ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/chrisDissertation/Ice.tex
Revision: 3045
Committed: Fri Oct 13 21:03:14 2006 UTC (17 years, 10 months ago) by chrisfen
Content type: application/x-tex
File size: 30997 byte(s)
Log Message:
fixin figures and tables of contents

File Contents

# User Rev Content
1 chrisfen 3045 \chapter[PHASE BEHAVIOR OF WATER IN COMPUTER \\ SIMULATIONS]{\label{chap:ice}PHASE BEHAVIOR OF WATER IN COMPUTER SIMULATIONS}
2 chrisfen 2987
3 chrisfen 3028 As mentioned in the previous chapter, water has proven to be a
4 chrisfen 2987 challenging substance to depict in simulations, and a variety of
5     models have been developed to describe its behavior under varying
6     simulation
7     conditions.\cite{Stillinger74,Rahman75,Berendsen81,Jorgensen83,Bratko85,Berendsen87,Caldwell95,Liu96,vanderSpoel98,Urbic00,Mahoney00,Fennell04}
8     These models have been used to investigate important physical
9     phenomena like phase transitions and the hydrophobic
10     effect.\cite{Yamada02,Marrink94,Gallagher03} With the choice of models
11     available, it is only natural to compare them under interesting
12     thermodynamic conditions in an attempt to clarify the limitations of
13     each.\cite{Jorgensen83,Jorgensen98b,Baez94,Mahoney01} Two important
14 chrisfen 3028 properties to quantify are the Gibbs and Helmholtz free energies,
15 chrisfen 2987 particularly for the solid forms of water, as these predict the
16     thermodynamic stability of the various phases. Water has a
17 chrisfen 3028 particularly rich phase diagram and takes on a number of different
18 chrisfen 2987 stable crystalline structures as the temperature and pressure are
19     varied. This complexity makes it a challenging task to investigate the
20     entire free energy landscape.\cite{Sanz04} Ideally, research is
21     focused on the phases having the lowest free energy at a given state
22     point, because these phases will dictate the relevant transition
23     temperatures and pressures for the model.
24    
25 chrisfen 3028 The high-pressure phases of water (ice II-ice X as well as ice XII-ice XIV)
26 chrisfen 2987 have been studied extensively both experimentally and
27     computationally. In this chapter, standard reference state methods
28     were applied in the {\it low} pressure regime to evaluate the free
29     energies for a few known crystalline water polymorphs that might be
30     stable at these pressures. This work is unique in the fact that one of
31     the crystal lattices was arrived at through crystallization of a
32     computationally efficient water model under constant pressure and
33     temperature conditions.
34    
35     While performing a series of melting simulations on an early iteration
36     of SSD/E, we observed several recrystallization events at a constant
37     pressure of 1 atm. After melting from ice I$_\textrm{h}$ at 235~K, two
38     of five systems recrystallized near 245~K. Crystallization events are
39     interesting in and of themselves;\cite{Matsumoto02,Yamada02} however,
40     the crystal structure extracted from these systems is different from
41     any previously observed ice polymorphs in experiment or
42 chrisfen 3028 simulation.\cite{Fennell04} We have named this structure ``imaginary
43     ice'' (Ice-$i$) to indicate its origin in computer simulations. The
44     unit cell of Ice-$i$ and an axially elongated variant named
45     Ice-$i^\prime$ both consist of eight water molecules that stack in
46     rows of interlocking water tetramers as illustrated in figure
47     \ref{fig:iceiCell}. These tetramers form a crystal structure
48     similar in appearance to a recently simulated two-dimensional surface
49     tessellation of water on silica.\cite{Yang04} As expected in an ice
50     crystal constructed of water tetramers, the hydrogen bonds are not as
51     linear as those present in ice I$_\textrm{h}$; however, the
52     interlocking of these subunits appears to provide significant
53     stabilization to the overall crystal. The arrangement of these
54     tetramers results in open octagonal cavities that are typically
55     greater than 6.3~\AA\ in diameter (see figure
56 chrisfen 2987 \ref{fig:protOrder}). This open structure leads to crystals that are
57     typically 0.07~g/cm$^3$ less dense than ice I$_\textrm{h}$.
