ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/chrisDissertation/Introduction.tex
Revision: 3023
Committed: Tue Sep 26 03:07:59 2006 UTC (17 years, 9 months ago) by chrisfen
Content type: application/x-tex
File size: 39049 byte(s)
Log Message:
Refinements.  Just a little proof checking left...

File Contents

# Content
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 were 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 chapter gives a general overview of molecular simulations, with
15 particular emphasis on considerations that need to be made in order to
16 apply the technique properly. This leads quite naturally into chapter
17 \ref{chap:electrostatics}, where we investigate correction techniques
18 for proper handling of long-ranged electrostatic interactions. In
19 particular we develop and evaluate some new simple pairwise
20 methods. These techniques make an appearance in the later chapters, as
21 they are applied to specific systems of interest, showing how it they
22 can improve the quality of various molecular simulations.
23
24 In chapter \ref{chap:water}, we focus on simple water models,
25 specifically the single-point soft sticky dipole (SSD) family of water
26 models. These single-point models are more efficient than the common
27 multi-point partial charge models and, in many cases, better capture
28 the dynamic properties of water. We discuss improvements to these
29 models in regards to long-range electrostatic corrections and show
30 that these models can work well with the techniques discussed in
31 chapter \ref{chap:electrostatics}. By investigating and improving
32 simple water models, we are extending the limits of computational
33 efficiency for systems that depend heavily on water calculations.
34
35 In chapter \ref{chap:ice}, we study a unique polymorph of ice that we
36 discovered while performing water simulations with the fast simple
37 water models discussed in the previous chapter. This form of ice,
38 which we called ``imaginary ice'' (Ice-$i$), has a low-density
39 structure which is different from any known polymorph from either
40 experiment or other simulations. In this study, we perform a free
41 energy analysis and see that this structure is in fact the
42 thermodynamically preferred form of ice for both the single-point and
43 commonly used multi-point water models under the chosen simulation
44 conditions. We also consider electrostatic corrections, again
45 including the techniques discussed in chapter
46 \ref{chap:electrostatics}, to see how they influence the free energy
47 results. This work, to some degree, addresses the appropriateness of
48 using these simplistic water models outside of the conditions for
49 which they were developed.
50
51 Finally, in chapter \ref{chap:conclusion}, we summarize the work
52 presented in the previous chapters and connect ideas together with
53 some final comments. The supporting information follows in the
54 appendix, and it gives a more detailed look at systems discussed in
55 chapter \ref{chap:electrostatics}.
56
57 \section{On Molecular Simulation}
58
59 In order to advance our understanding of natural chemical and physical
60 processes, researchers develop explanations for events observed
61 experimentally. These hypotheses, supported by a body of corroborating
62 observations, can develop into accepted theories for how these
63 processes occur. This validation process, as well as testing the
64 limits of these theories, is essential in developing a firm foundation
65 for our knowledge. Theories involving molecular scale systems cause
66 particular difficulties in this process because their size and often
67 fast motions make them difficult to observe experimentally. One useful
68 tool for addressing these difficulties is computer simulation.
69
70 In computer simulations, we can develop models from either the theory
71 or experimental knowledge and then test them in a controlled
72 environment. Work done in this manner allows us to further refine
73 theories, more accurately represent what is happening in experimental
74 observations, and even make predictions about what one will see in
75 experiments. Thus, computer simulations of molecular systems act as a
76 bridge between theory and experiment.
77
78 Depending on the system of interest, there are a variety of different
79 computational techniques that can used to test and gather information
80 from the developed models. In the study of classical systems, the two
81 most commonly used techniques are Monte Carlo and molecular
82 dynamics. Both of these methods operate by calculating interactions
83 between particles of our model systems; however, the progression of
84 the simulation under the different techniques is vastly
85 different. Monte Carlo operates through random configuration changes
86 that that follow rules adhering to a specific statistical mechanics
87 ensemble, while molecular dynamics is chiefly concerned with solving
88 the classic equations of motion to move between configurations within
89 an ensemble. Thermodynamic properties can be calculated from both
90 techniques; but because of the random nature of Monte Carlo, only
91 molecular dynamics can be used to investigate dynamical
92 quantities. The research presented in the following chapters utilized
93 molecular dynamics near exclusively, so we will present a general
94 introduction to molecular dynamics. There are several resources
95 available for those desiring a more in-depth presentation of either of
96 these techniques.\cite{Allen87,Frenkel02,Leach01}
97
98 \section{\label{sec:MolecularDynamics}Molecular Dynamics}
99
100 As stated above, in molecular dynamics we focus on evolving
101 configurations of molecules over time. In order to use this as a tool
102 for understanding experiments and testing theories, we want the
103 configuration to evolve in a manner that mimics real molecular
104 systems. To do this, we start with clarifying what we know about a
105 given configuration of particles at time $t_1$, basically that each
106 particle in the configuration has a position ($\mathbf{q}$) and velocity
107 ($\dot{\mathbf{q}}$). We now want to know what the configuration will be at
108 time $t_2$. To find out, we need the classical equations of
109 motion, and one useful formulation of them is the Lagrangian form.
