ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/fennellDissertation/iceChapter.tex
Revision: 2979
Committed: Tue Aug 29 00:40:05 2006 UTC (17 years, 10 months ago) by chrisfen
Content type: application/x-tex
File size: 25640 byte(s)
Log Message:
more changes

File Contents

# User Rev Content
1 chrisfen 2977 \chapter{\label{chap:ice}PHASE BEHAVIOR OF WATER IN COMPUTER SIMULATIONS}
2    
3 chrisfen 2979 As discussed in the previous chapter, water has proven to be a
4     challenging substance to depict in simulations, and a variety of
5     models have been developed to describe its behavior under varying
6     simulation
7 chrisfen 2978 conditions.\cite{Stillinger74,Rahman75,Berendsen81,Jorgensen83,Bratko85,Berendsen87,Caldwell95,Liu96,vanderSpoel98,Urbic00,Mahoney00,Fennell04}
8 chrisfen 2977 These models have been used to investigate important physical
9     phenomena like phase transitions and the hydrophobic
10 chrisfen 2978 effect.\cite{Yamada02,Marrink94,Gallagher03} With the choice of models
11     available, it is only natural to compare the models under interesting
12     thermodynamic conditions in an attempt to clarify the limitations of
13 chrisfen 2979 each of the models.\cite{Jorgensen83,Jorgensen98b,Baez94,Mahoney01}
14     Two important property to quantify are the Gibbs and Helmholtz free
15 chrisfen 2978 energies, particularly for the solid forms of water, as these predict
16     the thermodynamic stability of the various phases. Water has a
17     particularly rich phase diagram and takes on a number of different and
18     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 chrisfen 2977 focused on the phases having the lowest free energy at a given state
22 chrisfen 2978 point, because these phases will dictate the relevant transition
23     temperatures and pressures for the model.
24 chrisfen 2977
25 chrisfen 2978 The high-pressure phases of water (ice II-ice X as well as ice XII)
26     have been studied extensively both experimentally and
27 chrisfen 2979 computationally. In this chapter, standard reference state methods
28 chrisfen 2978 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 chrisfen 2979 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 235K, two
38     of five systems recrystallized near 245K. 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     simulation.\cite{Fennell04} We have named this structure Ice-{\it i}
43     to indicate its origin in computational simulation. The unit cell of
44     Ice-$i$ and an axially elongated variant named Ice-$i^\prime$ both
45     consist of eight water molecules that stack in rows of interlocking
46     water tetramers as illustrated in figure \ref{fig:iceiCell}A,B. These
47     tetramers form a crystal structure similar in appearance to a recent
48     two-dimensional surface tessellation simulated on silica.\cite{Yang04}
49     As expected in an ice crystal constructed of water tetramers, the
50     hydrogen bonds are not as linear as those observed in ice
51     I$_\textrm{h}$; however, the interlocking of these subunits appears to
52     provide significant stabilization to the overall crystal. The
53     arrangement of these tetramers results in open octagonal cavities that
54     are typically greater than 6.3\AA\ in diameter (see figure
55 chrisfen 2978 \ref{fig:protOrder}). This open structure leads to crystals that are
56     typically 0.07 g/cm$^3$ less dense than ice I$_\textrm{h}$.
