ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/iceiPaper/iceiPaper.tex
Revision: 2134
Committed: Fri Mar 25 19:11:49 2005 UTC (19 years, 3 months ago) by chrisfen
Content type: application/x-tex
File size: 24317 byte(s)
Log Message:
added a reference

File Contents

# Content
1 %\documentclass[prb,aps,twocolumn,tabularx]{revtex4}
2 \documentclass[12pt]{article}
3 \usepackage{endfloat}
4 \usepackage{amsmath}
5 \usepackage{epsf}
6 \usepackage{times}
7 \usepackage{mathptm}
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{Computational free energy studies of a new ice polymorph which
24 exhibits greater stability than Ice I$_h$}
25
26 \author{Christopher J. Fennell and J. Daniel Gezelter \\
27 Department of Chemistry and Biochemistry\\
28 University of Notre Dame\\
29 Notre Dame, Indiana 46556}
30
31 \date{\today}
32
33 \maketitle
34 %\doublespacing
35
36 \begin{abstract}
37 The absolute free energies of several ice polymorphs were calculated
38 using thermodynamic integration. These polymorphs are predicted by
39 computer simulations using a variety of common water models to be
40 stable at low pressures. A recently discovered ice polymorph that has
41 as yet {\it only} been observed in computer simulations (Ice-{\it i}),
42 was determined to be the stable crystalline state for {\it all} the
43 water models investigated. Phase diagrams were generated, and phase
44 coexistence lines were determined for all of the known low-pressure
45 ice structures. Additionally, potential truncation was shown to play
46 a role in the resulting shape of the free energy landscape.
47 \end{abstract}
48
49 %\narrowtext
50
51 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
52 % BODY OF TEXT
53 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
54
55 \section{Introduction}
56
57 Water has proven to be a challenging substance to depict in
58 simulations, and a variety of models have been developed to describe
59 its behavior under varying simulation
60 conditions.\cite{Stillinger74,Rahman75,Berendsen81,Jorgensen83,Bratko85,Berendsen87,Caldwell95,Liu96,Berendsen98,Dill00,Mahoney00,Fennell04}
61 These models have been used to investigate important physical
62 phenomena like phase transitions, transport properties, and the
63 hydrophobic effect.\cite{Yamada02,Marrink94,Gallagher03} With the
64 choice of models available, it is only natural to compare the models
65 under interesting thermodynamic conditions in an attempt to clarify
66 the limitations of
67 each.\cite{Jorgensen83,Jorgensen98b,Clancy94,Mahoney01} Two important
68 properties to quantify are the Gibbs and Helmholtz free energies,
69 particularly for the solid forms of water as these predict the
70 thermodynamic stability of the various phases. Water has a
71 particularly rich phase diagram and takes on a number of different and
72 stable crystalline structures as the temperature and pressure are
73 varied. It is a challenging task to investigate the entire free
74 energy landscape\cite{Sanz04}; and ideally, research is focused on the
75 phases having the lowest free energy at a given state point, because
76 these phases will dictate the relevant transition temperatures and
77 pressures for the model.
78
79 The high-pressure phases of water (ice II - ice X as well as ice XII)
80 have been studied extensively both experimentally and
81 computationally. In this paper, standard reference state methods were
82 applied in the {\it low} pressure regime to evaluate the free energies
83 for a few known crystalline water polymorphs that might be stable at
84 these pressures. This work is unique in that one of the crystal
85 lattices was arrived at through crystallization of a computationally
86 efficient water model under constant pressure and temperature
87 conditions. Crystallization events are interesting in and of
88 themselves\cite{Matsumoto02,Yamada02}; however, the crystal structure
89 obtained in this case is different from any previously observed ice
90 polymorphs in experiment or simulation.\cite{Fennell04} We have named
91 this structure Ice-{\it i} to indicate its origin in computational
92 simulation. The unit cell of Ice-{\it i} and an axially-elongated
93 variant named Ice-{\it i}$^\prime$ both consist of eight water
94 molecules that stack in rows of interlocking water tetramers as
95 illustrated in figures \ref{unitcell}A and \ref{unitcell}B. These
96 tetramers form a crystal structure similar in appearance to a recent
97 two-dimensional surface tessellation simulated on silica.\cite{Yang04}
98 As expected in an ice crystal constructed of water tetramers, the
99 hydrogen bonds are not as linear as those observed in ice I$_h$,
100 however the interlocking of these subunits appears to provide
101 significant stabilization to the overall crystal. The arrangement of
102 these tetramers results in octagonal cavities that are typically
103 greater than 6.3 \AA\ in diameter (Fig. \ref{iCrystal}). This open
104 structure leads to crystals that are typically 0.07 g/cm$^3$ less
105 dense than ice I$_h$.