110
111 The Lagrangian ($L$) is a function of the positions and velocities that
112 takes the form,
113 \begin{equation}
114 L = K - V,
115 \label{eq:lagrangian}
116 \end{equation}
117 where $K$ is the kinetic energy and $V$ is the potential energy. We
118 can use Hamilton's principle, which states that the integral of the
119 Lagrangian over time has a stationary value for the correct path of
120 motion, to say that the variation of the integral of the Lagrangian
121 over time is zero,\cite{Tolman38}
122 \begin{equation}
123 \delta\int_{t_1}^{t_2}L(\mathbf{q},\dot{\mathbf{q}})dt = 0.
124 \end{equation}
125 The variation can be transferred to the variables that make up the
126 Lagrangian,
127 \begin{equation}
128 \int_{t_1}^{t_2}\sum_{i=1}^{3N}\left(
129 \frac{\partial L}{\partial \mathbf{q}_i}\delta \mathbf{q}_i
130 + \frac{\partial L}{\partial \dot{\mathbf{q}}_i}\delta
131 \dot{\mathbf{q}}_i\right)dt = 0.
132 \end{equation}
133 Using the fact that $\dot{\mathbf{q}}$ is the derivative of
134 $\mathbf{q}$ with respect to time and integrating the second partial
135 derivative in the parenthesis by parts, this equation simplifies to
136 \begin{equation}
137 \int_{t_1}^{t_2}\sum_{i=1}^{3N}\left(
138 \frac{d}{dt}\frac{\partial L}{\partial \dot{\mathbf{q}}_i}
139 - \frac{\partial L}{\partial \mathbf{q}_i}\right)
140 \delta {\mathbf{q}}_i dt = 0,
141 \end{equation}
142 and since each variable is independent, we can separate the
143 contribution from each of the variables:
144 \begin{equation}
145 \frac{d}{dt}\frac{\partial L}{\partial \dot{\mathbf{q}}_i}
146 - \frac{\partial L}{\partial \mathbf{q}_i} = 0
147 \quad\quad(i=1,2,\dots,3N).
148 \label{eq:formulation}
149 \end{equation}
150 To obtain the classical equations of motion for the particles, we can
151 substitute equation (\ref{eq:lagrangian}) into the above equation with
152 $m\dot{\mathbf{r}}^2/2$ for the kinetic energy, giving
153 \begin{equation}
154 \frac{d}{dt}(m\dot{\mathbf{r}})+\frac{dV}{d\mathbf{r}}=0,
155 \end{equation}
156 or more recognizably,
157 \begin{equation}
158 \mathbf{f} = m\mathbf{a},
159 \end{equation}
160 where $\mathbf{f} = -dV/d\mathbf{r}$ and $\mathbf{a} =
161 d^2\mathbf{r}/dt^2$. This Lagrangian formulation shown in equation
162 (\ref{eq:formulation}) is generalized, and it can be used to determine
163 equations of motion in forms outside of the typical Cartesian case
164 shown here.
165
166 \subsection{\label{sec:Verlet}Verlet Integration}
167
168 In order to perform molecular dynamics, we need an algorithm that
169 integrates the equations of motion described above. Ideal algorithms
170 are both simple in implementation and highly accurate. There have been
171 a large number of algorithms developed for this purpose; however, for
172 reasons discussed below, we are going to focus on the Verlet class of
173 integrators.\cite{Gear66,Beeman76,Berendsen86,Allen87,Verlet67,Swope82}
174
175 In Verlet's original study of computer ``experiments'', he directly
176 integrated the Newtonian second order differential equation of motion,
177 \begin{equation}
178 m\frac{d^2\mathbf{r}_i}{dt^2} = \sum_{j\ne i}\mathbf{f}(r_{ij}),
179 \end{equation}
180 with the following simple algorithm:
181 \begin{equation}
182 \mathbf{r}_i(t+\delta t) = -\mathbf{r}_i(t-\delta t) + 2\mathbf{r}_i(t)
183 + \sum_{j\ne i}\mathbf{f}(r_{ij}(t))\delta t^2,
184 \label{eq:verlet}
185 \end{equation}
186 where $\delta t$ is the time step of integration.\cite{Verlet67} It is
187 interesting to note that equation (\ref{eq:verlet}) does not include
188 velocities, and this makes sense since they are not present in the
189 differential equation. The velocities are necessary for the
190 calculation of the kinetic energy and can be calculated at each time
191 step with the equation:
192 \begin{equation}
193 \mathbf{v}_i(t) = \frac{\mathbf{r}_i(t+\delta t)-\mathbf{r}_i(t-\delta t)}
194 {2\delta t}.