57 chrisfen 2977
58     \begin{figure}
59     \includegraphics[width=\linewidth]{./figures/unitCell.pdf}
60     \caption{Unit cells for (A) Ice-{\it i} and (B) Ice-$i^\prime$, the
61     elongated variant of Ice-{\it i}. For Ice-{\it i}, the $a$ to $c$
62     relation is given by $a = 2.1214c$, while for Ice-$i^\prime$, $a =
63     1.7850c$.}
64 chrisfen 2978 \label{fig:iceiCell}
65 chrisfen 2977 \end{figure}
66    
67     \begin{figure}
68     \centering
69     \includegraphics[width=3.5in]{./figures/orderedIcei.pdf}
70     \caption{Image of a proton ordered crystal of Ice-{\it i} looking
71     down the (001) crystal face. The rows of water tetramers surrounded by
72     octagonal pores leads to a crystal structure that is significantly
73 chrisfen 2978 less dense than ice I$_\textrm{h}$.}
74     \label{fig:protOrder}
75 chrisfen 2977 \end{figure}
76    
77 chrisfen 2979 Results from our initial studies indicated that Ice-{\it i} is the
78 chrisfen 2978 minimum energy crystal structure for the single point water models
79 chrisfen 2977 investigated (for discussions on these single point dipole models, see
80     the previous work and related
81 chrisfen 2979 articles\cite{Fennell04,Liu96,Bratko85}). These earlier results only
82 chrisfen 2977 considered energetic stabilization and neglected entropic
83 chrisfen 2978 contributions to the overall free energy. To address this issue, we
84     have calculated the absolute free energy of this crystal using
85     thermodynamic integration and compared to the free energies of ice
86     I$_\textrm{c}$ and ice I$_\textrm{h}$ (the common low-density ice
87     polymorphs) and ice B (a higher density, but very stable crystal
88     structure observed by B\`{a}ez and Clancy in free energy studies of
89 chrisfen 2977 SPC/E).\cite{Baez95b} This work includes results for the water model
90     from which Ice-{\it i} was crystallized (SSD/E) in addition to several
91     common water models (TIP3P, TIP4P, TIP5P, and SPC/E) and a reaction
92 chrisfen 2978 field parametrized single point dipole water model (SSD/RF). The
93     axially elongated variant, Ice-$i^\prime$, was used in calculations
94     involving SPC/E, TIP4P, and TIP5P. The square tetramer in Ice-$i$
95     distorts in Ice-$i^\prime$ to form a rhombus with alternating 85 and
96     95 degree angles. Under SPC/E, TIP4P, and TIP5P, this geometry is
97     better at forming favorable hydrogen bonds. The degree of rhomboid
98     distortion depends on the water model used but is significant enough
99     to split the peak in the radial distribution function which corresponds
100     to diagonal sites in the tetramers.
101 chrisfen 2977
102 chrisfen 2979 \section{Methods and Thermodynamic Integration}
103 chrisfen 2977
104 chrisfen 2978 Canonical ensemble ({\it NVT}) molecular dynamics calculations were
105 chrisfen 2977 performed using the OOPSE molecular mechanics package.\cite{Meineke05}
106 chrisfen 2978 The densities chosen for the simulations were taken from
107     isobaric-isothermal ({\it NPT}) simulations performed at 1 atm and
108     200K. Each model (and each crystal structure) was allowed to relax for
109     300ps in the {\it NPT} ensemble before averaging the density to obtain
110     the volumes for the {\it NVT} simulations.All molecules were treated
111     as rigid bodies, with orientational motion propagated using the
112     symplectic DLM integration method described in section
113 chrisfen 2979 \ref{sec:IntroIntegrate}.
114 chrisfen 2977
115 chrisfen 2979
116 chrisfen 2978 We used thermodynamic integration to calculate the Helmholtz free
117     energies ({\it A}) of the listed water models at various state
118     points. Thermodynamic integration is an established technique that has
119     been used extensively in the calculation of free energies for
120     condensed phases of
121 chrisfen 2979 materials.\cite{Frenkel84,Hermens88,Meijer90,Baez95a,Vlot99}. This
122     method uses a sequence of simulations over which the system of
123 chrisfen 2978 interest is converted into a reference system for which the free
124 chrisfen 2979 energy is known analytically ($A_0$). The difference in potential
125     energy between the reference system and the system of interest
126     ($\Delta V$) is then integrated in order to determine the free energy
127     difference between the two states:
128 chrisfen 2977 \begin{equation}
129 chrisfen 2979 A = A_0 + \int_0^1 \left\langle \Delta V \right\rangle_\lambda d\lambda.
130 chrisfen 2977 \end{equation}
131 chrisfen 2979 Here, $\lambda$ is the parameter that governs the transformation
132     between the reference system and the system of interest. For
133     crystalline phases, an harmonically-restrained (Einstein) crystal is
134     chosen as the reference state, while for liquid phases, the ideal gas
135     is taken as the reference state. Figure \ref{fig:integrationPath}
136     shows an example integration path for converting a crystalline system
137     to the Einstein crystal reference state.