106
107 \begin{figure}
108 \centering
109 \includegraphics[width=\linewidth]{unitCell.eps}
110 \caption{(A) Unit cells for Ice-{\it i} and (B) Ice-{\it i}$^\prime$.
111 The spheres represent the center-of-mass locations of the water
112 molecules. The $a$ to $c$ ratios for Ice-{\it i} and Ice-{\it
113 i}$^\prime$ are given by 2.1214 and 1.785 respectively.}
114 \label{unitcell}
115 \end{figure}
116
117 \begin{figure}
118 \centering
119 \includegraphics[width=\linewidth]{orderedIcei.eps}
120 \caption{A rendering of a proton ordered crystal of Ice-{\it i} looking
121 down the (001) crystal face. The presence of large octagonal pores
122 leads to a polymorph that is less dense than ice I$_h$.}
123 \label{iCrystal}
124 \end{figure}
125
126 Results from our previous study indicated that Ice-{\it i} is the
127 minimum energy crystal structure for the single point water models
128 investigated (for discussions on these single point dipole models, see
129 our previous work and related
130 articles).\cite{Fennell04,Liu96,Bratko85} Our earlier results
131 considered only energetic stabilization and neglected entropic
132 contributions to the overall free energy. To address this issue, we
133 have calculated the absolute free energy of this crystal using
134 thermodynamic integration and compared it to the free energies of ice
135 I$_c$ and ice I$_h$ (the common low density ice polymorphs) and ice B
136 (a higher density, but very stable crystal structure observed by
137 B\`{a}ez and Clancy in free energy studies of SPC/E).\cite{Baez95b}
138 This work includes results for the water model from which Ice-{\it i}
139 was crystallized (SSD/E) in addition to several common water models
140 (TIP3P, TIP4P, TIP5P, and SPC/E) and a reaction field parametrized
141 single point dipole water model (SSD/RF). The axially-elongated
142 variant, Ice-{\it i}$^\prime$, was used in calculations involving
143 SPC/E, TIP4P, and TIP5P. The square tetramers in Ice-{\it i} distort
144 in Ice-{\it i}$^\prime$ to form a rhombus with alternating 85 and 95
145 degree angles. Under SPC/E, TIP4P, and TIP5P, this geometry is better
146 at forming favorable hydrogen bonds. The degree of rhomboid
147 distortion depends on the water model used, but is significant enough
148 to split a peak in the radial distribution function which corresponds
149 to diagonal sites in the tetramers.
150
151 \section{Methods}
152
153 Canonical ensemble (NVT) molecular dynamics calculations were
154 performed using the OOPSE molecular mechanics program.\cite{Meineke05}
155 All molecules were treated as rigid bodies, with orientational motion
156 propagated using the symplectic DLM integration method. Details about
157 the implementation of this technique can be found in a recent
158 publication.\cite{Dullweber1997}
159
160 Thermodynamic integration was utilized to calculate the Helmholtz free
161 energies ($A$) of the listed water models at various state points
162 using the OOPSE molecular dynamics program.\cite{Meineke05}
163 Thermodynamic integration is an established technique that has been
164 used extensively in the calculation of free energies for condensed
165 phases of
166 materials.\cite{Frenkel84,Hermens88,Meijer90,Baez95a,Vlot99}. This
167 method uses a sequence of simulations during which the system of
168 interest is converted into a reference system for which the free
169 energy is known analytically ($A_0$). The difference in potential
170 energy between the reference system and the system of interest
171 ($\Delta V$) is then integrated in order to determine the free energy
172 difference between the two states:
173 \begin{equation}
174 A = A_0 + \int_0^1 \left\langle \Delta V \right\rangle_\lambda d\lambda.
175 \end{equation}
176 Here, $\lambda$ is the parameter that governs the transformation
177 between the reference system and the system of interest. For
178 crystalline phases, an harmonically-restrained (Einsten) crystal is
179 chosen as the reference state, while for liquid phases, the ideal gas
180 is taken as the reference state.