195 \end{equation}
196
197 Like the equation of motion it solves, the Verlet algorithm has the
198 beneficial property of being time-reversible, meaning that you can
199 integrate the configuration forward and then backward and end up at
200 the original configuration. Some other methods for integration, like
201 predictor-corrector methods, lack this property in that they require
202 higher order information that is discarded after integrating
203 steps. Another interesting property of this algorithm is that it is
204 symplectic, meaning that it preserves area in phase-space. Symplectic
205 algorithms keep the system evolving in the region of phase-space
206 dictated by the ensemble and enjoy a greater degree of energy
207 conservation.\cite{Frenkel02}
208
209 While the error in the positions calculated using the Verlet algorithm
210 is small ($\mathcal{O}(\delta t^4)$), the error in the velocities is
211 substantially larger ($\mathcal{O}(\delta t^2)$).\cite{Allen87} Swope
212 {\it et al.} developed a corrected version of this algorithm, the
213 `velocity Verlet' algorithm, which improves the error of the velocity
214 calculation and thus the energy conservation.\cite{Swope82} This
215 algorithm involves a full step of the positions,
216 \begin{equation}
217 \mathbf{r}(t+\delta t) = \mathbf{r}(t) + \delta t\mathbf{v}(t)
218 + \frac{1}{2}\delta t^2\mathbf{a}(t),
219 \end{equation}
220 and a half step of the velocities,
221 \begin{equation}
222 \mathbf{v}\left(t+\frac{1}{2}\delta t\right) = \mathbf{v}(t)
223 + \frac{1}{2}\delta t\mathbf{a}(t).
224 \end{equation}
225 After forces are calculated at the new positions, the velocities can
226 be updated to a full step,
227 \begin{equation}
228 \mathbf{v}(t+\delta t) = \mathbf{v}\left(t+\frac{1}{2}\delta t\right)
229 + \frac{1}{2}\delta t\mathbf{a}(t+\delta t).
230 \end{equation}
231 By integrating in this manner, the error in the velocities reduces to
232 $\mathcal{O}(\delta t^3)$. It should be noted that the error in the
233 positions increases to $\mathcal{O}(\delta t^3)$, but the resulting
234 improvement in the energies coupled with the maintained simplicity,
235 time-reversibility, and symplectic nature make it an improvement over
236 the original form.
237
238 \subsection{\label{sec:IntroIntegrate}Rigid Body Motion}
239
240 Rigid bodies are non-spherical particles or collections of particles
241 (e.g. $\mbox{C}_{60}$) that have a constant internal potential and
242 move collectively.\cite{Goldstein01} Discounting iterative constraint
243 procedures like {\sc shake} and {\sc rattle} for approximating rigid
244 bodies, they are not included in most simulation packages because of
245 the algorithmic complexity involved in propagating orientational
246 degrees of freedom.\cite{Ryckaert77,Andersen83,Krautler01} Integrators
247 which propagate orientational motion with an acceptable level of
248 energy conservation for molecular dynamics are relatively new
249 inventions.
250
251 Moving a rigid body involves determination of both the force and
252 torque applied by the surroundings, which directly affect the
253 translational and rotational motion in turn. In order to accumulate
254 the total force on a rigid body, the external forces and torques must
255 first be calculated for all the internal particles. The total force on
256 the rigid body is simply the sum of these external forces.
257 Accumulation of the total torque on the rigid body is more complex
258 than the force because the torque is applied to the center of mass of
259 the rigid body. The space-fixed torque on rigid body $i$ is
260 \begin{equation}
261 \boldsymbol{\tau}_i=
262 \sum_{a}\biggl[(\mathbf{r}_{ia}-\mathbf{r}_i)\times \mathbf{f}_{ia}
263 + \boldsymbol{\tau}_{ia}\biggr],
264 \label{eq:torqueAccumulate}
265 \end{equation}
266 where $\boldsymbol{\tau}_i$ and $\mathbf{r}_i$ are the torque on and
267 position of the center of mass respectively, while $\mathbf{f}_{ia}$,
268 $\mathbf{r}_{ia}$, and $\boldsymbol{\tau}_{ia}$ are the force on,
269 position of, and torque on the component particles.
270
271 The summation of the total torque is done in the body fixed axis. In
272 order to move between the space fixed and body fixed coordinate axes,
273 parameters describing the orientation must be maintained for each
274 rigid body. At a minimum, the rotation matrix ($\mathsf{A}$) can be
275 described by the three Euler angles ($\phi, \theta,$ and $\psi$),
276 where the elements of $\mathsf{A}$ are composed of trigonometric
277 operations involving $\phi, \theta,$ and $\psi$.\cite{Goldstein01}
278 Direct propagation of the Euler angles has a known $1/\sin\theta$
279 divergence in the equations of motion for $\phi$ and $\psi$, leading
280 to numerical instabilities any time one of the directional atoms or
281 rigid bodies has an orientation near $\theta=0$ or
282 $\theta=\pi$.\cite{Allen87} One of the most practical ways to avoid
283 this ``gimbal point'' is to switch to another angular set defining the
284 orientation of the rigid body near this point.\cite{Barojas73} This
285 procedure results in additional book-keeping and increased algorithm
286 complexity. In the search for more elegant alternative methods, Evans
287 proposed the use of quaternions to describe and propagate
288 orientational motion.\cite{Evans77}
289
290 The quaternion method for integration involves a four dimensional
291 representation of the orientation of a rigid
292 body.\cite{Evans77,Evans77b,Allen87} Thus, the elements of
293 $\mathsf{A}$ can be expressed as arithmetic operations involving the
294 four quaternions ($q_0, q_1, q_2,$ and $q_3$),
295 \begin{equation}
296 \mathsf{A} = \left( \begin{array}{l@{\quad}l@{\quad}l}
297 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) \\
298 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) \\
299 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 \\
300 \end{array}\right).