58    
59     \begin{figure}
60     \includegraphics[width=\linewidth]{./figures/unitCell.pdf}
61 chrisfen 3028 \caption{Unit cells for (A) Ice-$i$ and (B) Ice-$i^\prime$, the
62     elongated variant of Ice-$i$. For Ice-$i$, the $a$ to $c$
63 chrisfen 2987 relation is given by $a = 2.1214c$, while for Ice-$i^\prime$, $a =
64     1.7850c$.}
65     \label{fig:iceiCell}
66     \end{figure}
67    
68     \begin{figure}
69     \centering
70     \includegraphics[width=3.5in]{./figures/orderedIcei.pdf}
71 chrisfen 3028 \caption{Image of a proton ordered crystal of Ice-$i$ looking
72 chrisfen 2987 down the (001) crystal face. The rows of water tetramers surrounded by
73     octagonal pores leads to a crystal structure that is significantly
74     less dense than ice I$_\textrm{h}$.}
75     \label{fig:protOrder}
76     \end{figure}
77    
78 chrisfen 3028 Results from our initial studies indicated that Ice-$i$ is the
79 chrisfen 2987 minimum energy crystal structure for the single point water models
80     investigated (for discussions on these single point dipole models, see
81 chrisfen 3028 the previous chapter and related
82 chrisfen 2987 articles\cite{Fennell04,Liu96,Bratko85}). These earlier results only
83     considered energetic stabilization and neglected entropic
84     contributions to the overall free energy. To address this issue, we
85     have calculated the absolute free energy of this crystal using
86 chrisfen 3042 thermodynamic integration and compared it to the free energies of ice
87 chrisfen 2987 I$_\textrm{c}$ and ice I$_\textrm{h}$ (the common low-density ice
88     polymorphs) and ice B (a higher density, but very stable crystal
89     structure observed by B\'{a}ez and Clancy in free energy studies of
90     SPC/E).\cite{Baez95b} This work includes results for the water model
91 chrisfen 3028 from which Ice-$i$ was crystallized (SSD/E) in addition to several
92 chrisfen 2987 common water models (TIP3P, TIP4P, TIP5P, and SPC/E) and a reaction
93     field parametrized single point dipole water model (SSD/RF). The
94     axially elongated variant, Ice-$i^\prime$, was used in calculations
95     involving SPC/E, TIP4P, and TIP5P. The square tetramer in Ice-$i$
96     distorts in Ice-$i^\prime$ to form a rhombus with alternating 85 and
97     95 degree angles. Under SPC/E, TIP4P, and TIP5P, this geometry is
98     better at forming favorable hydrogen bonds. The degree of rhomboid
99     distortion depends on the water model used but is significant enough
100 chrisfen 3028 to split the peak in the radial distribution function which
101     corresponds to diagonal sites in the tetramers.
102 chrisfen 2987
103     \section{Methods and Thermodynamic Integration}
104    
105 chrisfen 3028 Canonical ensemble ($NVT$) molecular dynamics calculations were
106     performed using the {\sc oopse} molecular mechanics
107     package.\cite{Meineke05} The densities chosen for the simulations were
108     taken from isobaric-isothermal ($NPT$) simulations performed at 1~atm
109     and 200~K. Each model (and each crystal structure) was allowed to
110     relax for 300~ps in the $NPT$ ensemble before averaging the density to
111     obtain the volumes for the $NVT$ simulations. All molecules were
112     treated as rigid bodies, with orientational motion propagated using
113     the symplectic {\sc dlm} integration method described in section
114 chrisfen 2987 \ref{sec:IntroIntegrate}.
115    
116    
117     We used thermodynamic integration to calculate the Helmholtz free
118     energies ({\it A}) of the listed water models at various state
119     points. Thermodynamic integration is an established technique that has
120     been used extensively in the calculation of free energies for
121     condensed phases of
122     materials.\cite{Frenkel84,Hermens88,Meijer90,Baez95a,Vlot99}. This
123 chrisfen 3023 method uses a sequence of simulations during which the system of
124 chrisfen 2987 interest is converted into a reference system for which the free
125     energy is known analytically ($A_0$). The difference in potential
126     energy between the reference system and the system of interest
127     ($\Delta V$) is then integrated in order to determine the free energy
128     difference between the two states:
129     \begin{equation}
130     A = A_0 + \int_0^1 \left\langle \Delta V \right\rangle_\lambda d\lambda.
131     \end{equation}
132     Here, $\lambda$ is the parameter that governs the transformation
133     between the reference system and the system of interest. For
134     crystalline phases, an harmonically-restrained (Einstein) crystal is
135     chosen as the reference state, while for liquid phases, the ideal gas
136     is taken as the reference state. Figure \ref{fig:integrationPath}
137     shows an example integration path for converting a crystalline system
138     to the Einstein crystal reference state.