181
182 In an Einstein crystal, the molecules are restrained at their ideal
183 lattice locations and orientations. Using harmonic restraints, as
184 applied by B\`{a}ez and Clancy, the total potential for this reference
185 crystal ($V_\mathrm{EC}$) is the sum of all the harmonic restraints,
186 \begin{equation}
187 V_\mathrm{EC} = \frac{K_\mathrm{v}r^2}{2} + \frac{K_\theta\theta^2}{2} +
188 \frac{K_\omega\omega^2}{2},
189 \end{equation}
190 where $K_\mathrm{v}$, $K_\mathrm{\theta}$, and $K_\mathrm{\omega}$ are
191 the spring constants restraining translational motion and deflection
192 of and rotation around the principle axis of the molecule
193 respectively. These spring constants are typically calculated from
194 the mean-square displacements of water molecules in an unrestrained
195 ice crystal at 200 K. For these studies, $K_\mathrm{v} = 4.29$ kcal
196 mol$^{-1}$ \AA$^{-2}$, $K_\theta\ = 13.88$ kcal mol$^{-1}$ rad$^{-2}$,
197 and $K_\omega\ = 17.75$ kcal mol$^{-1}$ rad$^{-2}$. It is clear from
198 Fig. \ref{waterSpring} that the values of $\theta$ range from $0$ to
199 $\pi$, while $\omega$ ranges from $-\pi$ to $\pi$. The partition
200 function for a molecular crystal restrained in this fashion can be
201 evaluated analytically, and the Helmholtz Free Energy ({\it A}) is
202 given by
203 \begin{eqnarray}
204 A = E_m\ -\ kT\ln \left (\frac{kT}{h\nu}\right )^3&-&kT\ln \left
205 [\pi^\frac{1}{2}\left (\frac{8\pi^2I_\mathrm{A}kT}{h^2}\right
206 )^\frac{1}{2}\left (\frac{8\pi^2I_\mathrm{B}kT}{h^2}\right
207 )^\frac{1}{2}\left (\frac{8\pi^2I_\mathrm{C}kT}{h^2}\right
208 )^\frac{1}{2}\right ] \nonumber \\ &-& kT\ln \left [\frac{kT}{2(\pi
209 K_\omega K_\theta)^{\frac{1}{2}}}\exp\left
210 (-\frac{kT}{2K_\theta}\right)\int_0^{\left (\frac{kT}{2K_\theta}\right
211 )^\frac{1}{2}}\exp(t^2)\mathrm{d}t\right ],
212 \label{ecFreeEnergy}
213 \end{eqnarray}
214 where $2\pi\nu = (K_\mathrm{v}/m)^{1/2}$, and $E_m$ is the minimum
215 potential energy of the ideal crystal.\cite{Baez95a}
216
217 \begin{figure}
218 \centering
219 \includegraphics[width=4in]{rotSpring.eps}
220 \caption{Possible orientational motions for a restrained molecule.
221 $\theta$ angles correspond to displacement from the body-frame {\it
222 z}-axis, while $\omega$ angles correspond to rotation about the
223 body-frame {\it z}-axis. $K_\theta$ and $K_\omega$ are spring
224 constants for the harmonic springs restraining motion in the $\theta$
225 and $\omega$ directions.}
226 \label{waterSpring}
227 \end{figure}
228
229 In the case of molecular liquids, the ideal vapor is chosen as the
230 target reference state. There are several examples of liquid state
231 free energy calculations of water models present in the
232 literature.\cite{Hermens88,Quintana92,Mezei92,Baez95b} These methods
233 typically differ in regard to the path taken for switching off the
234 interaction potential to convert the system to an ideal gas of water
235 molecules. In this study, we applied one of the most convenient
236 methods and integrated over the $\lambda^4$ path, where all
237 interaction parameters are scaled equally by this transformation
238 parameter. This method has been shown to be reversible and provide
239 results in excellent agreement with other established
240 methods.\cite{Baez95b}
241
242 Near the cutoff radius ($0.85 * r_{cut}$), charge, dipole, and
243 Lennard-Jones interactions were gradually reduced by a cubic switching
244 function. By applying this function, these interactions are smoothly
245 truncated, thereby avoiding the poor energy conservation which results
246 from harsher truncation schemes. The effect of a long-range
247 correction was also investigated on select model systems in a variety
248 of manners. For the SSD/RF model, a reaction field with a fixed
249 dielectric constant of 80 was applied in all
250 simulations.\cite{Onsager36} For a series of the least computationally
251 expensive models (SSD/E, SSD/RF, TIP3P, and SPC/E), simulations were
252 performed with longer cutoffs of 10.5, 12, 13.5, and 15 \AA\ to
253 compare with the 9 \AA\ cutoff results. Finally, the effects of using
254 the Ewald summation were estimated for TIP3P and SPC/E by performing
255 single configuration Particle-Mesh Ewald (PME)
256 calculations~\cite{Tinker} for each of the ice polymorphs. The
257 calculated energy difference in the presence and absence of PME was
258 applied to the previous results in order to predict changes to the
259 free energy landscape.