301 \end{equation}
302 Integration of the equations of motion involves a series of arithmetic
303 operations involving the quaternions and angular momenta and leads to
304 performance enhancements over Euler angles, particularly for very
305 small systems.\cite{Evans77} This integration method works well for
306 propagating orientational motion in the canonical ensemble ($NVT$);
307 however, energy conservation concerns arise when using the simple
308 quaternion technique under the microcanonical ($NVE$) ensemble. An
309 earlier implementation of our simulation code utilized quaternions for
310 propagation of rotational motion; however, a detailed investigation
311 showed that they resulted in a steady drift in the total energy,
312 something that has been observed by Kol {\it et al.} (also see
313 section~\ref{sec:waterSimMethods}).\cite{Kol97}
314
315 Because of the outlined issues involving integration of the
316 orientational motion using both Euler angles and quaternions, we
317 decided to focus on a relatively new scheme that propagates the entire
318 nine parameter rotation matrix. This techniques is a velocity-Verlet
319 version of the symplectic splitting method proposed by Dullweber,
320 Leimkuhler and McLachlan ({\sc dlm}).\cite{Dullweber97} When there are
321 no directional atoms or rigid bodies present in the simulation, this
322 integrator becomes the standard velocity-Verlet integrator which is
323 known to effectively sample the microcanonical ($NVE$)
324 ensemble.\cite{Frenkel02}
325
326 The key aspect of the integration method proposed by Dullweber
327 \emph{et al.} is that the entire $3 \times 3$ rotation matrix is
328 propagated from one time step to the next. In the past, this would not
329 have been as feasible, since the rotation matrix for a single body has
330 nine elements compared with the more memory-efficient methods (using
331 three Euler angles or four quaternions). Computer memory has become
332 much less costly in recent years, and this can be translated into
333 substantial benefits in energy conservation.
334
335 The integration of the equations of motion is carried out in a
336 velocity-Verlet style two-part algorithm.\cite{Swope82} The first-part
337 ({\tt moveA}) consists of a half-step ($t + \delta t/2$) of the linear
338 velocity (${\bf v}$) and angular momenta ({\bf j}) and a full-step ($t
339 + \delta t$) of the positions ({\bf r}) and rotation matrix,
340 \begin{equation*}
341 {\tt moveA} = \left\{\begin{array}{r@{\quad\leftarrow\quad}l}
342 {\bf v}\left(t + \delta t / 2\right) & {\bf v}(t)
343 + \left( {\bf f}(t) / m \right)(\delta t/2), \\
344 %
345 {\bf r}(t + \delta t) & {\bf r}(t)
346 + {\bf v}\left(t + \delta t / 2 \right)\delta t, \\
347 %
348 {\bf j}\left(t + \delta t / 2 \right) & {\bf j}(t)
349 + \boldsymbol{\tau}^b(t)(\delta t/2), \\
350 %
351 \mathsf{A}(t + \delta t) & \mathrm{rotate}\left( {\bf j}
352 (t + \delta t / 2)\delta t \cdot
353 \overleftrightarrow{\mathsf{I}}^{-1} \right),
354 \end{array}\right.
355 \end{equation*}
356 where $\overleftrightarrow{\mathsf{I}}^{-1}$ is the inverse of the
357 moment of inertia tensor. The $\mathrm{rotate}$ function is the
358 product of rotations about the three body-fixed axes,
359 \begin{equation}
360 \mathrm{rotate}({\bf a}) = \mathsf{G}_x(a_x / 2) \cdot
361 \mathsf{G}_y(a_y / 2) \cdot \mathsf{G}_z(a_z) \cdot \mathsf{G}_y(a_y /
362 2) \cdot \mathsf{G}_x(a_x /2),
363 \label{eq:dlmTrot}
364 \end{equation}
365 where each rotational propagator, $\mathsf{G}_\alpha(\theta)$, rotates
366 both the rotation matrix ($\mathsf{A}$) and the body-fixed angular
367 momentum (${\bf j}$) by an angle $\theta$ around body-fixed axis
368 $\alpha$,
369 \begin{equation}
370 \mathsf{G}_\alpha( \theta ) = \left\{
371 \begin{array}{l@{\quad\leftarrow\quad}l}
372 \mathsf{A}(t) & \mathsf{A}(0) \cdot \mathsf{R}_\alpha(\theta)^\textrm{T},\\
373 {\bf j}(t) & \mathsf{R}_\alpha(\theta) \cdot {\bf j}(0).
374 \end{array}
375 \right.
376 \end{equation}
377 $\mathsf{R}_\alpha$ is a quadratic approximation to the single-axis
378 rotation matrix. For example, in the small-angle limit, the rotation
379 matrix around the body-fixed x-axis can be approximated as
380 \begin{equation}
381 \mathsf{R}_x(\theta) \approx \left(
382 \begin{array}{ccc}
383 1 & 0 & 0 \\
384 0 & \frac{1-\theta^2 / 4}{1 + \theta^2 / 4} & -\frac{\theta}{1+\theta^2 / 4} \\
385 0 & \frac{\theta}{1+\theta^2 / 4} & \frac{1-\theta^2 / 4}{1 + \theta^2 / 4}
386 \end{array}
387 \right).
