ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/chrisDissertation/Introduction.tex
Revision: 3001
Committed: Thu Sep 7 20:42:27 2006 UTC (18 years ago) by chrisfen
Content type: application/x-tex
File size: 26493 byte(s)
Log Message:
Thar be changes in these documents

File Contents

# User Rev Content
1 chrisfen 2987 \chapter{\label{chap:intro}INTRODUCTION AND BACKGROUND}
2    
3 chrisfen 3001 In order to advance our understanding of natural chemical and physical
4     processes, researchers develop explanations for events observed
5     experimentally. These hypotheses, supported by a body of corroborating
6     observations, can develop into accepted theories for how these
7     processes occur. This validation process, as well as testing the
8     limits of these theories, is essential in developing a firm foundation
9     for our knowledge. Theories involving molecular scale systems cause
10     particular difficulties in this process because their size and often
11     fast motions make them difficult to observe experimentally. One useful
12     tool for addressing these difficulties is computer simulation.
13    
14     In computer simulations, we can develop models from either the theory
15     or experimental knowledge and then test them in a controlled
16     environment. Work done in this manner allows us to further refine
17     theories, more accurately represent what is happening in experimental
18     observations, and even make predictions about what one will see in
19     experiments. Thus, computer simulations of molecular systems act as a
20     bridge between theory and experiment.
21    
22     Depending on the system of interest, there are a variety of different
23     computational techniques that can used to test and gather information
24     from the developed models. In the study of classical systems, the two
25     most commonly used techniques are Monte Carlo and molecular
26     dynamics. Both of these methods operate by calculating interactions
27     between particles of our model systems; however, the progression of
28     the simulation under the different techniques is vastly
29     different. Monte Carlo operates through random configuration changes
30     that that follow rules adhering to a specific statistical mechanics
31     ensemble, while molecular dynamics is chiefly concerned with solving
32     the classic equations of motion to move between configurations within
33     an ensemble. Thermodynamic properties can be calculated from both
34     techniques; but because of the random nature of Monte Carlo, only
35     molecular dynamics can be used to investigate dynamical
36     quantities. The research presented in the following chapters utilized
37     molecular dynamics near exclusively, so we will present a general
38     introduction to molecular dynamics and not Monte Carlo. There are
39     several resources available for those desiring a more in-depth
40     presentation of either of these
41     techniques.\cite{Allen87,Frenkel02,Leach01}
42    
43     \section{\label{sec:MolecularDynamics}Molecular Dynamics}
44    
45     \section{\label{sec:MovingParticles}Propagating Particle Motion}
46    
47     \subsection{\label{sec:IntroIntegrate}Rigid Body Motion}
48    
49     Rigid bodies are non-spherical particles or collections of particles
50     (e.g. $\mbox{C}_{60}$) that have a constant internal potential and
51     move collectively.\cite{Goldstein01} Discounting iterative constraint
52     procedures like {\sc shake} and {\sc rattle}, they are not included in
53     most simulation packages because of the algorithmic complexity
54     involved in propagating orientational degrees of
55     freedom.\cite{Ryckaert77,Andersen83,Krautler01} Integrators which
56     propagate orientational motion with an acceptable level of energy
57     conservation for molecular dynamics are relatively new inventions.
58    
59     Moving a rigid body involves determination of both the force and
60     torque applied by the surroundings, which directly affect the
61     translational and rotational motion in turn. In order to accumulate
62     the total force on a rigid body, the external forces and torques must
63     first be calculated for all the internal particles. The total force on
64     the rigid body is simply the sum of these external forces.
65     Accumulation of the total torque on the rigid body is more complex
66     than the force because the torque is applied to the center of mass of
67     the rigid body. The space-fixed torque on rigid body $i$ is
68     \begin{equation}
69     \boldsymbol{\tau}_i=
70     \sum_{a}\biggl[(\mathbf{r}_{ia}-\mathbf{r}_i)\times \mathbf{f}_{ia}
71     + \boldsymbol{\tau}_{ia}\biggr],
72     \label{eq:torqueAccumulate}
73     \end{equation}
74     where $\boldsymbol{\tau}_i$ and $\mathbf{r}_i$ are the torque on and
75     position of the center of mass respectively, while $\mathbf{f}_{ia}$,
76     $\mathbf{r}_{ia}$, and $\boldsymbol{\tau}_{ia}$ are the force on,
77     position of, and torque on the component particles.