260
261 \section{Results and Discussion}
262
263 The calculated free energies of proton-ordered variants of three low
264 density polymorphs (I$_h$, I$_c$, and Ice-{\it i} or Ice-{\it
265 i}$^\prime$) and the stable higher density ice B are listed in Table
266 \ref{freeEnergy}. Ice B was included because it has been
267 shown to be a minimum free energy structure for SPC/E at ambient
268 conditions.\cite{Baez95b} In addition to the free energies, the
269 relevant transition temperatures at standard pressure are also
270 displayed in Table \ref{freeEnergy}. These free energy values
271 indicate that Ice-{\it i} is the most stable state for all of the
272 investigated water models. With the free energy at these state
273 points, the Gibbs-Helmholtz equation was used to project to other
274 state points and to build phase diagrams. Figure \ref{tp3PhaseDia} is
275 an example diagram built from the results for the TIP3P water model.
276 All other models have similar structure, although the crossing points
277 between the phases move to different temperatures and pressures as
278 indicated from the transition temperatures in Table \ref{freeEnergy}.
279 It is interesting to note that ice I$_h$ (and ice I$_c$ for that
280 matter) do not appear in any of the phase diagrams for any of the
281 models. For purposes of this study, ice B is representative of the
282 dense ice polymorphs. A recent study by Sanz {\it et al.} provides
283 details on the phase diagrams for SPC/E and TIP4P at higher pressures
284 than those studied here.\cite{Sanz04}
285
286 \begin{table*}
287 \begin{minipage}{\linewidth}
288 \begin{center}
289 \caption{Calculated free energies for several ice polymorphs along
290 with the calculated melting (or sublimation) and boiling points for
291 the investigated water models. All free energy calculations used a
292 cutoff radius of 9.0 \AA\ and were performed at 200 K and $\sim$1 atm.
293 Units of free energy are kcal/mol, while transition temperature are in
294 Kelvin. Calculated error of the final digits is in parentheses.}
295 \begin{tabular}{lccccccc}
296 \hline
297 Water Model & I$_h$ & I$_c$ & B & Ice-{\it i} & Ice-{\it i}$^\prime$ & $T_m$ (*$T_s$) & $T_b$\\
298 \hline
299 TIP3P & -11.41(2) & -11.23(3) & -11.82(3) & -12.30(3) & - & 269(7) & 357(4)\\
300 TIP4P & -11.84(3) & -12.04(2) & -12.08(3) & - & -12.33(3) & 262(6) & 354(4)\\
301 TIP5P & -11.85(3) & -11.86(2) & -11.96(2) & - & -12.29(2) & 266(7) & 337(4)\\
302 SPC/E & -12.87(2) & -13.05(2) & -13.26(3) & - & -13.55(2) & 299(6) & 396(4)\\
303 SSD/E & -11.27(2) & -11.19(4) & -12.09(2) & -12.54(2) & - & *355(4) & -\\
304 SSD/RF & -11.96(2) & -11.60(2) & -12.53(3) & -12.79(2) & - & 278(7) & 382(4)\\
305 \end{tabular}
306 \label{freeEnergy}
307 \end{center}
308 \end{minipage}
309 \end{table*}
310
311 \begin{figure}
312 \includegraphics[width=\linewidth]{tp3PhaseDia.eps}
313 \caption{Phase diagram for the TIP3P water model in the low pressure
314 regime. The displayed $T_m$ and $T_b$ values are good predictions of
315 the experimental values; however, the solid phases shown are not the
316 experimentally observed forms. Both cubic and hexagonal ice $I$ are
317 higher in energy and don't appear in the phase diagram.}
318 \label{tp3PhaseDia}
319 \end{figure}
320
321 Most of the water models have melting points that compare quite
322 favorably with the experimental value of 273 K. The unfortunate
323 aspect of this result is that this phase change occurs between
324 Ice-{\it i} and the liquid state rather than ice I$_h$ and the liquid
325 state. These results do not contradict other studies. Studies of ice
326 I$_h$ using TIP4P predict a $T_m$ ranging from 191 to 238 K
327 (differences being attributed to choice of interaction truncation and
328 different ordered and disordered molecular
329 arrangements).\cite{Nada03,Vlot99,Gao00,Sanz04} If the presence of ice B and
