi$.\cite{Allen87} One of the most practical ways to avoid
287 > this ``gimbal point'' is to switch to another angular set defining the
288 > orientation of the rigid body near this point.\cite{Barojas73} This
289 > procedure results in additional book-keeping and increased algorithm
290 > complexity. In the search for more elegant alternative methods, Evans
291 > proposed the use of quaternions to describe and propagate
292 > orientational motion.\cite{Evans77}
293 >
294 > The quaternion method for integration involves a four dimensional
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 ($q_0, q_1, q_2,$ 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) \\
302 > 2(q_1q_2-q_0q_3) & q_0^2-q_1^2+q_2^2-q_3^2 & 2(q_2q_3+q_0q_1) \\
303 > 2(q_1q_3+q_0q_2) & 2(q_2q_3-q_0q_1) & q_0^2-q_1^2-q_2^2+q_3^2 \\
304 > \end{array}\right).
305 > \end{equation}
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 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 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 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 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
332 > \emph{et al.} is that the entire $3 \times 3$ rotation matrix is
333 > propagated from one time step to the next. In the past, this would not
334 > have been as feasible, since the rotation matrix for a single body has
335 > nine elements compared with the more memory-efficient methods (using
336 > three Euler angles or four quaternions).  Computer memory has become
337 > much less costly in recent years, and this can be translated into
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
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,
345 > \begin{equation*}
346 > {\tt moveA} = \left\{\begin{array}{r@{\quad\leftarrow\quad}l}
347 > {\bf v}\left(t + \delta t / 2\right) & {\bf v}(t)
348 >        + \left( {\bf f}(t) / m \right)(\delta t/2), \\
349 > %
350 > {\bf r}(t + \delta t) & {\bf r}(t)
351 >        + {\bf v}\left(t + \delta t / 2 \right)\delta t, \\
352 > %
353 > {\bf j}\left(t + \delta t / 2 \right) & {\bf j}(t)
354 >        + \boldsymbol{\tau}^b(t)(\delta t/2), \\
355 > %
356 > \mathsf{A}(t + \delta t) & \mathrm{rotate}\left( {\bf j}
357 >        (t + \delta t / 2)\delta t \cdot
358 >                \overleftrightarrow{\mathsf{I}}^{-1} \right),
359 > \end{array}\right.
360 > \end{equation*}
361 > where $\overleftrightarrow{\mathsf{I}}^{-1}$ is the inverse of the
362 > moment of inertia tensor. The $\mathrm{rotate}$ function is the
363 > product of rotations about the three body-fixed axes,
364 > \begin{equation}
365 > \mathrm{rotate}({\bf a}) = \mathsf{G}_x(a_x / 2) \cdot
366 > \mathsf{G}_y(a_y / 2) \cdot \mathsf{G}_z(a_z) \cdot \mathsf{G}_y(a_y /
367 > 2) \cdot \mathsf{G}_x(a_x /2),
368 > \label{eq:dlmTrot}
369 > \end{equation}
370 > where each rotational propagator, $\mathsf{G}_\alpha(\theta)$, rotates
371 > both the rotation matrix ($\mathsf{A}$) and the body-fixed angular
372 > momentum (${\bf j}$) by an angle $\theta$ around body-fixed axis
373 > $\alpha$,
374 > \begin{equation}
375 > \mathsf{G}_\alpha( \theta ) = \left\{
376 > \begin{array}{l@{\quad\leftarrow\quad}l}
377 > \mathsf{A}(t) & \mathsf{A}(0) \cdot \mathsf{R}_\alpha(\theta)^\textrm{T},\\
378 > {\bf j}(t) & \mathsf{R}_\alpha(\theta) \cdot {\bf j}(0).
379 > \end{array}
380 > \right.