78    
79     The summation of the total torque is done in the body fixed axis. In
80     order to move between the space fixed and body fixed coordinate axes,
81     parameters describing the orientation must be maintained for each
82     rigid body. At a minimum, the rotation matrix ($\mathsf{A}$) can be
83     described by the three Euler angles ($\phi, \theta,$ and $\psi$),
84     where the elements of $\mathsf{A}$ are composed of trigonometric
85     operations involving $\phi, \theta,$ and $\psi$.\cite{Goldstein01}
86     Direct propagation of the Euler angles has a known $1/\sin\theta$
87     divergence in the equations of motion for $\phi$ and $\psi$, leading
88     to numerical instabilities any time one of the directional atoms or
89     rigid bodies has an orientation near $\theta=0$ or
90     $\theta=\pi$.\cite{Allen87} One of the most practical ways to avoid
91     this ``gimbal point'' is to switch to another angular set defining the
92     orientation of the rigid body near this point.\cite{Barojas73} This
93     procedure results in additional book-keeping and increased algorithm
94     complexity. In the search for more elegant alternative methods, Evans
95     proposed the use of quaternions to describe and propagate
96     orientational motion.\cite{Evans77}
97    
98     The quaternion method for integration involves a four dimensional
99     representation of the orientation of a rigid
100     body.\cite{Evans77,Evans77b,Allen87} Thus, the elements of
101     $\mathsf{A}$ can be expressed as arithmetic operations involving the
102     four quaternions ($q_w, q_x, q_y,$ and $q_z$),
103     \begin{equation}
104     \mathsf{A} = \left( \begin{array}{l@{\quad}l@{\quad}l}
105     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) \\
106     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) \\
107     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 \\
108     \end{array}\right).
109     \end{equation}
110     Integration of the equations of motion involves a series of arithmetic
111     operations involving the quaternions and angular momenta and leads to
112     performance enhancements over Euler angles, particularly for very
113     small systems.\cite{Evans77} This integration methods works well for
114     propagating orientational motion in the canonical ensemble ($NVT$);
115     however, energy conservation concerns arise when using the simple
116     quaternion technique under the microcanonical ($NVE$) ensemble. An
117     earlier implementation of {\sc oopse} utilized quaternions for
118     propagation of rotational motion; however, a detailed investigation
119     showed that they resulted in a steady drift in the total energy,
120     something that has been observed by Kol {\it et al.} (also see
121     section~\ref{sec:waterSimMethods}).\cite{Kol97}
122    
123     Because of the outlined issues involving integration of the
124     orientational motion using both Euler angles and quaternions, we
125     decided to focus on a relatively new scheme that propagates the entire
126     nine parameter rotation matrix. This techniques is a velocity-Verlet
127     version of the symplectic splitting method proposed by Dullweber,
128     Leimkuhler and McLachlan ({\sc dlm}).\cite{Dullweber97} When there are
129     no directional atoms or rigid bodies present in the simulation, this
130     integrator becomes the standard velocity-Verlet integrator which is
131     known to effectively sample the microcanonical ($NVE$)
132     ensemble.\cite{Frenkel02}
133    
134     The key aspect of the integration method proposed by Dullweber
135     \emph{et al.} is that the entire $3 \times 3$ rotation matrix is
136     propagated from one time step to the next. In the past, this would not
137     have been as feasible, since the rotation matrix for a single body has
138     nine elements compared with the more memory-efficient methods (using
139     three Euler angles or four quaternions). Computer memory has become
140     much less costly in recent years, and this can be translated into
141     substantial benefits in energy conservation.
