ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/iceiPaper/iceiPaper.tex
Revision: 1464
Committed: Wed Sep 15 21:44:27 2004 UTC (19 years, 9 months ago) by chrisfen
Content type: application/x-tex
File size: 23978 byte(s)
Log Message:
minor changes

File Contents

# Content
1 %\documentclass[prb,aps,twocolumn,tabularx]{revtex4}
2 \documentclass[11pt]{article}
3 %\documentclass[11pt]{article}
4 \usepackage{endfloat}
5 \usepackage{amsmath}
6 \usepackage{epsf}
7 \usepackage{berkeley}
8 \usepackage{setspace}
9 \usepackage{tabularx}
10 \usepackage{graphicx}
11 \usepackage[ref]{overcite}
12 \pagestyle{plain}
13 \pagenumbering{arabic}
14 \oddsidemargin 0.0cm \evensidemargin 0.0cm
15 \topmargin -21pt \headsep 10pt
16 \textheight 9.0in \textwidth 6.5in
17 \brokenpenalty=10000
18 \renewcommand{\baselinestretch}{1.2}
19 \renewcommand\citemid{\ } % no comma in optional reference note
20
21 \begin{document}
22
23 \title{Ice-{\it i}: a novel ice polymorph predicted via computer simulation}
24
25 \author{Christopher J. Fennell and J. Daniel Gezelter \\
26 Department of Chemistry and Biochemistry\\ University of Notre Dame\\
27 Notre Dame, Indiana 46556}
28
29 \date{\today}
30
31 \maketitle
32 %\doublespacing
33
34 \begin{abstract}
35 The free energies of several ice polymorphs in the low pressure regime
36 were calculated using thermodynamic integration. These integrations
37 were done for most of the common water models. Ice-{\it i}, a
38 structure we recently observed to be stable in one of the single-point
39 water models, was determined to be the stable crystalline state (at 1
40 atm) for {\it all} the water models investigated. Phase diagrams were
41 generated, and phase coexistence lines were determined for all of the
42 known low-pressure ice structures under all of the common water
43 models. Additionally, potential truncation was shown to have an
44 effect on the calculated free energies, and can result in altered free
45 energy landscapes.
46 \end{abstract}
47
48 %\narrowtext
49
50 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
51 % BODY OF TEXT
52 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
53
54 \section{Introduction}
55
56 Molecular dynamics is a valuable tool for studying the phase behavior
57 of systems ranging from small or simple
58 molecules\cite{Matsumoto02andOthers} to complex biological
59 species.\cite{bigStuff} Many techniques have been developed to
60 investigate the thermodynamic properites of model substances,
61 providing both qualitative and quantitative comparisons between
62 simulations and experiment.\cite{thermMethods} Investigation of these
63 properties leads to the development of new and more accurate models,
64 leading to better understanding and depiction of physical processes
65 and intricate molecular systems.
66
67 Water has proven to be a challenging substance to depict in
68 simulations, and a variety of models have been developed to describe
69 its behavior under varying simulation
70 conditions.\cite{Berendsen81,Jorgensen83,Bratko85,Berendsen87,Liu96,Mahoney00,Fennell04}
71 These models have been used to investigate important physical
72 phenomena like phase transitions and the hydrophobic
73 effect.\cite{Yamada02} With the choice of models available, it
74 is only natural to compare the models under interesting thermodynamic
75 conditions in an attempt to clarify the limitations of each of the
76 models.\cite{modelProps} Two important property to quantify are the
77 Gibbs and Helmholtz free energies, particularly for the solid forms of
78 water. Difficulty in these types of studies typically arises from the
79 assortment of possible crystalline polymorphs that water adopts over a
80 wide range of pressures and temperatures. There are currently 13
81 recognized forms of ice, and it is a challenging task to investigate
82 the entire free energy landscape.\cite{Sanz04} Ideally, research is
83 focused on the phases having the lowest free energy at a given state
84 point, because these phases will dictate the true transition
85 temperatures and pressures for their respective model.