381 > \end{equation}
382 > $\mathsf{R}_\alpha$ is a quadratic approximation to the single-axis
383 > rotation matrix.  For example, in the small-angle limit, the rotation
384 > matrix around the body-fixed x-axis can be approximated as
385 > \begin{equation}
386 > \mathsf{R}_x(\theta) \approx \left(
387 > \begin{array}{ccc}
388 > 1 & 0 & 0 \\
389 > 0 & \frac{1-\theta^2 / 4}{1 + \theta^2 / 4} & -\frac{\theta}{1+\theta^2 / 4} \\
390 > 0 & \frac{\theta}{1+\theta^2 / 4} & \frac{1-\theta^2 / 4}{1 + \theta^2 / 4}
391 > \end{array}
392 > \right).
393 > \end{equation}
394 > The remaining rotations follow in a straightforward manner. As seen
395 > from the form of equation~(\ref{eq:dlmTrot}), the {\sc dlm} method
396 > uses a Trotter factorization of the orientational
397 > propagator.\cite{Trotter59} This has three effects:
398 > \begin{enumerate}
399 > \item the integrator is area-preserving in phase space (i.e. it is
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 > $\mathcal{O}\left(\delta t^3\right)$ for time steps of length $\delta t$.
404 > \end{enumerate}
405 >
406 > After the initial half-step ({\tt moveA}), the forces and torques are
407 > evaluated for all of the particles. Once completed, the velocities can
408 > be advanced to complete the second-half of the two-part algorithm
409 > ({\tt moveB}), resulting an a completed full step of both the
410 > positions and momenta,
411 > \begin{equation*}
412 > {\tt moveB} = \left\{\begin{array}{r@{\quad\leftarrow\quad}l}
413 > {\bf v}\left(t + \delta t \right) &
414 >        {\bf v}\left(t + \delta t / 2 \right)
415 >        + \left({\bf f}(t + \delta t) / m \right)(\delta t/2), \\
416 > %
417 > {\bf j}\left(t + \delta t \right) &
418 >        {\bf j}\left(t + \delta t / 2 \right)
419 >        + \boldsymbol{\tau}^b(t + \delta t)(\delta t/2).
420 > \end{array}\right.
421 > \end{equation*}
422 >
423 > The matrix rotations used in the {\sc dlm} method end up being more
424 > costly computationally than the simpler arithmetic quaternion
425 > propagation. With the same time step, a 1024-molecule water simulation
426 > incurs approximately a 10\% increase in computation time using the
427 > {\sc dlm} method in place of quaternions. This cost is more than
428 > justified when comparing the energy conservation achieved by the two
429 > methods. Figure \ref{fig:quatdlm} provides a comparative analysis of
430 > the {\sc dlm} method versus the traditional quaternion scheme.
431 >
432 > \begin{figure}
433 > \centering
434 > \includegraphics[width=3.5in]{./figures/dlmVsQuat.pdf}
435 > \caption[Energy conservation analysis of the {\sc dlm} and quaternion
436 > integration methods]{Analysis of the energy conservation of the {\sc
437 > dlm} and quaternion integration methods.  $\delta \mathrm{E}_1$ is the
438 > linear drift in energy over time and $\delta \mathrm{E}_0$ is the
439 > standard deviation of energy fluctuations around this drift.  All
440 > simulations were of a 1024 SSD water system at 298 K starting from the
441 > same initial configuration. Note that the {\sc dlm} method provides
442 > more than an order-of-magnitude improvement in both the energy drift
443 > and the size of the energy fluctuations when compared with the
444 > quaternion method at any given time step.  At time steps larger than 4
445 > fs, the quaternion scheme resulted in rapidly rising energies which
446 > eventually lead to simulation failure.  Using the {\sc dlm} method,
447 > time steps up to 8 fs can be taken before this behavior is evident.}
448 > \label{fig:quatdlm}
449 > \end{figure}
450 >
451 > In figure \ref{fig:quatdlm}, $\delta \mbox{E}_1$ is a measure of the
452 > linear energy drift in units of $\mbox{kcal mol}^{-1}$ per particle
453 > over a nanosecond of simulation time, and $\delta \mbox{E}_0$ is the
454 > standard deviation of the energy fluctuations in units of $\mbox{kcal
455 > mol}^{-1}$ per particle. In the top plot, it is apparent that the
456 > energy drift is reduced by a significant amount (2 to 3 orders of
457 > magnitude improvement at all tested time steps) by choosing the {\sc
458 > dlm} method over the simple non-symplectic quaternion integration
459 > method.  In addition to this improvement in energy drift, the
460 > fluctuations in the total energy are also dampened by 1 to 2 orders of
461 > magnitude by utilizing the {\sc dlm} method.