139     \begin{figure}
140     \includegraphics[width=\linewidth]{./figures/integrationPath.pdf}
141     \caption{An example integration path to convert an unrestrained
142     crystal ($\lambda = 1$) to the Einstein crystal reference state
143     ($\lambda = 0$). Note the increase in samples at either end of the
144     path to improve the smoothness of the curve. For reversible processes,
145     conversion of the Einstein crystal back to the system of interest will
146     give an identical plot, thereby integrating to the same result.}
147     \label{fig:integrationPath}
148     \end{figure}
149    
150     In an Einstein crystal, the molecules are restrained at their ideal
151     lattice locations and orientations. Using harmonic restraints, as
152     applied by B\'{a}ez and Clancy, the total potential for this reference
153     crystal ($V_\mathrm{EC}$) is the sum of all the harmonic restraints,
154     \begin{equation}
155     V_\mathrm{EC} = \frac{K_\mathrm{v}r^2}{2} + \frac{K_\theta\theta^2}{2} +
156     \frac{K_\omega\omega^2}{2},
157     \end{equation}
158     where $K_\mathrm{v}$, $K_\mathrm{\theta}$, and $K_\mathrm{\omega}$ are
159     the spring constants restraining translational motion and deflection
160 chrisfen 3028 of and rotation around the principle axis of the molecule respectively
161     (see figure \ref{fig:waterSpring}). These spring constants are
162     typically calculated from the mean-square displacements of water
163     molecules in an unrestrained ice crystal at 200~K. For these studies,
164     $K_\mathrm{v} = 4.29$~kcal~mol$^{-1}$~\AA$^{-2}$, $K_\theta\ =
165 chrisfen 2987 13.88$~kcal~mol$^{-1}$~rad$^{-2}$, and $K_\omega\ =
166 chrisfen 3028 17.75$~kcal~mol$^{-1}$~rad$^{-2}$. It is clear from figure
167     \ref{fig:waterSpring} that the values of $\theta$ range from $0$ to
168     $\pi$, while $\omega$ ranges from $-\pi$ to $\pi$. The partition
169 chrisfen 2987 function for a molecular crystal restrained in this fashion can be
170     evaluated analytically, and the Helmholtz Free Energy ({\it A}) is
171     given by
172     \begin{equation}
173     \begin{split}
174     A = E_m &- kT\ln\left(\frac{kT}{h\nu}\right)^3 \\
175     &- kT\ln\left[\pi^\frac{1}{2}\left(
176     \frac{8\pi^2I_\mathrm{A}kT}{h^2}\right)^\frac{1}{2}
177     \left(\frac{8\pi^2I_\mathrm{B}kT}{h^2}\right)^\frac{1}{2}
178     \left(\frac{8\pi^2I_\mathrm{C}kT}{h^2}\right)^\frac{1}{2}
179     \right] \\
180     &- kT\ln\left[\frac{kT}{2(\pi K_\omega K_\theta)^{\frac{1}{2}}}
181     \exp\left(-\frac{kT}{2K_\theta}\right)
182     \int_0^{\left(\frac{kT}{2K_\theta}\right)^\frac{1}{2}}
183     \exp(t^2)\mathrm{d}t\right],
184     \end{split}
185     \label{eq:ecFreeEnergy}
186     \end{equation}
187     where $2\pi\nu = (K_\mathrm{v}/m)^{1/2}$, and $E_m$ is the minimum
188 chrisfen 3028 potential energy of the ideal crystal.\cite{Baez95a}
189    
190     The choice of an Einstein crystal reference state is somewhat
191     arbitrary. Any ideal system for which the partition function is known
192     exactly could be used as a reference point, as long as the system does
193     not undergo a phase transition during the integration path between the
194     real and ideal systems. Nada and van der Eerden have shown that the
195     use of different force constants in the Einstein crystal does not
196     affect the total free energy, and Gao {\it et al.} have shown that
197     free energies computed with the Debye crystal reference state differ
198     from the Einstein crystal by only a few tenths of a
199 chrisfen 2987 kJ~mol$^{-1}$.\cite{Nada03,Gao00} These free energy differences can
200     lead to some uncertainty in the computed melting point of the solids.
201     \begin{figure}
202     \centering
203     \includegraphics[width=3.5in]{./figures/rotSpring.pdf}
204     \caption{Possible orientational motions for a restrained molecule.