86
87 In this paper, standard reference state methods were applied to the
88 study of crystalline water polymorphs in the low pressure regime. This
89 work is unique in the fact that one of the crystal lattices was
90 arrived at through crystallization of a computationally efficient
91 water model under constant pressure and temperature
92 conditions. Crystallization events are interesting in and of
93 themselves\cite{Matsumoto02,Yamada02}; however, the crystal structure
94 obtained in this case was different from any previously observed ice
95 polymorphs, in experiment or simulation.\cite{Fennell04} This crystal
96 was termed Ice-{\it i} in homage to its origin in computational
97 simulation. The unit cell (Fig. \ref{iceiCell}A) consists of eight
98 water molecules that stack in rows of interlocking water
99 tetramers. Proton ordering can be accomplished by orienting two of the
100 waters so that both of their donated hydrogen bonds are internal to
101 their tetramer (Fig. \ref{protOrder}). As expected in an ice crystal
102 constructed of water tetramers, the hydrogen bonds are not as linear
103 as those observed in ice $I_h$, however the interlocking of these
104 subunits appears to provide significant stabilization to the overall
105 crystal. The arrangement of these tetramers results in surrounding
106 open octagonal cavities that are typically greater than 6.3 \AA\ in
107 diameter. This relatively open overall structure leads to crystals
108 that are 0.07 g/cm$^3$ less dense on average than ice $I_h$.
109
110 \begin{figure}
111 \includegraphics[width=\linewidth]{unitCell.eps}
112 \caption{Unit cells for (A) Ice-{\it i} and (B) Ice-2{\it i}, the elongated variant of Ice-{\it i}. For Ice-{\it i}, the $a$ to $c$ relation is given by $a = 2.1214c$, while for Ice-2{\it i}, $a = 1.7850c$.}
113 \label{iceiCell}
114 \end{figure}
115
116 \begin{figure}
117 \includegraphics[width=\linewidth]{orderedIcei.eps}
118 \caption{Image of a proton ordered crystal of Ice-{\it i} looking
119 down the (001) crystal face. The rows of water tetramers surrounded by
120 octagonal pores leads to a crystal structure that is significantly
121 less dense than ice $I_h$.}
122 \label{protOrder}
123 \end{figure}
124
125 Results in the previous study indicated that Ice-{\it i} is the
126 minimum energy crystal structure for the single point water models
127 being studied (for discussions on these single point dipole models,
128 see the previous work and related
129 articles\cite{Fennell04,Ichiye96,Bratko85}). Those results only
130 consider energetic stabilization and neglect entropic contributions to
131 the overall free energy. To address this issue, the absolute free
132 energy of this crystal was calculated using thermodynamic integration
133 and compared to the free energies of cubic and hexagonal ice $I$ (the
134 experimental low density ice polymorphs) and ice B (a higher density,
135 but very stable crystal structure observed by B\`{a}ez and Clancy in
136 free energy studies of SPC/E).\cite{Baez95b} This work includes
137 results for the water model from which Ice-{\it i} was crystallized
138 (soft sticky dipole extended, SSD/E) in addition to several common
139 water models (TIP3P, TIP4P, TIP5P, and SPC/E) and a reaction field
140 parametrized single point dipole water model (soft sticky dipole
141 reaction field, SSD/RF). In should be noted that a second version of
142 Ice-{\it i} (Ice-2{\it i}) was used in calculations involving SPC/E,
143 TIP4P, and TIP5P. The unit cell of this crystal (Fig. \ref{iceiCell}B)
144 is similar to the Ice-{\it i} unit it is extended in the direction of
145 the (001) face and compressed along the other two faces.
146
147 \section{Methods}
148
149 Canonical ensemble (NVT) molecular dynamics calculations were
150 performed using the OOPSE (Object-Oriented Parallel Simulation Engine)
151 molecular mechanics package. All molecules were treated as rigid
152 bodies, with orientational motion propagated using the symplectic DLM
153 integration method. Details about the implementation of these
154 techniques can be found in a recent publication.\cite{Meineke05}
155
156 Thermodynamic integration was utilized to calculate the free energy of
157 several ice crystals at 200 K using the TIP3P, TIP4P, TIP5P, SPC/E,
158 SSD/RF, and SSD/E water models. Liquid state free energies at 300 and
159 400 K for all of these water models were also determined using this
160 same technique, in order to determine melting points and generate
161 phase diagrams. All simulations were carried out at densities
162 resulting in a pressure of approximately 1 atm at their respective
163 temperatures.