388 \end{equation}
389 The remaining rotations follow in a straightforward manner. As seen
390 from the form of equation~(\ref{eq:dlmTrot}), the {\sc dlm} method
391 uses a Trotter factorization of the orientational
392 propagator.\cite{Trotter59} This has three effects:
393 \begin{enumerate}
394 \item the integrator is area-preserving in phase space (i.e. it is
395 {\it symplectic}),
396 \item the integrator is time-{\it reversible}, and
397 \item the error for a single time step is of order
398 $\mathcal{O}\left(\delta t^3\right)$ for time steps of length $\delta t$.
399 \end{enumerate}
400
401 After the initial half-step ({\tt moveA}), the forces and torques are
402 evaluated for all of the particles. Once completed, the velocities can
403 be advanced to complete the second-half of the two-part algorithm
404 ({\tt moveB}), resulting an a completed full step of both the
405 positions and momenta,
406 \begin{equation*}
407 {\tt moveB} = \left\{\begin{array}{r@{\quad\leftarrow\quad}l}
408 {\bf v}\left(t + \delta t \right) &
409 {\bf v}\left(t + \delta t / 2 \right)
410 + \left({\bf f}(t + \delta t) / m \right)(\delta t/2), \\
411 %
412 {\bf j}\left(t + \delta t \right) &
413 {\bf j}\left(t + \delta t / 2 \right)
414 + \boldsymbol{\tau}^b(t + \delta t)(\delta t/2).
415 \end{array}\right.
416 \end{equation*}
417
418 The matrix rotations used in the {\sc dlm} method end up being more
419 costly computationally than the simpler arithmetic quaternion
420 propagation. With the same time step, a 1024-molecule water simulation
421 incurs approximately a 10\% increase in computation time using the
422 {\sc dlm} method in place of quaternions. This cost is more than
423 justified when comparing the energy conservation achieved by the two
424 methods. Figure \ref{fig:quatdlm} provides a comparative analysis of
425 the {\sc dlm} method versus the traditional quaternion scheme.
426
427 \begin{figure}
428 \centering
429 \includegraphics[width=3.5in]{./figures/dlmVsQuat.pdf}
430 \caption[Energy conservation analysis of the {\sc dlm} and quaternion
431 integration methods]{Analysis of the energy conservation of the {\sc
432 dlm} and quaternion integration methods. $\delta \mathrm{E}_1$ is the
433 linear drift in energy over time and $\delta \mathrm{E}_0$ is the
434 standard deviation of energy fluctuations around this drift. All
435 simulations were of a 1024 SSD water system at 298 K starting from the
436 same initial configuration. Note that the {\sc dlm} method provides
437 more than an order-of-magnitude improvement in both the energy drift
438 and the size of the energy fluctuations when compared with the
439 quaternion method at any given time step. At time steps larger than 4
440 fs, the quaternion scheme resulted in rapidly rising energies which
441 eventually lead to simulation failure. Using the {\sc dlm} method,
442 time steps up to 8 fs can be taken before this behavior is evident.}
443 \label{fig:quatdlm}
444 \end{figure}
445
446 In figure \ref{fig:quatdlm}, $\delta \mbox{E}_1$ is a measure of the
447 linear energy drift in units of $\mbox{kcal mol}^{-1}$ per particle
448 over a nanosecond of simulation time, and $\delta \mbox{E}_0$ is the
449 standard deviation of the energy fluctuations in units of $\mbox{kcal
450 mol}^{-1}$ per particle. In the top plot, it is apparent that the
451 energy drift is reduced by a significant amount (2 to 3 orders of
452 magnitude improvement at all tested time steps) by choosing the {\sc
453 dlm} method over the simple non-symplectic quaternion integration
454 method. In addition to this improvement in energy drift, the
455 fluctuations in the total energy are also dampened by 1 to 2 orders of
456 magnitude by utilizing the {\sc dlm} method.
457
458 \begin{figure}
459 \centering
460 \includegraphics[width=\linewidth]{./figures/compCost.pdf}
461 \caption[Energy drift as a function of required simulation run
462 time]{Energy drift as a function of required simulation run time.
463 $\delta \mathrm{E}_1$ is the linear drift in energy over time.
464 Simulations were performed on a single 2.5 GHz Pentium 4
465 processor. Simulation time comparisons can be made by tracing
466 horizontally from one curve to the other. For example, a simulation
467 that takes 24 hours using the {\sc dlm} method will take roughly
468 210 hours using the simple quaternion method if the same degree of
469 energy conservation is desired.}
470 \label{fig:cpuCost}
471 \end{figure}
472 Although the {\sc dlm} method is more computationally expensive than
473 the traditional quaternion scheme for propagating of a time step,
474 consideration of the computational cost for a long simulation with a
475 particular level of energy conservation is in order. A plot of energy
476 drift versus computational cost was generated
477 (Fig.~\ref{fig:cpuCost}). This figure provides an estimate of the CPU
478 time required under the two integration schemes for 1 nanosecond of
479 simulation time for the model 1024-molecule system. By choosing a
480 desired energy drift value it is possible to determine the CPU time
481 required for both methods. If a $\delta \mbox{E}_1$ of
482 0.001~kcal~mol$^{-1}$ per particle is desired, a nanosecond of
483 simulation time will require ~19 hours of CPU time with the {\sc dlm}
484 integrator, while the quaternion scheme will require ~154 hours of CPU
485 time. This demonstrates the computational advantage of the {\sc dlm}
486 integration scheme.