330 Ice-{\it i} were omitted, a $T_m$ value around 200 K would be
331 predicted from this work. However, the $T_m$ from Ice-{\it i} is
332 calculated to be 262 K, indicating that these simulation based
333 structures ought to be included in studies probing phase transitions
334 with this model. Also of interest in these results is that SSD/E does
335 not exhibit a melting point at 1 atm but does sublime at 355 K. This
336 is due to the significant stability of Ice-{\it i} over all other
337 polymorphs for this particular model under these conditions. While
338 troubling, this behavior resulted in the spontaneous crystallization
339 of Ice-{\it i} which led us to investigate this structure. These
340 observations provide a warning that simulations of SSD/E as a
341 ``liquid'' near 300 K are actually metastable and run the risk of
342 spontaneous crystallization. However, when a longer cutoff radius is
343 used, SSD/E prefers the liquid state under standard temperature and
344 pressure.
345
346 \begin{figure}
347 \includegraphics[width=\linewidth]{cutoffChange.eps}
348 \caption{Free energy as a function of cutoff radius for SSD/E, TIP3P,
349 SPC/E, SSD/RF with a reaction field, and the TIP3P and SPC/E models
350 with an added Ewald correction term. Error for the larger cutoff
351 points is equivalent to that observed at 9.0\AA\ (see Table
352 \ref{freeEnergy}). Data for ice I$_c$ with TIP3P using both 12 and
353 13.5 \AA\ cutoffs were omitted because the crystal was prone to
354 distortion and melting at 200 K. Ice-{\it i}$^\prime$ is the form of
355 Ice-{\it i} used in the SPC/E simulations.}
356 \label{incCutoff}
357 \end{figure}
358
359 For the more computationally efficient water models, we have also
360 investigated the effect of potential trunctaion on the computed free
361 energies as a function of the cutoff radius. As seen in
362 Fig. \ref{incCutoff}, the free energies of the ice polymorphs with
363 water models lacking a long-range correction show significant cutoff
364 dependence. In general, there is a narrowing of the free energy
365 differences while moving to greater cutoff radii. As the free
366 energies for the polymorphs converge, the stability advantage that
367 Ice-{\it i} exhibits is reduced. Adjacent to each of these plots are
368 results for systems with applied or estimated long-range corrections.
369 SSD/RF was parametrized for use with a reaction field, and the benefit
370 provided by this computationally inexpensive correction is apparent.
371 The free energies are largely independent of the size of the reaction
372 field cavity in this model, so small cutoff radii mimic bulk
373 calculations quite well under SSD/RF.
374
375 Although TIP3P was paramaterized for use without the Ewald summation,
376 we have estimated the effect of this method for computing long-range
377 electrostatics for both TIP3P and SPC/E. This was accomplished by
378 calculating the potential energy of identical crystals both with and
379 without particle mesh Ewald (PME). Similar behavior to that observed
380 with reaction field is seen for both of these models. The free
381 energies show reduced dependence on cutoff radius and span a narrower
382 range for the various polymorphs. Like the dipolar water models,
383 TIP3P displays a relatively constant preference for the Ice-{\it i}
384 polymorph. Crystal preference is much more difficult to determine for
385 SPC/E. Without a long-range correction, each of the polymorphs
386 studied assumes the role of the preferred polymorph under different
387 cutoff radii. The inclusion of the Ewald correction flattens and
388 narrows the gap in free energies such that the polymorphs are
389 isoenergetic within statistical uncertainty. This suggests that other
390 conditions, such as the density in fixed-volume simulations, can
391 influence the polymorph expressed upon crystallization.