142    
143     The integration of the equations of motion is carried out in a
144     velocity-Verlet style two-part algorithm.\cite{Swope82} The first-part
145     ({\tt moveA}) consists of a half-step ($t + \delta t/2$) of the linear
146     velocity (${\bf v}$) and angular momenta ({\bf j}) and a full-step ($t
147     + \delta t$) of the positions ({\bf r}) and rotation matrix,
148     \begin{equation*}
149     {\tt moveA} = \left\{\begin{array}{r@{\quad\leftarrow\quad}l}
150     {\bf v}\left(t + \delta t / 2\right) & {\bf v}(t)
151     + \left( {\bf f}(t) / m \right)(\delta t/2), \\
152     %
153     {\bf r}(t + \delta t) & {\bf r}(t)
154     + {\bf v}\left(t + \delta t / 2 \right)\delta t, \\
155     %
156     {\bf j}\left(t + \delta t / 2 \right) & {\bf j}(t)
157     + \boldsymbol{\tau}^b(t)(\delta t/2), \\
158     %
159     \mathsf{A}(t + \delta t) & \mathrm{rotate}\left( {\bf j}
160     (t + \delta t / 2)\delta t \cdot
161     \overleftrightarrow{\mathsf{I}}^{-1} \right),
162     \end{array}\right.
163     \end{equation*}
164     where $\overleftrightarrow{\mathsf{I}}^{-1}$ is the inverse of the
165     moment of inertia tensor. The $\mathrm{rotate}$ function is the
166     product of rotations about the three body-fixed axes,
167     \begin{equation}
168     \mathrm{rotate}({\bf a}) = \mathsf{G}_x(a_x / 2) \cdot
169     \mathsf{G}_y(a_y / 2) \cdot \mathsf{G}_z(a_z) \cdot \mathsf{G}_y(a_y /
170     2) \cdot \mathsf{G}_x(a_x /2),
171     \label{eq:dlmTrot}
172     \end{equation}
173     where each rotational propagator, $\mathsf{G}_\alpha(\theta)$, rotates
174     both the rotation matrix ($\mathsf{A}$) and the body-fixed angular
175     momentum (${\bf j}$) by an angle $\theta$ around body-fixed axis
176     $\alpha$,
177     \begin{equation}
178     \mathsf{G}_\alpha( \theta ) = \left\{
179     \begin{array}{l@{\quad\leftarrow\quad}l}
180     \mathsf{A}(t) & \mathsf{A}(0) \cdot \mathsf{R}_\alpha(\theta)^\textrm{T},\\
181     {\bf j}(t) & \mathsf{R}_\alpha(\theta) \cdot {\bf j}(0).
182     \end{array}
183     \right.
184     \end{equation}
185     $\mathsf{R}_\alpha$ is a quadratic approximation to the single-axis
186     rotation matrix. For example, in the small-angle limit, the rotation
187     matrix around the body-fixed x-axis can be approximated as
188     \begin{equation}
189     \mathsf{R}_x(\theta) \approx \left(
190     \begin{array}{ccc}
191     1 & 0 & 0 \\
192     0 & \frac{1-\theta^2 / 4}{1 + \theta^2 / 4} & -\frac{\theta}{1+\theta^2 / 4} \\
193     0 & \frac{\theta}{1+\theta^2 / 4} & \frac{1-\theta^2 / 4}{1 + \theta^2 / 4}
194     \end{array}
195     \right).
196     \end{equation}
197     The remaining rotations follow in a straightforward manner. As seen
198     from the form of equation~(\ref{eq:dlmTrot}), the {\sc dlm} method
199     uses a Trotter factorization of the orientational
200     propagator.\cite{Trotter59} This has three effects:
201     \begin{enumerate}
202     \item the integrator is area-preserving in phase space (i.e. it is
203     {\it symplectic}),
204     \item the integrator is time-{\it reversible}, and
205     \item the error for a single time step is of order
206     $\mathcal{O}\left(\delta t^4\right)$ for time steps of length $\delta t$.