164
165 A single thermodynamic integration involves a sequence of simulations
166 over which the system of interest is converted into a reference system
167 for which the free energy is known. This transformation path is then
168 integrated in order to determine the free energy difference between
169 the two states:
170 \begin{equation}
171 \Delta A = \int_0^1\left\langle\frac{\partial V(\lambda
172 )}{\partial\lambda}\right\rangle_\lambda d\lambda,
173 \end{equation}
174 where $V$ is the interaction potential and $\lambda$ is the
175 transformation parameter that scales the overall
176 potential. Simulations are distributed unevenly along this path in
177 order to sufficiently sample the regions of greatest change in the
178 potential. Typical integrations in this study consisted of $\sim$25
179 simulations ranging from 300 ps (for the unaltered system) to 75 ps
180 (near the reference state) in length.
181
182 For the thermodynamic integration of molecular crystals, the Einstein
183 Crystal is chosen as the reference state that the system is converted
184 to over the course of the simulation. In an Einstein Crystal, the
185 molecules are harmonically restrained at their ideal lattice locations
186 and orientations. The partition function for a molecular crystal
187 restrained in this fashion has been evaluated, and the Helmholtz Free
188 Energy ({\it A}) is given by
189 \begin{eqnarray}
190 A = E_m\ -\ kT\ln \left (\frac{kT}{h\nu}\right )^3&-&kT\ln \left
191 [\pi^\frac{1}{2}\left (\frac{8\pi^2I_\mathrm{A}kT}{h^2}\right
192 )^\frac{1}{2}\left (\frac{8\pi^2I_\mathrm{B}kT}{h^2}\right
193 )^\frac{1}{2}\left (\frac{8\pi^2I_\mathrm{C}kT}{h^2}\right
194 )^\frac{1}{2}\right ] \nonumber \\ &-& kT\ln \left [\frac{kT}{2(\pi
195 K_\omega K_\theta)^{\frac{1}{2}}}\exp\left
196 (-\frac{kT}{2K_\theta}\right)\int_0^{\left (\frac{kT}{2K_\theta}\right
197 )^\frac{1}{2}}\exp(t^2)\mathrm{d}t\right ],
198 \label{ecFreeEnergy}
199 \end{eqnarray}
200 where $2\pi\nu = (K_\mathrm{v}/m)^{1/2}$.\cite{Baez95a} In equation
201 \ref{ecFreeEnergy}, $K_\mathrm{v}$, $K_\mathrm{\theta}$, and
202 $K_\mathrm{\omega}$ are the spring constants restraining translational
203 motion and deflection of and rotation around the principle axis of the
204 molecule respectively (Fig. \ref{waterSpring}), and $E_m$ is the
205 minimum potential energy of the ideal crystal. In the case of
206 molecular liquids, the ideal vapor is chosen as the target reference
207 state.
208
209 \begin{figure}
210 \includegraphics[width=\linewidth]{rotSpring.eps}
211 \caption{Possible orientational motions for a restrained molecule.
212 $\theta$ angles correspond to displacement from the body-frame {\it
213 z}-axis, while $\omega$ angles correspond to rotation about the
214 body-frame {\it z}-axis. $K_\theta$ and $K_\omega$ are spring
215 constants for the harmonic springs restraining motion in the $\theta$
216 and $\omega$ directions.}
217 \label{waterSpring}
218 \end{figure}
219
220 Charge, dipole, and Lennard-Jones interactions were modified by a
221 cubic switching between 100\% and 85\% of the cutoff value (9 \AA
222 ). By applying this function, these interactions are smoothly
223 truncated, thereby avoiding poor energy conserving dynamics resulting
224 from harsher truncation schemes. The effect of a long-range correction
225 was also investigated on select model systems in a variety of
226 manners. For the SSD/RF model, a reaction field with a fixed
227 dielectric constant of 80 was applied in all
228 simulations.\cite{Onsager36} For a series of the least computationally
229 expensive models (SSD/E, SSD/RF, and TIP3P), simulations were
230 performed with longer cutoffs of 12 and 15 \AA\ to compare with the 9
231 \AA\ cutoff results. Finally, results from the use of an Ewald
232 summation were estimated for TIP3P and SPC/E by performing
233 calculations with Particle-Mesh Ewald (PME) in the TINKER molecular
234 mechanics software package.\cite{Tinker} TINKER was chosen because it
235 can also propagate the motion of rigid-bodies, and provides the most
236 direct comparison to the results from OOPSE. The calculated energy
237 difference in the presence and absence of PME was applied to the
238 previous results in order to predict changes in the free energy
239 landscape.