462 >
463 > \begin{figure}
464 > \centering
465 > \includegraphics[width=\linewidth]{./figures/compCost.pdf}
466 > \caption[Energy drift as a function of required simulation run
467 > time]{Energy drift as a function of required simulation run time.
468 > $\delta \mathrm{E}_1$ is the linear drift in energy over time.
469 > Simulations were performed on a single 2.5 GHz Pentium 4
470 > processor. Simulation time comparisons can be made by tracing
471 > horizontally from one curve to the other. For example, a simulation
472 > that takes 24 hours using the {\sc dlm} method will take roughly
473 > 210 hours using the simple quaternion method if the same degree of
474 > energy conservation is desired.}
475 > \label{fig:cpuCost}
476 > \end{figure}
477 > Although the {\sc dlm} method is more computationally expensive than
478 > the traditional quaternion scheme for propagating of a time step,
479 > consideration of the computational cost for a long simulation with a
480 > particular level of energy conservation is in order.  A plot of energy
481 > drift versus computational cost was generated
482 > (Fig.~\ref{fig:cpuCost}). This figure provides an estimate of the CPU
483 > time required under the two integration schemes for 1 nanosecond of
484 > simulation time for the model 1024-molecule system.  By choosing a
485 > desired energy drift value it is possible to determine the CPU time
486 > required for both methods. If a $\delta \mbox{E}_1$ of
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 {\sc dlm}
491 > integration scheme.
492 >
493 > \section{Accumulating Interactions}
494 >
495 > In the force calculation between {\tt moveA} and {\tt moveB} mentioned
496 > in section \ref{sec:IntroIntegrate}, we need to accumulate the
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 $\mathcal{O}(N(N-1)/2)$ 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
511 > \includegraphics[width=3.5in]{./figures/sphericalCut.pdf}
512 > \caption{When using a spherical cutoff, only particles within a chosen
513 > cutoff radius distance, $R_\textrm{c}$, of the central particle are
514 > included in the pairwise summation. This reduces a problem that scales
515 > by $\sim\mathcal{O}(N^2)$ to one that scales by $\sim\mathcal{O}(N)$.}
516 > \label{fig:sphereCut}
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 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 >
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
541 > possible causes.\cite{Harvey98,Sagui99}
542 >
543 > \subsection{Correcting Cutoff Discontinuities}
544 > \begin{figure}
545 > \centering
546 > \includegraphics[width=3.5in]{./figures/ljCutoffSquare.pdf}
547 > \caption{The common methods to smooth the potential discontinuity
548 > introduced when using a cutoff include a shifted potential, a shifted
549 > force, and a switching function. The shifted potential and shifted
550 > force both lift the whole potential so that it zeroes at
551 > $R_\textrm{c}$, thereby reducing the strength of the interaction. The
552 > (cubic) switching function only alters the potential in the switching
553 > region in order to smooth out the discontinuity.}
554 > \label{fig:ljCutoff}
555 > \end{figure}
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}, \\
569 >        0 & r_{ij} \geqslant R_\textrm{c},
570 > \end{array}\right.