487
488 \section{Accumulating Interactions}
489
490 In the force calculation between {\tt moveA} and {\tt moveB} mentioned
491 in section \ref{sec:IntroIntegrate}, we need to accumulate the
492 potential and forces (and torques if the particle is a rigid body or
493 multipole) on each particle from their surroundings. This can quickly
494 become a cumbersome task for large systems since the number of pair
495 interactions increases by $\mathcal{O}(N(N-1)/2)$ if you accumulate
496 interactions between all particles in the system. (Note the
497 utilization of Newton's third law to reduce the interaction number
498 from $\mathcal{O}(N^2)$.) The case of periodic boundary conditions
499 further complicates matters by turning the finite system into an
500 infinitely repeating one. Fortunately, we can reduce the scale of this
501 problem by using spherical cutoff methods.
502
503 \begin{figure}
504 \centering
505 \includegraphics[width=3.5in]{./figures/sphericalCut.pdf}
506 \caption{When using a spherical cutoff, only particles within a chosen
507 cutoff radius distance, $R_\textrm{c}$, of the central particle are
508 included in the pairwise summation. This reduces a problem that scales
509 by $\sim\mathcal{O}(N^2)$ to one that scales by $\sim\mathcal{O}(N)$.}
510 \label{fig:sphereCut}
511 \end{figure}
512 With spherical cutoffs, rather than accumulating the full set of
513 interactions between all particles in the simulation, we only
514 explicitly consider interactions between particles separated by less
515 than a specified cutoff radius distance, $R_\textrm{c}$, (see figure
516 \ref{fig:sphereCut}). This reduces the scaling of the interaction to
517 $\mathcal{O}(N\cdot\textrm{c})$, where `c' is a value that depends on
518 the size of $R_\textrm{c}$ (c $\approx R_\textrm{c}^3$). Determination
519 of which particles are within the cutoff is also an issue, because
520 this process requires a full loop over all $N(N-1)/2$ pairs. To reduce
521 the this expense, we can use neighbor lists.\cite{Verlet67,Thompson83}
522 With neighbor lists, we have a second list of particles built from a
523 list radius $R_\textrm{l}$, which is larger than $R_\textrm{c}$. Once
524 any particle within $R_\textrm{l}$ moves half the distance of
525 $R_\textrm{l}-R_\textrm{c}$ (the ``skin'' thickness), we rebuild the
526 list with the full $N(N-1)/2$ loop.\cite{Verlet67} With an appropriate
527 skin thickness, these updates are only performed every $\sim$20 time
528 steps, significantly reducing the time spent on pair-list bookkeeping
529 operations.\cite{Allen87} If these neighbor lists are utilized, it is
530 important that these list updates occur regularly. Incorrect
531 application of this technique leads to non-physical dynamics, such as
532 the ``flying block of ice'' behavior for which improper neighbor list
533 handling was identified a one of the possible
534 causes.\cite{Harvey98,Sagui99}
535
536 \subsection{Correcting Cutoff Discontinuities}
537 \begin{figure}
538 \centering
539 \includegraphics[width=3.5in]{./figures/ljCutoffSquare.pdf}
540 \caption{The common methods to smooth the potential discontinuity
541 introduced when using a cutoff include a shifted potential, a shifted
542 force, and a switching function. The shifted potential and shifted
543 force both lift the whole potential so that it zeroes at
544 $R_\textrm{c}$, thereby reducing the strength of the interaction. The
545 (cubic) switching function only alters the potential in the switching
546 region in order to smooth out the discontinuity.}
547 \label{fig:ljCutoff}
548 \end{figure}
549 As the distance between a pair of particles fluctuates around
550 $R_\textrm{c}$, there will be sudden discontinuous jumps in the
551 potential (and forces) due to their inclusion and exclusion from the
552 interaction loop. In order to prevent heating and poor energy
553 conservation in the simulations, this discontinuity needs to be
554 smoothed out. There are several ways to modify the potential function
555 to eliminate these discontinuties, and the easiest methods is shifting
556 the potential. To shift the potential, we simply subtract out the
557 value we calculate at $R_\textrm{c}$ from the whole potential. The
558 shifted form of the Lennard-Jones potential is
559 \begin{equation}
560 V_\textrm{SLJ} = \left\{\begin{array}{l@{\quad\quad}l}
561 V_\textrm{LJ}(r_{ij}) - V_\textrm{LJ}(R_\textrm{c}) & r_{ij} < R_\textrm{c}, \\
562 0 & r_{ij} \geqslant R_\textrm{c},
563 \end{array}\right.