205     $\theta$ angles correspond to displacement from the body-frame {\it
206     z}-axis, while $\omega$ angles correspond to rotation about the
207     body-frame {\it z}-axis. $K_\theta$ and $K_\omega$ are spring
208     constants for the harmonic springs restraining motion in the $\theta$
209     and $\omega$ directions.}
210     \label{fig:waterSpring}
211     \end{figure}
212    
213     In the case of molecular liquids, the ideal vapor is chosen as the
214     target reference state. There are several examples of liquid state
215     free energy calculations of water models present in the
216     literature.\cite{Hermens88,Quintana92,Mezei92,Baez95b} These methods
217     typically differ in regard to the path taken for switching off the
218     interaction potential to convert the system to an ideal gas of water
219     molecules. In this study, we applied one of the most convenient
220     methods and integrated over the $\lambda^4$ path, where all
221     interaction parameters are scaled equally by this transformation
222 chrisfen 3028 parameter. This method is reversible and provides results in excellent
223     agreement with other established methods.\cite{Baez95b}
224 chrisfen 2987
225 chrisfen 3004 The Helmholtz free energy error was determined in the same manner in
226 chrisfen 3028 both the solid and the liquid free energy calculations. At each point
227 chrisfen 3004 along the integration path, we calculated the standard deviation of
228     the potential energy difference. Addition or subtraction of these
229 chrisfen 3028 deviations to each of their respective points and integrating the
230     curve again provides the upper and lower bounds of the uncertainty in
231     the Helmholtz free energy.
232 chrisfen 3004
233 chrisfen 2987 Near the cutoff radius ($0.85 * r_{cut}$), charge, dipole, and
234     Lennard-Jones interactions were gradually reduced by a cubic switching
235     function. By applying this function, these interactions are smoothly
236     truncated, thereby avoiding the poor energy conservation which results
237     from harsher truncation schemes. The effect of a long-range
238     correction was also investigated on select model systems in a variety
239     of manners. For the SSD/RF model, a reaction field with a fixed
240     dielectric constant of 80 was applied in all
241     simulations.\cite{Onsager36} For a series of the least computationally
242     expensive models (SSD/E, SSD/RF, TIP3P, and SPC/E), simulations were
243     performed with longer cutoffs of 10.5, 12, 13.5, and 15~\AA\ to
244     compare with the 9~\AA\ cutoff results. Finally, the effects of using
245     the Ewald summation were estimated for TIP3P and SPC/E by performing
246 chrisfen 3028 single configuration smooth particle-mesh Ewald ({\sc spme})
247     calculations for each of the ice polymorphs.\cite{Ponder87} The
248     calculated energy difference in the presence and absence of {\sc spme}
249     was applied to the previous results in order to predict changes to the
250     free energy landscape.
251 chrisfen 2987
252 chrisfen 3019 In addition to the above procedures, we also tested how the inclusion
253     of the Lennard-Jones long-range correction affects the free energy
254 chrisfen 3028 results. The correction for the Lennard-Jones truncation was included
255 chrisfen 3019 by integration of the equation discussed in section
256     \ref{sec:LJCorrections}. Rather than discuss its affect alongside the
257     free energy results, we will just mention that while the correction
258     does lower the free energy of the higher density states more than the
259     lower density states, the effect is so small that it is entirely
260 chrisfen 3028 overwhelmed by the error in the free energy calculation. Since its
261 chrisfen 3019 inclusion does not influence the results, the Lennard-Jones correction
262     was omitted from all the calculations below.
263    
264 chrisfen 2987 \section{Initial Free Energy Results}
265    
266     \begin{table}
267     \centering
268     \caption{HELMHOLTZ FREE ENERGIES AND TRANSITION TEMPERATURES AT 1
269     ATMOSPHERE FOR SEVERAL WATER MODELS}
270    
271     \footnotesize
272     \begin{tabular}{lccccccc}
273     \toprule
274     \toprule
275 chrisfen 3028 Water Model & I$_\textrm{h}$ & I$_\textrm{c}$ & B & Ice-$i$ & Ice-$i^\prime$ & $T_\textrm{m}$ (*$T_\textrm{s}$) & $T_\textrm{b}$\\
276 chrisfen 2987 \cmidrule(lr){2-6}
277     \cmidrule(l){7-8}
278     & \multicolumn{5}{c}{(kcal mol$^{-1}$)} & \multicolumn{2}{c}{(K)}\\
279     \midrule
280     TIP5P & -11.85(3) & -11.86(2) & -11.96(2) & - & -12.29(2) & 266(7) & 337(4)\\
281     TIP4P & -11.84(3) & -12.04(2) & -12.08(3) & - & -12.33(3) & 262(6) & 354(4)\\
282     TIP3P & -11.41(2) & -11.23(3) & -11.82(3) & -12.30(3) & - & 269(7) & 357(4)\\
283     SPC/E & -12.87(2) & -13.05(2) & -13.26(3) & - & -13.55(2) & 299(6) & 396(4)\\
284     SSD/E & -11.27(2) & -11.19(4) & -12.09(2) & -12.54(2) & - & *355(4) & -\\
285     SSD/RF & -11.96(2) & -11.60(2) & -12.53(3) & -12.79(2) & - & 278(7) & 382(4)\\
286     \bottomrule
287     \end{tabular}
288     \label{tab:freeEnergy}
289     \end{table}
290 chrisfen 3028 The calculated free energies of proton-ordered variants of three low
291     density polymorphs (I$_\textrm{h}$, I$_\textrm{c}$, and Ice-$i$ or
292     Ice-$i^\prime$) and the stable higher density ice B are listed in
293     table \ref{tab:freeEnergy}. Ice B was included because it has been