138     \begin{figure}
139     \includegraphics[width=\linewidth]{./figures/integrationPath.pdf}
140     \caption{An example integration path to convert an unrestrained
141     crystal ($\lambda = 1$) to the Einstein crystal reference state
142     ($\lambda = 0$). Note the increase in samples at either end of the
143     path to improve the smoothness of the curve. For reversible processes,
144     conversion of the Einstein crystal back to the system of interest will
145     give an identical plot, thereby integrating to the same result.}
146     \label{fig:integrationPath}
147     \end{figure}
148 chrisfen 2977
149 chrisfen 2979 In an Einstein crystal, the molecules are restrained at their ideal
150     lattice locations and orientations. Using harmonic restraints, as
151     applied by B\'{a}ez and Clancy, the total potential for this reference
152     crystal ($V_\mathrm{EC}$) is the sum of all the harmonic restraints,
153     \begin{equation}
154     V_\mathrm{EC} = \frac{K_\mathrm{v}r^2}{2} + \frac{K_\theta\theta^2}{2} +
155     \frac{K_\omega\omega^2}{2},
156     \end{equation}
157     where $K_\mathrm{v}$, $K_\mathrm{\theta}$, and $K_\mathrm{\omega}$ are
158     the spring constants restraining translational motion and deflection
159     of and rotation around the principle axis of the molecule
160     respectively. These spring constants are typically calculated from
161     the mean-square displacements of water molecules in an unrestrained
162     ice crystal at 200 K. For these studies, $K_\mathrm{v} = 4.29$ kcal
163     mol$^{-1}$ \AA$^{-2}$, $K_\theta\ = 13.88$ kcal mol$^{-1}$ rad$^{-2}$,
164     and $K_\omega\ = 17.75$ kcal mol$^{-1}$ rad$^{-2}$. It is clear from
165     Fig. \ref{fig:waterSpring} that the values of $\theta$ range from $0$ to
166     $\pi$, while $\omega$ ranges from $-\pi$ to $\pi$. The partition
167     function for a molecular crystal restrained in this fashion can be
168     evaluated analytically, and the Helmholtz Free Energy ({\it A}) is
169     given by
170     \begin{equation}
171     \begin{split}
172     A = E_m &- kT\ln\left(\frac{kT}{h\nu}\right)^3 \\
173     &- kT\ln\left[\pi^\frac{1}{2}\left(
174     \frac{8\pi^2I_\mathrm{A}kT}{h^2}\right)^\frac{1}{2}
175     \left(\frac{8\pi^2I_\mathrm{B}kT}{h^2}\right)^\frac{1}{2}
176     \left(\frac{8\pi^2I_\mathrm{C}kT}{h^2}\right)^\frac{1}{2}
177     \right] \\
178     &- kT\ln\left[\frac{kT}{2(\pi K_\omega K_\theta)^{\frac{1}{2}}}
179     \exp\left(-\frac{kT}{2K_\theta}\right)
180     \int_0^{\left(\frac{kT}{2K_\theta}\right)^\frac{1}{2}}
181     \exp(t^2)\mathrm{d}t\right],
182     \end{split}
183 chrisfen 2978 \label{eq:ecFreeEnergy}
184 chrisfen 2979 \end{equation}
185     where $2\pi\nu = (K_\mathrm{v}/m)^{1/2}$, and $E_m$ is the minimum
186     potential energy of the ideal crystal.\cite{Baez95a} The choice of an
187     Einstein crystal reference stat is somewhat arbitrary. Any ideal
188     system for which the partition function is known exactly could be used
189     as a reference point as long as the system does not undergo a phase
190     transition during the integration path between the real and ideal
191     systems. Nada and van der Eerden have shown that the use of different
192     force constants in the Einstein crystal doesn not affect the total
193     free energy, and Gao {\it et al.} have shown that free energies
194     computed with the Debye crystal reference state differ from the
195     Einstein crystal by only a few tenths of a kJ
196     mol$^{-1}$.\cite{Nada03,Gao00} These free energy differences can lead
197     to some uncertainty in the computed melting point of the solids.
198 chrisfen 2977 \begin{figure}
199     \centering
200     \includegraphics[width=3.5in]{./figures/rotSpring.pdf}
201     \caption{Possible orientational motions for a restrained molecule.