564 \end{equation}
565 where
566 \begin{equation}
567 V_\textrm{LJ}(r_{ij}) =
568 4\epsilon\left[\left(\frac{\sigma}{r_{ij}}\right)^{12} -
569 \left(\frac{\sigma}{r_{ij}}\right)^6\right].
570 \end{equation}
571 In figure \ref{fig:ljCutoff}, the shifted form of the potential
572 reaches zero at the cutoff radius at the expense of the correct
573 magnitude inside $R_\textrm{c}$. This correction method also does
574 nothing to correct the discontinuity in the forces. We can account for
575 this force discontinuity by shifting in the by applying the shifting
576 in the forces as well as in the potential via
577 \begin{equation}
578 V_\textrm{SFLJ} = \left\{\begin{array}{l@{\quad\quad}l}
579 V_\textrm{LJ}({r_{ij}}) - V_\textrm{LJ}(R_\textrm{c}) -
580 \left(\frac{d V(r_{ij})}{d r_{ij}}\right)_{r_{ij}=R_\textrm{c}}
581 (r_{ij} - R_\textrm{c}) & r_{ij} < R_\textrm{c}, \\
582 0 & r_{ij} \geqslant R_\textrm{c}.
583 \end{array}\right.
584 \end{equation}
585 The forces are continuous with this potential; however, the inclusion
586 of the derivative term distorts the potential even further than the
587 simple shifting as shown in figure \ref{fig:ljCutoff}. The method for
588 correcting the discontinuity which results in the smallest
589 perturbation in both the potential and the forces is the use of a
590 switching function. The cubic switching function,
591 \begin{equation}
592 S(r) = \left\{\begin{array}{l@{\quad\quad}l}
593 1 & r_{ij} \leqslant R_\textrm{sw}, \\
594 \frac{(R_\textrm{c} + 2r_{ij} - 3R_\textrm{sw})
595 (R_\textrm{c} - r_{ij})^2}{(R_\textrm{c} - R_\textrm{sw})^3}
596 & R_\textrm{sw} < r_{ij} \leqslant R_\textrm{c}, \\
597 0 & r_{ij} > R_\textrm{c},
598 \end{array}\right.
599 \end{equation}
600 is sufficient to smooth the potential (again, see figure
601 \ref{fig:ljCutoff}) and the forces by only perturbing the potential in
602 the switching region. If smooth second derivatives are required, a
603 higher order polynomial switching function (i.e. fifth order
604 polynomial) can be used.\cite{Andrea83,Leach01} It should be noted
605 that the higher the order of the polynomial, the more abrupt the
606 switching transition.
607
608 \subsection{\label{sec:LJCorrections}Long-Range Interaction Corrections}
609
610 While a good approximation, accumulating interactions only from nearby
611 particles can lead to some issues, because particles at distances
612 greater than $R_\textrm{c}$ do still have a small effect. For
613 instance, while the strength of the Lennard-Jones interaction is quite
614 weak at $R_\textrm{c}$ in figure \ref{fig:ljCutoff}, we are discarding
615 all of the attractive interactions that extend out to extremely long
616 distances. Thus, the potential is a little too high and the pressure
617 on the central particle from the surroundings is a little too low. For
618 homogeneous Lennard-Jones systems, we can correct for this effect by
619 assuming a uniform density and integrating the missing part,
620 \begin{equation}
621 V_\textrm{full}(r_{ij}) \approx V_\textrm{c}
622 + 2\pi N\rho\int^\infty_{R_\textrm{c}}r^2V_\textrm{LJ}(r)dr,
623 \end{equation}
624 where $V_\textrm{c}$ is the truncated Lennard-Jones
625 potential.\cite{Allen87} Like the potential, other properties can be
626 corrected by integration over the relevant function. Note that with
627 heterogeneous systems, this correction breaks down because the density
628 is no longer uniform.
629
630 Correcting long-range electrostatic interactions is a topic of great
631 importance in the field of molecular simulations. There have been
632 several techniques developed to address this issue, like the Ewald
633 summation and the reaction field technique. An in-depth analysis of
634 this problem, as well as useful corrective techniques, is presented in
635 chapter \ref{chap:electrostatics}.
636
637 \subsection{Periodic Boundary Conditions}
638
639 In typical molecular dynamics simulations there are no restrictions
640 placed on the motion of particles outside of what the inter-particle
641 interactions dictate. This means that if a particle collides with
642 other particles, it is free to move away from the site of the
643 collision. If we consider the entire system as a collection of
644 particles, they are not confined by walls of the ``simulation box''
645 and can freely move away from the other particles. With no boundary
646 considerations, particles moving outside of the simulation box
647 enter a vacuum. This is correct behavior for cluster simulations in a
648 vacuum; however, if we want to simulate bulk or spatially infinite
649 systems, we need to use periodic boundary conditions.