240
241 \section{Results and discussion}
242
243 The free energy of proton ordered Ice-{\it i} was calculated and
244 compared with the free energies of proton ordered variants of the
245 experimentally observed low-density ice polymorphs, $I_h$ and $I_c$,
246 as well as the higher density ice B, observed by B\`{a}ez and Clancy
247 and thought to be the minimum free energy structure for the SPC/E
248 model at ambient conditions (Table \ref{freeEnergy}).\cite{Baez95b}
249 Ice XI, the experimentally observed proton ordered variant of ice
250 $I_h$, was investigated initially, but it was found not to be as
251 stable as antiferroelectric variants of proton ordered or even proton
252 disordered ice$I_h$.\cite{Davidson84} The proton ordered variant of
253 ice $I_h$ used here is a simple antiferroelectric version that has an
254 8 molecule unit cell. The crystals contained 648 or 1728 molecules for
255 ice B, 1024 or 1280 molecules for ice $I_h$, 1000 molecules for ice
256 $I_c$, or 1024 molecules for Ice-{\it i}. The larger crystal sizes
257 were necessary for simulations involving larger cutoff values.
258
259 \begin{table*}
260 \begin{minipage}{\linewidth}
261 \renewcommand{\thefootnote}{\thempfootnote}
262 \begin{center}
263 \caption{Calculated free energies for several ice polymorphs with a
264 variety of common water models. All calculations used a cutoff radius
265 of 9 \AA\ and were performed at 200 K and $\sim$1 atm. Units are
266 kcal/mol. *Ice $I_c$ is unstable at 200 K using SSD/RF.}
267 \begin{tabular}{ l c c c c }
268 \hline
269 \ \quad \ Water Model\ \ & \ \quad \ \ \ \ $I_h$ \ \ & \ \quad \ \ \ \ $I_c$ \ \ & \ \quad \ \ \ \ B \ \ & \ \quad \ \ \ Ice-{\it i} \ \quad \ \\
270 \hline
271 \ \quad \ TIP3P & \ \quad \ -11.41 & \ \quad \ -11.23 & \ \quad \ -11.82 & \quad -12.30\\
272 \ \quad \ TIP4P & \ \quad \ -11.84 & \ \quad \ -12.04 & \ \quad \ -12.08 & \quad -12.33\\
273 \ \quad \ TIP5P & \ \quad \ -11.85 & \ \quad \ -11.86 & \ \quad \ -11.96 & \quad -12.29\\
274 \ \quad \ SPC/E & \ \quad \ -12.67 & \ \quad \ -12.96 & \ \quad \ -13.25 & \quad -13.55\\
275 \ \quad \ SSD/E & \ \quad \ -11.27 & \ \quad \ -11.19 & \ \quad \ -12.09 & \quad -12.54\\
276 \ \quad \ SSD/RF & \ \quad \ -11.51 & \ \quad \ NA* & \ \quad \ -12.08 & \quad -12.29\\
277 \end{tabular}
278 \label{freeEnergy}
279 \end{center}
280 \end{minipage}
281 \end{table*}
282
283 The free energy values computed for the studied polymorphs indicate
284 that Ice-{\it i} is the most stable state for all of the common water
285 models studied. With the free energy at these state points, the
286 temperature and pressure dependence of the free energy was used to
287 project to other state points and build phase diagrams. Figures
288 \ref{tp3phasedia} and \ref{ssdrfphasedia} are example diagrams built
289 from the free energy results. All other models have similar structure,
290 only the crossing points between these phases exist at different
291 temperatures and pressures. It is interesting to note that ice $I$
292 does not exist in either cubic or hexagonal form in any of the phase
293 diagrams for any of the models. For purposes of this study, ice B is
294 representative of the dense ice polymorphs. A recent study by Sanz
295 {\it et al.} goes into detail on the phase diagrams for SPC/E and
296 TIP4P in the high pressure regime.\cite{Sanz04}
297
298 \begin{figure}
299 \includegraphics[width=\linewidth]{tp3PhaseDia.eps}
300 \caption{Phase diagram for the TIP3P water model in the low pressure
301 regime. The displayed $T_m$ and $T_b$ values are good predictions of
302 the experimental values; however, the solid phases shown are not the
303 experimentally observed forms. Both cubic and hexagonal ice $I$ are
304 higher in energy and don't appear in the phase diagram.}
305 \label{tp3phasedia}
306 \end{figure}
307
308 \begin{figure}
309 \includegraphics[width=\linewidth]{ssdrfPhaseDia.eps}
310 \caption{Phase diagram for the SSD/RF water model in the low pressure
311 regime. Calculations producing these results were done under an
312 applied reaction field. It is interesting to note that this
313 computationally efficient model (over 3 times more efficient than