571 > \end{equation}
572 > where
573 > \begin{equation}
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 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 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}) -
587 >                \left(\frac{d V(r_{ij})}{d r_{ij}}\right)_{r_{ij}=R_\textrm{c}}
588 >                (r_{ij} - R_\textrm{c}) & r_{ij} < R_\textrm{c}, \\
589 >        0 & r_{ij} \geqslant 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
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}, \\
602 >        \frac{(R_\textrm{c} + 2r_{ij} - 3R_\textrm{sw})
603 >                (R_\textrm{c} - r_{ij})^2}{(R_\textrm{c} - R_\textrm{sw})^3}
604 >                & R_\textrm{sw} < r_{ij} \leqslant R_\textrm{c}, \\
605 >        0 & r_{ij} > R_\textrm{c},
606 >        \end{array}\right.
607 > \end{equation}
608 > is sufficient to smooth the potential (again, see figure
609 > \ref{fig:ljCutoff}) and the forces by only perturbing the potential in
610 > the switching region. If smooth second derivatives are required, a
611 > higher order polynomial switching function (i.e. fifth order
612 > polynomial) can be used.\cite{Andrea83,Leach01} It should be noted
613 > that the higher the order of the polynomial, the more abrupt the
614 > switching transition.
615 >
616 > \subsection{\label{sec:LJCorrections}Long-Range Interaction Corrections}
617 >
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 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 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 correction techniques, is presented in
644 > chapter \ref{chap:electrostatics}.
645 >
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 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
653 > particles, they are not confined by walls of the ``simulation box''
654 > and can freely move away from the other particles. With no boundary
655 > considerations, particles moving outside of the simulation box
656 > enter a vacuum. This is correct behavior for cluster simulations in a
657 > vacuum; however, if we want to simulate bulk or spatially infinite
658 > systems, we need to use periodic boundary conditions.
659 >
660 > \begin{figure}
661 > \centering
662 > \includegraphics[width=4.5in]{./figures/periodicImage.pdf}
663 > \caption{With periodic boundary conditions imposed, when particles
664 > move out of one side the simulation box, they wrap back in the
665 > opposite side. In this manner, a finite system of particles behaves as
666 > an infinite system.}
667 > \label{fig:periodicImage}
668 > \end{figure}
669 > In periodic boundary conditions, as a particle moves outside one wall
670 > of the simulation box, the coordinates are remapped such that the
671 > particle enters the opposing side of the box. This process is easy to
672 > visualize in two dimensions as shown in figure \ref{fig:periodicImage}
673 > and can occur in three dimensions, though it is not as easy to
674 > visualize. Remapping the actual coordinates of the particles can be
675 > problematic in that the we are restricting the distance a particle can
676 > move from it's point of origin to a diagonal of the simulation
677 > box. Thus, even though we are not confining the system with hard
678 > walls, we are confining the particle coordinates to a particular
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 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 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
693 > the system and make connections between these properties and the
694 > experimental thermodynamic properties through statistical mechanics.
695 >
696 > The work presented in the later chapters use the canonical ($NVT$),
697 > isobaric-isothermal ($NPT$), and microcanonical ($NVE$) statistical
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 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}
712 > while the Gibbs free energy ($G$) depends on the $NPT$ partition
713 > function ($Q_{NPT}$),
714 > \begin{equation}
715 > G = -k_\textrm{B}T\ln Q_{NPT}.  
716 > \end{equation}
717 > It is also useful to note that the conserved quantities of the $NVT$
718 > and $NPT$ ensembles are related to the Helmholtz and Gibbs free
719 > energies respectively.\cite{Melchionna93}
720 >
721 > Integrating the equations of motion is a simple method to obtain a
722 > sequence of configurations that sample the chosen ensemble. For each
723 > of these configurations, we can calculate an instantaneous value for a
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. 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 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 > 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{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
751 > simulation packages available, both free and commercial, which
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 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 >
768 >

Diff Legend

Removed lines
+ Added lines
< Changed lines
> Changed lines