207     \end{enumerate}
208    
209     After the initial half-step ({\tt moveA}), the forces and torques are
210     evaluated for all of the particles. Once completed, the velocities can
211     be advanced to complete the second-half of the two-part algorithm
212     ({\tt moveB}), resulting an a completed full step of both the
213     positions and momenta,
214     \begin{equation*}
215     {\tt moveB} = \left\{\begin{array}{r@{\quad\leftarrow\quad}l}
216     {\bf v}\left(t + \delta t \right) &
217     {\bf v}\left(t + \delta t / 2 \right)
218     + \left({\bf f}(t + \delta t) / m \right)(\delta t/2), \\
219     %
220     {\bf j}\left(t + \delta t \right) &
221     {\bf j}\left(t + \delta t / 2 \right)
222     + \boldsymbol{\tau}^b(t + \delta t)(\delta t/2).
223     \end{array}\right.
224     \end{equation*}
225    
226     The matrix rotations used in the {\sc dlm} method end up being more
227     costly computationally than the simpler arithmetic quaternion
228     propagation. With the same time step, a 1024-molecule water simulation
229     incurs approximately a 10\% increase in computation time using the
230     {\sc dlm} method in place of quaternions. This cost is more than
231     justified when comparing the energy conservation achieved by the two
232     methods. Figure \ref{fig:quatdlm} provides a comparative analysis of
233     the {\sc dlm} method versus the traditional quaternion scheme.
234    
235     \begin{figure}
236     \centering
237     \includegraphics[width=3.5in]{./figures/dlmVsQuat.pdf}
238     \caption[Energy conservation analysis of the {\sc dlm} and quaternion
239     integration methods]{Analysis of the energy conservation of the {\sc
240     dlm} and quaternion integration methods. $\delta \mathrm{E}_1$ is the
241     linear drift in energy over time and $\delta \mathrm{E}_0$ is the
242     standard deviation of energy fluctuations around this drift. All
243     simulations were of a 1024 SSD water system at 298 K starting from the
244     same initial configuration. Note that the {\sc dlm} method provides
245     more than an order-of-magnitude improvement in both the energy drift
246     and the size of the energy fluctuations when compared with the
247     quaternion method at any given time step. At time steps larger than 4
248     fs, the quaternion scheme resulted in rapidly rising energies which
249     eventually lead to simulation failure. Using the {\sc dlm} method,
250     time steps up to 8 fs can be taken before this behavior is evident.}
251     \label{fig:quatdlm}
252     \end{figure}
253    
254     In figure \ref{fig:quatdlm}, $\delta \mbox{E}_1$ is a measure of the
255     linear energy drift in units of $\mbox{kcal mol}^{-1}$ per particle
256     over a nanosecond of simulation time, and $\delta \mbox{E}_0$ is the
257     standard deviation of the energy fluctuations in units of $\mbox{kcal
258     mol}^{-1}$ per particle. In the top plot, it is apparent that the
259     energy drift is reduced by a significant amount (2 to 3 orders of
260     magnitude improvement at all tested time steps) by choosing the {\sc
261     dlm} method over the simple non-symplectic quaternion integration
262     method. In addition to this improvement in energy drift, the
263     fluctuations in the total energy are also dampened by 1 to 2 orders of
264     magnitude by utilizing the {\sc dlm} method.
265    
266     \begin{figure}
267     \centering
268     \includegraphics[width=\linewidth]{./figures/compCost.pdf}
269     \caption[Energy drift as a function of required simulation run
270     time]{Energy drift as a function of required simulation run time.
271     $\delta \mathrm{E}_1$ is the linear drift in energy over time.
272     Simulations were performed on a single 2.5 GHz Pentium 4
273     processor. Simulation time comparisons can be made by tracing
274     horizontally from one curve to the other. For example, a simulation
275     that takes 24 hours using the {\sc dlm} method will take roughly
276     210 hours using the simple quaternion method if the same degree of
277     energy conservation is desired.}
278     \label{fig:cpuCost}
279     \end{figure}
280     Although the {\sc dlm} method is more computationally expensive than
281     the traditional quaternion scheme for propagating of a time step,
282     consideration of the computational cost for a long simulation with a
283     particular level of energy conservation is in order. A plot of energy
284     drift versus computational cost was generated
285     (Fig.~\ref{fig:cpuCost}). This figure provides an estimate of the CPU
286     time required under the two integration schemes for 1 nanosecond of
287     simulation time for the model 1024-molecule system. By choosing a
288     desired energy drift value it is possible to determine the CPU time
289     required for both methods. If a $\delta \mbox{E}_1$ of
290     0.001~kcal~mol$^{-1}$ per particle is desired, a nanosecond of
291     simulation time will require ~19 hours of CPU time with the {\sc dlm}
292     integrator, while the quaternion scheme will require ~154 hours of CPU
293     time. This demonstrates the computational advantage of the integration
294     scheme utilized in {\sc oopse}.