392
393 \section{Conclusions}
394
395 In this work, thermodynamic integration was used to determine the
396 absolute free energies of several ice polymorphs. The new polymorph,
397 Ice-{\it i} was observed to be the stable crystalline state for {\it
398 all} the water models when using a 9.0 \AA\ cutoff. However, the free
399 energy partially depends on simulation conditions (particularly on the
400 choice of long range correction method). Regardless, Ice-{\it i} was
401 still observered to be a stable polymorph for all of the studied water
402 models.
403
404 So what is the preferred solid polymorph for simulated water? As
405 indicated above, the answer appears to be dependent both on the
406 conditions and the model used. In the case of short cutoffs without a
407 long-range interaction correction, Ice-{\it i} and Ice-{\it
408 i}$^\prime$ have the lowest free energy of the studied polymorphs with
409 all the models. Ideally, crystallization of each model under constant
410 pressure conditions, as was done with SSD/E, would aid in the
411 identification of their respective preferred structures. This work,
412 however, helps illustrate how studies involving one specific model can
413 lead to insight about important behavior of others.
414
415 We also note that none of the water models used in this study are
416 polarizable or flexible models. It is entirely possible that the
417 polarizability of real water makes Ice-{\it i} substantially less
418 stable than ice I$_h$. However, the calculations presented above seem
419 interesting enough to communicate before the role of polarizability
420 (or flexibility) has been thoroughly investigated.
421
422 Finally, due to the stability of Ice-{\it i} in the investigated
423 simulation conditions, the question arises as to possible experimental
424 observation of this polymorph. The rather extensive past and current
425 experimental investigation of water in the low pressure regime makes
426 us hesitant to ascribe any relevance to this work outside of the
427 simulation community. It is for this reason that we chose a name for
428 this polymorph which involves an imaginary quantity. That said, there
429 are certain experimental conditions that would provide the most ideal
430 situation for possible observation. These include the negative
431 pressure or stretched solid regime, small clusters in vacuum
432 deposition environments, and in clathrate structures involving small
433 non-polar molecules. For the purpose of comparison with experimental
434 results, we have calculated the oxygen-oxygen pair correlation
435 function, $g_{OO}(r)$, and the structure factor, $S(\vec{q})$ for the
436 two Ice-{\it i} variants (along with example ice I$_h$ and I$_c$
437 plots) at 77K, and they are shown in figures \ref{fig:gofr} and
438 \ref{fig:sofq} respectively. It is interesting to note that the
439 structure factors for Ice-{\it i}$^\prime$ and Ice-I$_c$ are quite similar.
440 The primary differences are small peaks at 1.125, 2.29, and 2.53
441 \AA${-1}$, so particular attention to these regions would be needed
442 to identify the new {\it i}$^\prime$ variant from the I$_{c}$ variant.
443
444 \begin{figure}
445 \centering
446 \includegraphics[width=\linewidth]{iceGofr.eps}
447 \caption{Radial distribution functions of ice I$_h$, I$_c$, and
448 Ice-{\it i} calculated from from simulations of the SSD/RF water model
449 at 77 K. The Ice-{\it i} distribution function was obtained from
450 simulations composed of TIP4P water.}
451 \label{fig:gofr}
452 \end{figure}
453
454 \begin{figure}
455 \centering
456 \includegraphics[width=\linewidth]{sofq.eps}
457 \caption{Predicted structure factors for ice I$_h$, I$_c$, Ice-{\it i},
458 and Ice-{\it i}$^\prime$ at 77 K. The raw structure factors have
459 been convoluted with a gaussian instrument function (0.075 \AA$^{-1}$
460 width) to compensate for the trunction effects in our finite size
461 simulations.}
462 \label{fig:sofq}
463 \end{figure}
464
465 \section{Acknowledgments}
466 Support for this project was provided by the National Science
467 Foundation under grant CHE-0134881. Computation time was provided by
468 the Notre Dame High Performance Computing Cluster and the Notre Dame
469 Bunch-of-Boxes (B.o.B) computer cluster (NSF grant DMR-0079647).
470
471 \newpage
472
473 \bibliographystyle{achemso}
474 \bibliography{iceiPaper}
475
476
477 \end{document}