294     shown to be a minimum free energy structure for SPC/E at ambient
295     conditions.\cite{Baez95b} In addition to the free energies, the
296     relevant transition temperatures at standard pressure are also
297     displayed in table \ref{tab:freeEnergy}. These free energy values
298     indicate that Ice-$i$ is the thermodynamically preferred state for
299     all of the investigated water models under the chosen simulation
300     conditions. With the free energy at these state points, the
301     Gibbs-Helmholtz equation was used to project to other state points and
302     to build phase diagrams. Figures \ref{fig:tp3PhaseDia} and
303     \ref{fig:ssdrfPhaseDia} are example phase diagrams built from the
304     results for the TIP3P and SSD/RF water models. All other models have
305     similar structure, although the crossing points between the phases
306     move to different temperatures and pressures as indicated from the
307     transition temperatures in table
308     \ref{tab:freeEnergy}. It is interesting to note that ice
309     I$_\textrm{h}$ (and ice I$_\textrm{c}$ for that matter) do not appear
310     in any of the phase diagrams for any of the models. For purposes of
311     this study, ice B is representative of the dense ice polymorphs. A
312     recent study by Sanz {\it et al.} provides details on the phase
313     diagrams for SPC/E and TIP4P at higher pressures than those studied
314     here.\cite{Sanz04}
315 chrisfen 2987 \begin{figure}
316     \centering
317     \includegraphics[width=\linewidth]{./figures/tp3PhaseDia.pdf}
318     \caption{Phase diagram for the TIP3P water model in the low pressure
319     regime. The displayed $T_m$ and $T_b$ values are good predictions of
320     the experimental values; however, the solid phases shown are not the
321 chrisfen 3023 experimentally observed forms. Both cubic and hexagonal ice I are
322 chrisfen 2987 higher in energy and don't appear in the phase diagram.}
323     \label{fig:tp3PhaseDia}
324     \end{figure}
325     \begin{figure}
326     \centering
327     \includegraphics[width=\linewidth]{./figures/ssdrfPhaseDia.pdf}
328     \caption{Phase diagram for the SSD/RF water model in the low pressure
329     regime. Calculations producing these results were done under an
330     applied reaction field. It is interesting to note that this
331     computationally efficient model (over 3 times more efficient than
332     TIP3P) exhibits phase behavior similar to the less computationally
333     conservative charge based models.}
334     \label{fig:ssdrfPhaseDia}
335     \end{figure}
336    
337     We note that all of the crystals investigated in this study are ideal
338     proton-ordered antiferroelectric structures. All of the structures
339     obey the Bernal-Fowler rules and should be able to form stable
340     proton-{\it disordered} crystals which have the traditional
341     $k_\textrm{B}$ln(3/2) residual entropy at 0~K.\cite{Bernal33,Pauling35}
342     Simulations of proton-disordered structures are relatively unstable
343     with all but the most expensive water models.\cite{Nada03} Our
344     simulations have therefore been performed with the ordered
345     antiferroelectric structures which do not require the residual entropy
346     term to be accounted for in the free energies. This may result in some
347     discrepancies when comparing our melting temperatures to the melting
348     temperatures that have been calculated via thermodynamic integrations
349     of the disordered structures.\cite{Sanz04}
350    
351     Most of the water models have melting points that compare quite
352     favorably with the experimental value of 273~K. The unfortunate
353     aspect of this result is that this phase change occurs between
354 chrisfen 3028 Ice-$i$ and the liquid state rather than ice I$_\textrm{h}$ and the liquid
355 chrisfen 2987 state. These results do not contradict other studies. Studies of ice
356 chrisfen 3028 I$_\textrm{h}$ using TIP4P predict a $T_m$ ranging from 191 to 238~K
357 chrisfen 2987 (differences being attributed to choice of interaction truncation and
358     different ordered and disordered molecular
359     arrangements).\cite{Nada03,Vlot99,Gao00,Sanz04} If the presence of ice
360 chrisfen 3028 B and Ice-$i$ were omitted, a $T_\textrm{m}$ value around 200~K
361 chrisfen 2987 would be predicted from this work. However, the $T_\textrm{m}$ from
362 chrisfen 3028 Ice-$i$ is calculated to be 262~K, indicating that these
363 chrisfen 3023 simulation-based structures ought to be included in studies probing
364 chrisfen 2987 phase transitions with this model. Also of interest in these results
365     is that SSD/E does not exhibit a melting point at 1 atm but does
366     sublime at 355~K. This is due to the significant stability of
367 chrisfen 3028 Ice-$i$ over all other polymorphs for this particular model under
368 chrisfen 2987 these conditions. While troubling, this behavior resulted in the
369 chrisfen 3028 spontaneous crystallization of Ice-$i$ which led us to investigate
370 chrisfen 2987 this structure. These observations provide a warning that simulations
371     of SSD/E as a ``liquid'' near 300~K are actually metastable and run
372     the risk of spontaneous crystallization. However, when a longer
373     cutoff radius is used, SSD/E prefers the liquid state under standard
374     temperature and pressure.