202     $\theta$ angles correspond to displacement from the body-frame {\it
203     z}-axis, while $\omega$ angles correspond to rotation about the
204     body-frame {\it z}-axis. $K_\theta$ and $K_\omega$ are spring
205     constants for the harmonic springs restraining motion in the $\theta$
206     and $\omega$ directions.}
207 chrisfen 2979 \label{fig:waterSpring}
208 chrisfen 2977 \end{figure}
209    
210 chrisfen 2979 In the case of molecular liquids, the ideal vapor is chosen as the
211     target reference state. There are several examples of liquid state
212     free energy calculations of water models present in the
213     literature.\cite{Hermens88,Quintana92,Mezei92,Baez95b} These methods
214     typically differ in regard to the path taken for switching off the
215     interaction potential to convert the system to an ideal gas of water
216     molecules. In this study, we applied one of the most convenient
217     methods and integrated over the $\lambda^4$ path, where all
218     interaction parameters are scaled equally by this transformation
219     parameter. This method has been shown to be reversible and provide
220     results in excellent agreement with other established
221     methods.\cite{Baez95b}
222 chrisfen 2977
223 chrisfen 2979 Near the cutoff radius ($0.85 * r_{cut}$), charge, dipole, and
224     Lennard-Jones interactions were gradually reduced by a cubic switching
225     function. By applying this function, these interactions are smoothly
226     truncated, thereby avoiding the poor energy conservation which results
227     from harsher truncation schemes. The effect of a long-range
228     correction was also investigated on select model systems in a variety
229     of manners. For the SSD/RF model, a reaction field with a fixed
230     dielectric constant of 80 was applied in all
231     simulations.\cite{Onsager36} For a series of the least computationally
232     expensive models (SSD/E, SSD/RF, TIP3P, and SPC/E), simulations were
233     performed with longer cutoffs of 10.5, 12, 13.5, and 15\AA\ to
234     compare with the 9\AA\ cutoff results. Finally, the effects of using
235     the Ewald summation were estimated for TIP3P and SPC/E by performing
236     single configuration Particle-Mesh Ewald (PME) calculations for each
237     of the ice polymorphs.\cite{Ponder87} The calculated energy difference
238     in the presence and absence of PME was applied to the previous results
239     in order to predict changes to the free energy landscape.
240 chrisfen 2977
241 chrisfen 2979 \section{Initial Free Energy Results}
242 chrisfen 2977
243 chrisfen 2979 The calculated free energies of proton-ordered variants of three low
244     density polymorphs (I$_\textrm{h}$, I$_\textrm{c}$, and Ice-{\it i} or
245     Ice-$i^\prime$) and the stable higher density ice B are listed in
246     table \ref{tab:freeEnergy}. Ice B was included because it has been
247     shown to be a minimum free energy structure for SPC/E at ambient
248     conditions.\cite{Baez95b} In addition to the free energies, the
249     relevant transition temperatures at standard pressure are also
250     displayed in table \ref{tab:freeEnergy}. These free energy values
251     indicate that Ice-{\it i} is the most stable state for all of the
252     investigated water models. With the free energy at these state
253     points, the Gibbs-Helmholtz equation was used to project to other
254     state points and to build phase diagrams. Figures
255     \ref{fig:tp3PhaseDia} and \ref{fig:ssdrfPhaseDia} are example diagrams
256     built from the results for the TIP3P and SSD/RF water models. All
257     other models have similar structure, although the crossing points
258     between the phases move to different temperatures and pressures as
259     indicated from the transition temperatures in table
260     \ref{tab:freeEnergy}. It is interesting to note that ice
261     I$_\textrm{h}$ (and ice I$_\textrm{c}$ for that matter) do not appear
262     in any of the phase diagrams for any of the models. For purposes of
263     this study, ice B is representative of the dense ice polymorphs. A
264     recent study by Sanz {\it et al.} provides details on the phase
265     diagrams for SPC/E and TIP4P at higher pressures than those studied
266     here.\cite{Sanz04}
267 chrisfen 2978 \begin{table}
268     \centering
269 chrisfen 2979 \caption{HELMHOLTZ FREE ENERGIES AND TRANSITION TEMPERATURES AT 1
270     ATMOSPHERE FOR SEVERAL WATER MODELS}
271    
272     \footnotesize
273     \begin{tabular}{lccccccc}
274 chrisfen 2978 \toprule
275     \toprule
276 chrisfen 2979 Water Model & I$_\textrm{h}$ & I$_\textrm{c}$ & B & Ice-{\it i} & Ice-$i^\prime$ & $T_\textrm{m}$ (*$T_\textrm{s}$) & $T_\textrm{b}$\\