314 TIP3P) exhibits phase behavior similar to the less computationally
315 conservative charge based models.}
316 \label{ssdrfphasedia}
317 \end{figure}
318
319 \begin{table*}
320 \begin{minipage}{\linewidth}
321 \renewcommand{\thefootnote}{\thempfootnote}
322 \begin{center}
323 \caption{Melting ($T_m$), boiling ($T_b$), and sublimation ($T_s$)
324 temperatures of several common water models compared with experiment.}
325 \begin{tabular}{ l c c c c c c c }
326 \hline
327 \ \ Equilibria Point\ \ & \ \ \ \ \ TIP3P \ \ & \ \ \ \ \ TIP4P \ \ & \ \quad \ \ \ \ TIP5P \ \ & \ \ \ \ \ SPC/E \ \ & \ \ \ \ \ SSD/E \ \ & \ \ \ \ \ SSD/RF \ \ & \ \ \ \ \ Exp. \ \ \\
328 \hline
329 \ \ $T_m$ (K) & \ \ 269 & \ \ 265 & \ \ 271 & 297 & \ \ - & \ \ 278 & \ \ 273\\
330 \ \ $T_b$ (K) & \ \ 357 & \ \ 354 & \ \ 337 & 396 & \ \ - & \ \ 349 & \ \ 373\\
331 \ \ $T_s$ (K) & \ \ - & \ \ - & \ \ - & - & \ \ 355 & \ \ - & \ \ -\\
332 \end{tabular}
333 \label{meltandboil}
334 \end{center}
335 \end{minipage}
336 \end{table*}
337
338 Table \ref{meltandboil} lists the melting and boiling temperatures
339 calculated from this work. Surprisingly, most of these models have
340 melting points that compare quite favorably with experiment. The
341 unfortunate aspect of this result is that this phase change occurs
342 between Ice-{\it i} and the liquid state rather than ice $I_h$ and the
343 liquid state. These results are actually not contrary to previous
344 studies in the literature. Earlier free energy studies of ice $I$
345 using TIP4P predict a $T_m$ ranging from 214 to 238 K (differences
346 being attributed to choice of interaction truncation and different
347 ordered and disordered molecular arrangements). If the presence of ice
348 B and Ice-{\it i} were omitted, a $T_m$ value around 210 K would be
349 predicted from this work. However, the $T_m$ from Ice-{\it i} is
350 calculated at 265 K, significantly higher in temperature than the
351 previous studies. Also of interest in these results is that SSD/E does
352 not exhibit a melting point at 1 atm, but it shows a sublimation point
353 at 355 K. This is due to the significant stability of Ice-{\it i} over
354 all other polymorphs for this particular model under these
355 conditions. While troubling, this behavior turned out to be
356 advantageous in that it facilitated the spontaneous crystallization of
357 Ice-{\it i}. These observations provide a warning that simulations of
358 SSD/E as a ``liquid'' near 300 K are actually metastable and run the
359 risk of spontaneous crystallization. However, this risk changes when
360 applying a longer cutoff.
361
362 \begin{figure}
363 \includegraphics[width=\linewidth]{cutoffChange.eps}
364 \caption{Free energy as a function of cutoff radius for (A) SSD/E, (B)
365 TIP3P, and (C) SSD/RF. Data points omitted include SSD/E: $I_c$ 12
366 \AA\, TIP3P: $I_c$ 12 \AA\ and B 12 \AA\, and SSD/RF: $I_c$ 9
367 \AA\. These crystals are unstable at 200 K and rapidly convert into a
368 liquid. The connecting lines are qualitative visual aid.}
369 \label{incCutoff}
370 \end{figure}
371
372 Increasing the cutoff radius in simulations of the more
373 computationally efficient water models was done in order to evaluate
374 the trend in free energy values when moving to systems that do not
375 involve potential truncation. As seen in Fig. \ref{incCutoff}, the
376 free energy of all the ice polymorphs show a substantial dependence on
377 cutoff radius. In general, there is a narrowing of the free energy
378 differences while moving to greater cutoff radius. Interestingly, by
379 increasing the cutoff radius, the free energy gap was narrowed enough
380 in the SSD/E model that the liquid state is preferred under standard
381 simulation conditions (298 K and 1 atm). Thus, it is recommended that
382 simulations using this model choose interaction truncation radii
383 greater than 9 \AA\. This narrowing trend is much more subtle in the
384 case of SSD/RF, indicating that the free energies calculated with a
385 reaction field present provide a more accurate picture of the free
386 energy landscape in the absence of potential truncation.