650
651 \begin{figure}
652 \centering
653 \includegraphics[width=4.5in]{./figures/periodicImage.pdf}
654 \caption{With periodic boundary conditions imposed, when particles
655 move out of one side the simulation box, they wrap back in the
656 opposite side. In this manner, a finite system of particles behaves as
657 an infinite system.}
658 \label{fig:periodicImage}
659 \end{figure}
660 In periodic boundary conditions, as a particle moves outside one wall
661 of the simulation box, the coordinates are remapped such that the
662 particle enters the opposing side of the box. This process is easy to
663 visualize in two dimensions as shown in figure \ref{fig:periodicImage}
664 and can occur in three dimensions, though it is not as easy to
665 visualize. Remapping the actual coordinates of the particles can be
666 problematic in that the we are restricting the distance a particle can
667 move from it's point of origin to a diagonal of the simulation
668 box. Thus, even though we are not confining the system with hard
669 walls, we are confining the particle coordinates to a particular
670 region in space. To avoid this ``soft'' confinement, it is common
671 practice to allow the particle coordinates to expand in an
672 unrestricted fashion while calculating interactions using a wrapped
673 set of ``minimum image'' coordinates. These coordinates need not be
674 stored because they are easily calculated while determining particle
675 distances.
676
677 \section{Calculating Properties}
678
679 In order to use simulations to model experimental processes and
680 evaluate theories, we need to be able to extract properties from the
681 results. In experiments, we can measure thermodynamic properties such
682 as the pressure and free energy. In computer simulations, we can
683 calculate properties from the motion and configuration of particles in
684 the system and make connections between these properties and the
685 experimental thermodynamic properties through statistical mechanics.
686
687 The work presented in the later chapters use the canonical ($NVT$),
688 isobaric-isothermal ($NPT$), and microcanonical ($NVE$) statistical
689 mechanical ensembles. The different ensembles lend themselves to more
690 effectively calculating specific properties. For instance, if we
691 concerned ourselves with the calculation of dynamic properties, which
692 are dependent upon the motion of the particles, it is better to choose
693 an ensemble that does not add artificial motions to keep properties
694 like the temperature or pressure constant. In this case, the $NVE$
695 ensemble would be the most appropriate choice. In chapter
696 \ref{chap:ice}, we discuss calculating free energies, which are
697 non-mechanical thermodynamic properties, and these calculations also
698 depend on the chosen ensemble.\cite{Allen87} The Helmholtz free energy
699 ($A$) depends on the $NVT$ partition function ($Q_{NVT}$),
700 \begin{equation}
701 A = -k_\textrm{B}T\ln Q_{NVT},
702 \end{equation}
703 while the Gibbs free energy ($G$) depends on the $NPT$ partition
704 function ($Q_{NPT}$),
705 \begin{equation}
706 G = -k_\textrm{B}T\ln Q_{NPT}.
707 \end{equation}
708 It is also useful to note that the conserved quantities of the $NVT$
709 and $NPT$ ensembles are related to the Helmholtz and Gibbs free
710 energies respectively.\cite{Melchionna93}
711
712 Integrating the equations of motion is a simple method to obtain a
713 sequence of configurations that sample the chosen ensemble. For each
714 of these configurations, we can calculate an instantaneous value for a
715 chosen property like the density in the $NPT$ ensemble, where the
716 volume is allowed to fluctuate. The density for the simulation is
717 calculated from an average over the densities for the individual
718 configurations. Since the configurations from the integration process
719 are related to one another by the time evolution of the interactions,
720 this average is technically a time average. In calculating
721 thermodynamic properties, we would actually prefer an ensemble average
722 that is representative of all accessible states of the system. We can
723 calculate thermodynamic properties from the time average by taking
724 advantage of the ergodic hypothesis, which states that for a
725 sufficiently chaotic system, and over a long enough period of time,
726 the time and ensemble averages are the same.
727
728 In addition to the average, the fluctuations of a particular property
729 can be determined via the standard deviation. Not only are
730 fluctuations useful for determining the spread of values around the
731 average and the error in the calculation of the value, they are also
732 useful for measuring various thermodynamic properties in computer
733 simulations. In section \ref{sec:t5peThermo}, we use fluctuations in
734 properties like the enthalpy and volume to calculate other
735 thermodynamic properties, such as the constant pressure heat capacity
736 and the isothermal compressibility.
737
738 \section{OOPSE}
739
740 In the following chapters, the above techniques are all utilized in
741 the study of molecular systems. There are a number of excellent
742 simulation packages available, both free and commercial, which
743 incorporate many of these
744 methods.\cite{Brooks83,MacKerell98,Pearlman95,Berendsen95,Lindahl01,Smith96,Ponder87}
745 Because of our interest in rigid body dynamics, point multipoles, and
746 systems where the orientational degrees of freedom cannot be handled
747 using the common constraint procedures (like {\sc shake}), the
748 majority of the following work was performed using {\sc oopse}, the
749 object-oriented parallel simulation engine.\cite{Meineke05} The {\sc
750 oopse} package started out as a collection of separate programs
751 written within our group, and has developed into one of the few
752 parallel molecular dynamics packages capable of accurately integrating
753 rigid bodies and point multipoles. This simulation package is
754 open-source software, available to anyone interested in performing
755 molecular dynamics simulations. More information about {\sc oopse} can
756 be found in reference \cite{Meineke05} or at the {\tt
757 http://oopse.org} website.
758
759