277     \cmidrule(lr){2-6}
278     \cmidrule(l){7-8}
279     & \multicolumn{5}{c}{(kcal mol$^{-1}$)} & \multicolumn{2}{c}{(K)}\\
280 chrisfen 2978 \midrule
281 chrisfen 2979 TIP3P & -11.41(2) & -11.23(3) & -11.82(3) & -12.30(3) & - & 269(7) & 357(4)\\
282     TIP4P & -11.84(3) & -12.04(2) & -12.08(3) & - & -12.33(3) & 262(6) & 354(4)\\
283     TIP5P & -11.85(3) & -11.86(2) & -11.96(2) & - & -12.29(2) & 266(7) & 337(4)\\
284     SPC/E & -12.87(2) & -13.05(2) & -13.26(3) & - & -13.55(2) & 299(6) & 396(4)\\
285     SSD/E & -11.27(2) & -11.19(4) & -12.09(2) & -12.54(2) & - & *355(4) & -\\
286     SSD/RF & -11.96(2) & -11.60(2) & -12.53(3) & -12.79(2) & - & 278(7) & 382(4)\\
287 chrisfen 2978 \bottomrule
288 chrisfen 2977 \end{tabular}
289 chrisfen 2978 \label{tab:freeEnergy}
290     \end{table}
291 chrisfen 2977
292     \begin{figure}
293     \centering
294     \includegraphics[width=\linewidth]{./figures/tp3PhaseDia.pdf}
295     \caption{Phase diagram for the TIP3P water model in the low pressure
296     regime. The displayed $T_m$ and $T_b$ values are good predictions of
297     the experimental values; however, the solid phases shown are not the
298     experimentally observed forms. Both cubic and hexagonal ice $I$ are
299     higher in energy and don't appear in the phase diagram.}
300 chrisfen 2979 \label{fig:tp3PhaseDia}
301 chrisfen 2977 \end{figure}
302    
303     \begin{figure}
304     \centering
305     \includegraphics[width=\linewidth]{./figures/ssdrfPhaseDia.pdf}
306     \caption{Phase diagram for the SSD/RF water model in the low pressure
307     regime. Calculations producing these results were done under an
308     applied reaction field. It is interesting to note that this
309     computationally efficient model (over 3 times more efficient than
310     TIP3P) exhibits phase behavior similar to the less computationally
311     conservative charge based models.}
312 chrisfen 2979 \label{fig:ssdrfPhaseDia}
313 chrisfen 2977 \end{figure}
314    
315 chrisfen 2979 We note that all of the crystals investigated in this study ar ideal
316     proton-ordered antiferroelectric structures. All of the structures
317     obey the Bernal-Fowler rules and should be able to form stable
318     proton-{\it disordered} crystals which have the traditional
319     $k_\textrm{B}$ln(3/2) residual entropy at 0K.\cite{Bernal33,Pauling35}
320     Simulations of proton-disordered structures are relatively unstable
321     with all but the most expensive water models.\cite{Nada03} Our
322     simulations have therefore been performed with the ordered
323     antiferroelectric structures which do not require the residual entropy
324     term to be accounted for in the free energies. This may result in some
325     discrepancies when comparing our melting temperatures to the melting
326     temperatures that have been calculated via thermodynamic integrations
327     of the disordered structures.\cite{Sanz04}
328 chrisfen 2977
329 chrisfen 2979 Most of the water models have melting points that compare quite
330     favorably with the experimental value of 273 K. The unfortunate
331     aspect of this result is that this phase change occurs between
332     Ice-{\it i} and the liquid state rather than ice I$_h$ and the liquid
333     state. These results do not contradict other studies. Studies of ice
334     I$_h$ using TIP4P predict a $T_m$ ranging from 191 to 238 K
335 chrisfen 2978 (differences being attributed to choice of interaction truncation and
336     different ordered and disordered molecular
337 chrisfen 2979 arrangements).\cite{Nada03,Vlot99,Gao00,Sanz04} If the presence of ice
338     B and Ice-{\it i} were omitted, a $T_\textrm{m}$ value around 200 K
339     would be predicted from this work. However, the $T_\textrm{m}$ from
340     Ice-{\it i} is calculated to be 262 K, indicating that these
341     simulation based structures ought to be included in studies probing
342     phase transitions with this model. Also of interest in these results
343     is that SSD/E does not exhibit a melting point at 1 atm but does
344     sublime at 355 K. This is due to the significant stability of
345     Ice-{\it i} over all other polymorphs for this particular model under
346     these conditions. While troubling, this behavior resulted in the
347     spontaneous crystallization of Ice-{\it i} which led us to investigate
348     this structure. These observations provide a warning that simulations
349     of SSD/E as a ``liquid'' near 300 K are actually metastable and run
350     the risk of spontaneous crystallization. However, when a longer
351     cutoff radius is used, SSD/E prefers the liquid state under standard
352     temperature and pressure.