375    
376     \section{Effects of Potential Truncation}
377    
378     \begin{figure}
379 chrisfen 3028 \centering
380 chrisfen 2987 \includegraphics[width=\linewidth]{./figures/cutoffChange.pdf}
381     \caption{Free energy as a function of cutoff radius for SSD/E, TIP3P,
382     SPC/E, SSD/RF with a reaction field, and the TIP3P and SPC/E models
383     with an added Ewald correction term. Error for the larger cutoff
384     points is equivalent to that observed at 9.0~\AA\ (see Table
385     \ref{tab:freeEnergy}). Data for ice I$_\textrm{c}$ with TIP3P using
386     both 12 and 13.5~\AA\ cutoffs were omitted because the crystal was
387     prone to distortion and melting at 200~K. Ice-$i^\prime$ is the
388 chrisfen 3028 form of Ice-$i$ used in the SPC/E simulations.}
389 chrisfen 2987 \label{fig:incCutoff}
390     \end{figure}
391     For the more computationally efficient water models, we have also
392     investigated the effect of potential truncation on the computed free
393 chrisfen 3028 energies as a function of the cutoff radius. As seen in figure
394     \ref{fig:incCutoff}, the free energies of the ice polymorphs with
395 chrisfen 2987 water models lacking a long-range correction show significant cutoff
396 chrisfen 3028 radius dependence. In general, there is a narrowing of the free
397     energy differences while moving to greater cutoff radii. As the free
398 chrisfen 2987 energies for the polymorphs converge, the stability advantage that
399 chrisfen 3028 Ice-$i$ exhibits is reduced. Adjacent to each of these plots are
400 chrisfen 2987 results for systems with applied or estimated long-range corrections.
401     SSD/RF was parametrized for use with a reaction field, and the benefit
402     provided by this computationally inexpensive correction is apparent.
403     The free energies are largely independent of the size of the reaction
404     field cavity in this model, so small cutoff radii mimic bulk
405     calculations quite well under SSD/RF.
406    
407 chrisfen 3028 Although the point-charge models investigated here were parametrized
408     for use without the Ewald summation, we have estimated the effect of
409     this method for computing long-range electrostatics for both TIP3P and
410     SPC/E. This was accomplished by calculating the potential energy of
411     identical crystals both with and without {\sc spme}. Similar behavior
412     to that observed with reaction field is seen for both of these models.
413     The free energies show reduced dependence on cutoff radius and span a
414     narrower range for the various polymorphs. Like the dipolar water
415     models, TIP3P displays a relatively constant preference for the
416     Ice-$i$ polymorph. The crystal preference is much more difficult to
417     determine for SPC/E. Without a long-range correction, each of the
418     polymorphs studied assumes the role of the thermodynamically preferred
419     state under different cutoff radii. The inclusion of the Ewald
420     correction flattens and narrows the gap in free energies such that the
421     polymorphs are isoenergetic within statistical uncertainty. This
422     suggests that other conditions, such as the density in fixed-volume
423     simulations, can influence the polymorph expressed upon
424     crystallization.