295    
296     \subsection{Periodic Boundary Conditions}
297    
298     \begin{figure}
299     \centering
300     \includegraphics[width=4.5in]{./figures/periodicImage.pdf}
301     \caption{With periodic boundary conditions imposed, when particles
302     move out of one side the simulation box, they wrap back in the
303     opposite side. In this manner, a finite system of particles behaves as
304     an infinite system.}
305     \label{fig:periodicImage}
306     \end{figure}
307    
308     \section{Accumulating Interactions}
309    
310     In the force calculation between {\tt moveA} and {\tt moveB} mentioned
311     in section \ref{sec:IntroIntegrate}, we need to accumulate the
312     potential and forces (and torques if the particle is a rigid body or
313     multipole) on each particle from their surroundings. This can quickly
314     become a cumbersome task for large systems since the number of pair
315     interactions increases by $\mathcal{O}(N(N-1)/2)$ if you accumulate
316     interactions between all particles in the system, note the utilization
317     of Newton's third law to reduce the interaction number from
318     $\mathcal{O}(N^2)$. The case of periodic boundary conditions further
319     complicates matters by turning the finite system into an infinitely
320     repeating one. Fortunately, we can reduce the scale of this problem by
321     using spherical cutoff methods.
322    
323     \begin{figure}
324     \centering
325     \includegraphics[width=3.5in]{./figures/sphericalCut.pdf}
326     \caption{When using a spherical cutoff, only particles within a chosen
327     cutoff radius distance, $R_\textrm{c}$, of the central particle are
328     included in the pairwise summation. This reduces a problem that scales
329     by $\sim\mathcal{O}(N^2)$ to one that scales by $\sim\mathcal{O}(N)$.}
330     \label{fig:sphereCut}
331     \end{figure}
332     With spherical cutoffs, rather than accumulating the full set of
333     interactions between all particles in the simulation, we only
334     explicitly consider interactions between local particles out to a
335     specified cutoff radius distance, $R_\textrm{c}$, (see figure
336     \ref{fig:sphereCut}). This reduces the scaling of the interaction to
337     $\mathcal{O}(N\cdot\textrm{c})$, where `c' is a value that depends on
338     the size of $R_\textrm{c}$ (c $\approx R_\textrm{c}^3$). Determination
339     of which particles are within the cutoff is also an issue, because
340     this process requires a full loop over all $N(N-1)/2$ pairs. To reduce
341     the this expense, we can use neighbor lists.\cite{Verlet67,Thompson83}
342     With neighbor lists, we have a second list of particles built from a
343     list radius $R_\textrm{l}$, which is larger than $R_\textrm{c}$. Once
344     any of the particles within $R_\textrm{l}$ move a distance of
345     $R_\textrm{l}-R_\textrm{c}$ (the ``skin'' thickness), we rebuild the
346     list with the full $N(N-1)/2$ loop.\cite{Verlet67} With an appropriate
347     skin thickness, these updates are only performed every $\sim$20
348     time steps, significantly reducing the time spent on pair-list
349     bookkeeping operations.\cite{Allen87} If these neighbor lists are
350     utilized, it is important that these list updates occur
351     regularly. Incorrect application of this technique leads to
352     non-physical dynamics, such as the ``flying block of ice'' behavior
353     for which improper neighbor list handling was identified a one of the
354     possible causes.\cite{Harvey98,Sagui99}
355    
356     \subsection{Correcting Cutoff Discontinuities}
357     \begin{figure}
358     \centering
359     \includegraphics[width=3.5in]{./figures/ljCutoffSquare.pdf}