387
388 To further study the changes resulting to the inclusion of a
389 long-range interaction correction, the effect of an Ewald summation
390 was estimated by applying the potential energy difference do to its
391 inclusion in systems in the presence and absence of the
392 correction. This was accomplished by calculation of the potential
393 energy of identical crystals with and without PME using TINKER. The
394 free energies for the investigated polymorphs using the TIP3P and
395 SPC/E water models are shown in Table \ref{pmeShift}. TIP4P and TIP5P
396 are not fully supported in TINKER, so the results for these models
397 could not be estimated. The same trend pointed out through increase of
398 cutoff radius is observed in these PME results. Ice-{\it i} is the
399 preferred polymorph at ambient conditions for both the TIP3P and SPC/E
400 water models; however, there is a narrowing of the free energy
401 differences between the various solid forms. In the case of SPC/E this
402 narrowing is significant enough that it becomes less clear cut that
403 Ice-{\it i} is the most stable polymorph, and is possibly metastable
404 with respect to ice B and possibly ice $I_c$. However, these results
405 do not significantly alter the finding that the Ice-{\it i} polymorph
406 is a stable crystal structure that should be considered when studying
407 the phase behavior of water models.
408
409 \begin{table*}
410 \begin{minipage}{\linewidth}
411 \renewcommand{\thefootnote}{\thempfootnote}
412 \begin{center}
413 \caption{The free energy of the studied ice polymorphs after applying
414 the energy difference attributed to the inclusion of the PME
415 long-range interaction correction. Units are kcal/mol.}
416 \begin{tabular}{ l c c c c }
417 \hline
418 \ \ Water Model \ \ & \ \ \ \ \ $I_h$ \ \ & \ \ \ \ \ $I_c$ \ \ & \ \quad \ \ \ \ B \ \ & \ \ \ \ \ Ice-{\it i} \ \ \\
419 \hline
420 \ \ TIP3P & \ \ -11.53 & \ \ -11.24 & \ \ -11.51 & \ \ -11.67\\
421 \ \ SPC/E & \ \ -12.77 & \ \ -12.92 & \ \ -12.96 & \ \ -13.02\\
422 \end{tabular}
423 \label{pmeShift}
424 \end{center}
425 \end{minipage}
426 \end{table*}
427
428 \section{Conclusions}
429
430 The free energy for proton ordered variants of hexagonal and cubic ice
431 $I$, ice B, and recently discovered Ice-{\it i} where calculated under
432 standard conditions for several common water models via thermodynamic
433 integration. All the water models studied show Ice-{\it i} to be the
434 minimum free energy crystal structure in the with a 9 \AA\ switching
435 function cutoff. Calculated melting and boiling points show
436 surprisingly good agreement with the experimental values; however, the
437 solid phase at 1 atm is Ice-{\it i}, not ice $I_h$. The effect of
438 interaction truncation was investigated through variation of the
439 cutoff radius, use of a reaction field parameterized model, and
440 estimation of the results in the presence of the Ewald summation
441 correction. Interaction truncation has a significant effect on the
442 computed free energy values, and may significantly alter the free
443 energy landscape for the more complex multipoint water models. Despite
444 these effects, these results show Ice-{\it i} to be an important ice
445 polymorph that should be considered in simulation studies.
446
447 Due to this relative stability of Ice-{\it i} in all manner of
448 investigated simulation examples, the question arises as to possible
449 experimental observation of this polymorph. The rather extensive past
450 and current experimental investigation of water in the low pressure
451 regime leads the authors to be hesitant in ascribing relevance outside
452 of computational models, hence the descriptive name presented. That
453 being said, there are certain experimental conditions that would
454 provide the most ideal situation for possible observation. These
455 include the negative pressure or stretched solid regime, small
456 clusters in vacuum deposition environments, and in clathrate
457 structures involving small non-polar molecules.
458
459 \section{Acknowledgments}
460 Support for this project was provided by the National Science
461 Foundation under grant CHE-0134881. Computation time was provided by
462 the Notre Dame High Performance Computing Cluster and the Notre Dame
463 Bunch-of-Boxes (B.o.B) computer cluster (NSF grant DMR-0079647).
464
465 \newpage
466
467 \bibliographystyle{jcp}
468 \bibliography{iceiPaper}
469
470
471 \end{document}