425 chrisfen 2987
426     \section{Expanded Results Using Damped Shifted Force Electrostatics}
427    
428     In chapter \ref{chap:electrostatics}, we discussed in detail a
429     pairwise method for handling electrostatics (shifted force, {\sc sf})
430     that can be used as a simple and efficient replacement for the Ewald
431     summation. Answering the question of the free energies of these ice
432     polymorphs with varying water models would be an interesting
433     application of this technique. To this end, we set up thermodynamic
434     integrations of all of the previously discussed ice polymorphs using
435     the {\sc sf} technique with a cutoff radius of 12~\AA\ and an $\alpha$
436     of 0.2125~\AA . These calculations were performed on TIP5P-E and
437     TIP4P-Ew (variants of the root models optimized for the Ewald
438     summation) as well as SPC/E, SSD/RF, and TRED (see section
439     \ref{sec:tredWater}).
440    
441     \begin{table}
442     \centering
443     \caption{HELMHOLTZ FREE ENERGIES OF ICE POLYMORPHS USING THE DAMPED
444     SHIFTED FORCE CORRECTION}
445     \begin{tabular}{ lccccc }
446     \toprule
447     \toprule
448     Model & I$_\textrm{h}$ & I$_\textrm{c}$ & B & Ice-$i$ & Ice-$i^\prime$ \\
449     \cmidrule(lr){2-6}
450     & \multicolumn{5}{c}{(kcal mol$^{-1}$)} \\
451     \midrule
452 chrisfen 3019 TIP5P-E & -11.98(4) & -11.96(4) & -11.87(3) & - & -11.95(3) \\
453 chrisfen 3004 TIP4P-Ew & -13.11(3) & -13.09(3) & -12.97(3) & - & -12.98(3) \\
454     SPC/E & -12.99(3) & -13.00(3) & -13.03(3) & - & -12.99(3) \\
455 chrisfen 3001 SSD/RF & -11.83(3) & -11.66(4) & -12.32(3) & -12.39(3) & - \\
456     TRED & -12.61(3) & -12.43(3) & -12.89(3) & -13.12(3) & - \\
457 chrisfen 3042 \bottomrule
458 chrisfen 2987 \end{tabular}
459     \label{tab:dampedFreeEnergy}
460     \end{table}
461 chrisfen 3001 The results of these calculations in table \ref{tab:dampedFreeEnergy}
462     show similar behavior to the Ewald results in figure
463     \ref{fig:incCutoff}, at least for SSD/RF and SPC/E which are present
464 chrisfen 3023 in both. The Helmholtz free energies of the ice polymorphs for SSD/RF
465 chrisfen 3028 order in the same fashion; however, Ice-$i$ and ice B are quite a bit
466 chrisfen 3023 closer in free energy (nearly isoenergetic). The free energy
467     differences between ice polymorphs for TRED water parallel SSD/RF,
468 chrisfen 3028 with the exception that ice B is destabilized such that the free
469     energy is not very close to Ice-$i$. The SPC/E results really show the
470     near isoenergetic behavior when using the electrostatic
471     correction. Ice B has the lowest Helmholtz free energy; however, all
472     the polymorph results overlap within error.
473 chrisfen 2987
474 chrisfen 3004 The most interesting results from these calculations come from the
475     more expensive TIP4P-Ew and TIP5P-E results. Both of these models were
476     optimized for use with an electrostatic correction and are
477     geometrically arranged to mimic water following two different
478     ideas. In TIP5P-E, the primary location for the negative charge in the
479     molecule is assigned to the lone-pairs of the oxygen, while TIP4P-Ew
480     places the negative charge near the center-of-mass along the H-O-H
481     bisector. There is some debate as to which is the proper choice for
482     the negative charge location, and this has in part led to a six-site
483     water model that balances both of these options.\cite{Vega05,Nada03}
484     The limited results in table \ref{tab:dampedFreeEnergy} support the
485     results of Vega {\it et al.}, which indicate the TIP4P charge location
486     geometry is more physically valid.\cite{Vega05} With the TIP4P-Ew
487     water model, the experimentally observed polymorph (ice
488     I$_\textrm{h}$) is the preferred form with ice I$_\textrm{c}$ slightly
489 chrisfen 3028 higher in energy, though overlapping within error. Additionally, the
490     less realistic ice B and Ice-$i^\prime$ structures are destabilized
491 chrisfen 3023 relative to these polymorphs. TIP5P-E shows similar behavior to SPC/E,
492     where there is no real free energy distinction between the various
493 chrisfen 3028 polymorphs, because many overlap within error. While ice B is close in
494 chrisfen 3023 free energy to the other polymorphs, these results fail to support the
495     findings of other researchers indicating the preferred form of TIP5P
496     at 1~atm is a structure similar to ice
497     B.\cite{Yamada02,Vega05,Abascal05} It should be noted that we are
498     looking at TIP5P-E rather than TIP5P, and the differences in the
499     Lennard-Jones parameters could be a reason for this dissimilarity.
500     Overall, these results indicate that TIP4P-Ew is a better mimic of
501     real water than these other models when studying crystallization and
502     solid forms of water.