360     \caption{The common methods to smooth the potential discontinuity
361     introduced when using a cutoff include a shifted potential, a shifted
362     force, and a switching function. The shifted potential and shifted
363     force both lift the whole potential so that it zeroes at
364     $R_\textrm{c}$, thereby reducing the strength of the interaction. The
365     (cubic) switching function only alters the potential in the switching
366     region in order to smooth out the discontinuity.}
367     \label{fig:ljCutoff}
368     \end{figure}
369     As particle move in and out of $R_\textrm{c}$, there will be sudden
370     discontinuous jumps in the potential (and forces) due to their
371     appearance and disappearance. In order to prevent heating and poor
372     energy conservation in the simulations, this discontinuity needs to be
373     smoothed out. There are several ways to modify the function so that it
374     crosses $R_\textrm{c}$ in a continuous fashion, and the easiest
375     methods is shifting the potential. To shift the potential, we simply
376     subtract out the value we calculate at $R_\textrm{c}$ from the whole
377     potential. For the shifted form of the Lennard-Jones potential is
378     \begin{equation}
379     V_\textrm{SLJ} = \left\{\begin{array}{l@{\quad\quad}l}
380     V_\textrm{LJ}(r_{ij}) - V_\textrm{LJ}(R_\textrm{c}) & r_{ij} < R_\textrm{c}, \\
381     0 & r_{ij} \geqslant R_\textrm{c},
382     \end{array}\right.
383     \end{equation}
384     where
385     \begin{equation}
386     V_\textrm{LJ} = 4\epsilon\left[\left(\frac{\sigma}{r_{ij}}\right)^{12} -
387     \left(\frac{\sigma}{r_{ij}}\right)^6\right].
388     \end{equation}
389     In figure \ref{fig:ljCutoff}, the shifted form of the potential
390     reaches zero at the cutoff radius at the expense of the correct
391     magnitude below $R_\textrm{c}$. This correction method also does
392     nothing to correct the discontinuity in the forces. We can account for
393     this force discontinuity by shifting in the by applying the shifting
394     in the forces rather than just the potential via
395     \begin{equation}
396     V_\textrm{SFLJ} = \left\{\begin{array}{l@{\quad\quad}l}
397     V_\textrm{LJ}({r_{ij}}) - V_\textrm{LJ}(R_\textrm{c}) -
398     \left(\frac{d V(r_{ij})}{d r_{ij}}\right)_{r_{ij}=R_\textrm{c}}
399     (r_{ij} - R_\textrm{c}) & r_{ij} < R_\textrm{c}, \\
400     0 & r_{ij} \geqslant R_\textrm{c}.
401     \end{array}\right.
402     \end{equation}
403     The forces are continuous with this potential; however, the inclusion
404     of the derivative term distorts the potential even further than the
405     simple shifting as shown in figure \ref{fig:ljCutoff}. The method for
406     correcting the discontinuity which results in the smallest
407     perturbation in both the potential and the forces is the use of a
408     switching function. The cubic switching function,
409     \begin{equation}
410     S(r) = \left\{\begin{array}{l@{\quad\quad}l}
411     1 & r_{ij} \leqslant R_\textrm{sw}, \\
412     \frac{(R_\textrm{c} + 2r_{ij} - 3R_\textrm{sw})
413     (R_\textrm{c} - r_{ij})^2}{(R_\textrm{c} - R_\textrm{sw})^3}
414     & R_\textrm{sw} < r_{ij} \leqslant R_\textrm{c}, \\
415     0 & r_{ij} > R_\textrm{c},
416     \end{array}\right.
417     \end{equation}
418     is sufficient to smooth the potential (again, see figure
419     \ref{fig:ljCutoff}) and the forces by only perturbing the potential in
420     the switching region. If smooth second derivatives are required, a
421     higher order polynomial switching function (i.e. fifth order
422     polynomial) can be used.\cite{Andrea83,Leach01} It should be noted
423     that the higher the order of the polynomial, the more abrupt the
424     switching transition.
425    
426     \subsection{Long-Range Interaction Corrections}
427    
428     While a good approximation, accumulating interaction only from the
429     nearby particles can lead to some issues, because the further away
430     surrounding particles do still have a small effect. For instance,
431     while the strength of the Lennard-Jones interaction is quite weak at
432     $R_\textrm{c}$ in figure \ref{fig:ljCutoff}, we are discarding all of
433     the attractive interaction that extends out to extremely long
434     distances. Thus, the potential is a little too high and the pressure
435     on the central particle from the surroundings is a little too low. For
436     homogeneous Lennard-Jones systems, we can correct for this neglect by
437     assuming a uniform density and integrating the missing part,
438     \begin{equation}
439     V_\textrm{full}(r_{ij}) \approx V_\textrm{c}
440     + 2\pi N\rho\int^\infty_{R_\textrm{c}}r^2V_\textrm{LJ}(r)dr,
441     \end{equation}
442     where $V_\textrm{c}$ is the truncated Lennard-Jones
443     potential.\cite{Allen87} Like the potential, other Lennard-Jones
444     properties can be corrected by integration over the relevant
445     function. Note that with heterogeneous systems, this correction begins
446     to break down because the density is no longer uniform.
447    
448     Correcting long-range electrostatic interactions is a topic of great
449     importance in the field of molecular simulations. There have been
450     several techniques developed to address this issue, like the Ewald
451     summation and the reaction field technique. An in-depth analysis of
452     this problem, as well as useful corrective techniques, is presented in
453     chapter \ref{chap:electrostatics}.
454    
455     \section{Measuring Properties}
456    
457     \section{Application of the Techniques}
458    
459     In the following chapters, the above techniques are all utilized in
460     the study of molecular systems. There are a number of excellent
461     simulation packages available, both free and commercial, which
462     incorporate many of these
463     methods.\cite{Brooks83,MacKerell98,Pearlman95,Berendsen95,Lindahl01,Smith96,Ponder87}
464     Because of our interest in rigid body dynamics, point multipoles, and
465     systems where the orientational degrees cannot be handled using the
466     common constraint procedures (like {\sc shake}), the majority of the
467     following work was performed using {\sc oopse}, the object-oriented
468     parallel simulation engine.\cite{Meineke05} The {\sc oopse} package
469     started out as a collection of separate programs written within our
470     group, and has developed into one of the few parallel molecular
471     dynamics packages capable of accurately integrating rigid bodies and
472     point multipoles.
473    
474     In chapter \ref{chap:electrostatics}, we investigate correction
475     techniques for proper handling of long-ranged electrostatic
476     interactions. In particular we develop and evaluate some new pairwise
477     methods which we have incorporated into {\sc oopse}. These techniques
478     make an appearance in the later chapters, as they are applied to
479     specific systems of interest, showing how it they can improve the
480     quality of various molecular simulations.
481    
482     In chapter \ref{chap:water}, we focus on simple water models,
483     specifically the single-point soft sticky dipole (SSD) family of water
484     models. These single-point models are more efficient than the common
485     multi-point partial charge models and, in many cases, better capture
486     the dynamic properties of water. We discuss improvements to these
487     models in regards to long-range electrostatic corrections and show
488     that these models can work well with the techniques discussed in
489     chapter \ref{chap:electrostatics}. By investigating and improving
490     simple water models, we are extending the limits of computational
491     efficiency for systems that heavily depend on water calculations.
492    
493     In chapter \ref{chap:ice}, we study a unique polymorph of ice that we
494     discovered while performing water simulations with the fast simple
495     water models discussed in the previous chapter. This form of ice,
496     which we called ``imaginary ice'' (Ice-$i$), has a low-density
497     structure which is different from any known polymorph from either
498     experiment or other simulations. In this study, we perform a free
499     energy analysis and see that this structure is in fact the
500     thermodynamically preferred form of ice for both the single-point and
501     commonly used multi-point water models under the chosen simulation
502     conditions. We also consider electrostatic corrections, again
503     including the techniques discussed in chapter
504     \ref{chap:electrostatics}, to see how they influence the free energy
505     results. This work, to some degree, addresses the appropriateness of
506     using these simplistic water models outside of the conditions for
507     which they were developed.
508    
509     In the final chapter we summarize the work presented previously. We
510     also close with some final comments before the supporting information
511     presented in the appendices.