503 chrisfen 3004
504 chrisfen 2987 \section{Conclusions}
505    
506     In this work, thermodynamic integration was used to determine the
507     absolute free energies of several ice polymorphs. The new polymorph,
508 chrisfen 3028 Ice-$i$ was observed to be the stable crystalline state for {\it all}
509     the water models when using a 9.0~\AA\ cutoff. Additional work showed
510     that the free energy depends in part on simulation conditions - in
511     particular, the choice of long-range correction method. Regardless,
512     Ice-$i$ was still observed to be a stable polymorph for all of the
513     studied water models.
514 chrisfen 2987
515 chrisfen 3042 So what is the preferred solid polymorph for simulated water? The
516     answer appears to be dependent both on the conditions and the model
517     used. In the case of short cutoffs without a long-range interaction
518     correction, Ice-$i$ and Ice-$i^\prime$ have the lowest free energy of
519     the studied polymorphs with all the models. Ideally, crystallization
520     of each model under constant pressure conditions, as was done with
521     SSD/E, would aid in identifying their respective preferred structures.
522     This work, however, helps illustrate how studies involving one
523     specific model can lead to insight about important behavior of others.
524 chrisfen 2987
525     We also note that none of the water models used in this study are
526 chrisfen 3016 polarizable or flexible models. It is entirely possible that the
527 chrisfen 3023 polarizability of real water makes the Ice-$i$ structure substantially
528     less stable than ice I$_\textrm{h}$. The dipole moment of the water
529     molecules increases as the system becomes more condensed, and the
530     increasing dipole moment should destabilize the tetramer structures in
531 chrisfen 3016 Ice-$i$. Right now, using TIP4P-Ew with an electrostatic correction
532     gives the proper thermodynamically preferred state, and we recommend
533     this arrangement for study of crystallization processes if the
534     computational cost increase that comes with including polarizability
535     is an issue.
536 chrisfen 2987
537 chrisfen 3042 Finally, the stability of Ice-$i$ in these simulations raises the
538     possibility of experimental observation. The extensive body of
539     research on water in the low pressure regime makes us hesitant to
540     ascribe any relevance to this work outside the simulation community.
541     It is for this reason that we chose a name for this polymorph
542     involving an imaginary quantity. That said, there are certain
543     conditions that would be ideal for experimental observation of
544     Ice-$i$. These include the negative pressure or stretched solid
545     regime, clusters deposited in vacuum environments, and clathrate
546     structures involving small non-polar molecules. For the purpose of
547     comparison with future experimental results, we calculated the
548     oxygen-oxygen pair correlation function, $g_\textrm{OO}(r)$, and the
549     structure factor, $S(\vec{q})$ for the two Ice-$i$ variants (along
550     with ice I$_\textrm{h}$ and I$_\textrm{c}$) at 77~K (figures
551     \ref{fig:gofr} and \ref{fig:sofq}). It is interesting to note that
552     the structure factors for Ice-$i^\prime$ and ice I$_\textrm{c}$ are
553     quite similar. The primary differences are small peaks at 1.125,
554     2.29, and 2.53~\AA$^{-1}$, so particular attention to these regions
555     would be needed to distinguish Ice-$i^\prime$ from ice I$_\textrm{c}$.
556 chrisfen 2987
557     \begin{figure}
558     \includegraphics[width=\linewidth]{./figures/iceGofr.pdf}
559 chrisfen 3028 \caption{Radial distribution functions of Ice-$i$ and ice
560 chrisfen 2987 I$_\textrm{c}$ calculated from from simulations of the SSD/RF water
561     model at 77~K.}
562     \label{fig:gofr}
563     \end{figure}
564    
565     \begin{figure}
566     \includegraphics[width=\linewidth]{./figures/sofq.pdf}
567 chrisfen 3028 \caption{Predicted structure factors for Ice-$i$ and ice
568 chrisfen 2987 I$_\textrm{c}$ at 77~K. The raw structure factors have been
569 chrisfen 3023 convoluted with a Gaussian instrument function (0.075~\AA$^{-1}$
570 chrisfen 2987 width) to compensate for the truncation effects in our finite size
571 chrisfen 3023 simulations.}
572 chrisfen 2987 \label{fig:sofq}
573     \end{figure}
574