353 chrisfen 2977
354 chrisfen 2979 \section{Effects of Potential Trucation}
355    
356 chrisfen 2977 \begin{figure}
357     \includegraphics[width=\linewidth]{./figures/cutoffChange.pdf}
358 chrisfen 2979 \caption{Free energy as a function of cutoff radius for SSD/E, TIP3P,
359     SPC/E, SSD/RF with a reaction field, and the TIP3P and SPC/E models
360     with an added Ewald correction term. Error for the larger cutoff
361     points is equivalent to that observed at 9.0\AA\ (see Table
362     \ref{tab:freeEnergy}). Data for ice I$_\textrm{c}$ with TIP3P using
363     both 12 and 13.5\AA\ cutoffs were omitted because the crystal was
364     prone to distortion and melting at 200K. Ice-$i^\prime$ is the
365     form of Ice-{\it i} used in the SPC/E simulations.}
366 chrisfen 2978 \label{fig:incCutoff}
367 chrisfen 2977 \end{figure}
368    
369 chrisfen 2979 For the more computationally efficient water models, we have also
370     investigated the effect of potential trunctaion on the computed free
371     energies as a function of the cutoff radius. As seen in
372     Fig. \ref{fig:incCutoff}, the free energies of the ice polymorphs with
373     water models lacking a long-range correction show significant cutoff
374     dependence. In general, there is a narrowing of the free energy
375     differences while moving to greater cutoff radii. As the free
376     energies for the polymorphs converge, the stability advantage that
377     Ice-{\it i} exhibits is reduced. Adjacent to each of these plots are
378     results for systems with applied or estimated long-range corrections.
379     SSD/RF was parametrized for use with a reaction field, and the benefit
380     provided by this computationally inexpensive correction is apparent.
381     The free energies are largely independent of the size of the reaction
382     field cavity in this model, so small cutoff radii mimic bulk
383     calculations quite well under SSD/RF.
384    
385     Although TIP3P was paramaterized for use without the Ewald summation,
386     we have estimated the effect of this method for computing long-range
387     electrostatics for both TIP3P and SPC/E. This was accomplished by
388     calculating the potential energy of identical crystals both with and
389     without particle mesh Ewald (PME). Similar behavior to that observed
390     with reaction field is seen for both of these models. The free
391     energies show reduced dependence on cutoff radius and span a narrower
392     range for the various polymorphs. Like the dipolar water models,
393     TIP3P displays a relatively constant preference for the Ice-{\it i}
394     polymorph. Crystal preference is much more difficult to determine for
395     SPC/E. Without a long-range correction, each of the polymorphs
396     studied assumes the role of the preferred polymorph under different
397     cutoff radii. The inclusion of the Ewald correction flattens and
398     narrows the gap in free energies such that the polymorphs are
399     isoenergetic within statistical uncertainty. This suggests that other
400     conditions, such as the density in fixed-volume simulations, can
401     influence the polymorph expressed upon crystallization.
402 chrisfen 2977
403 chrisfen 2979 \section{Expanded Results Using Damped Shifted Force Electrostatics}
404 chrisfen 2977
405    
406     \section{Conclusions}
407    
408 chrisfen 2979 In this work, thermodynamic integration was used to determine the
409     absolute free energies of several ice polymorphs. The new polymorph,
410     Ice-{\it i} was observed to be the stable crystalline state for {\it
411     all} the water models when using a 9.0\AA\ cutoff. However, the free
412     energy partially depends on simulation conditions (particularly on the
413     choice of long range correction method). Regardless, Ice-{\it i} was
414     still observered to be a stable polymorph for all of the studied water
415     models.
416 chrisfen 2977
417 chrisfen 2979 So what is the preferred solid polymorph for simulated water? As
418     indicated above, the answer appears to be dependent both on the
419     conditions and the model used. In the case of short cutoffs without a
420     long-range interaction correction, Ice-{\it i} and Ice-$i^\prime$ have
421     the lowest free energy of the studied polymorphs with all the models.
422     Ideally, crystallization of each model under constant pressure
423     conditions, as was done with SSD/E, would aid in the identification of
424     their respective preferred structures. This work, however, helps
425     illustrate how studies involving one specific model can lead to
426     insight about important behavior of others.
427    
428     We also note that none of the water models used in this study are
429     polarizable or flexible models. It is entirely possible that the
430     polarizability of real water makes Ice-{\it i} substantially less
431     stable than ice I$_h$. However, the calculations presented above seem
432     interesting enough to communicate before the role of polarizability
433     (or flexibility) has been thoroughly investigated.
434    
435     Finally, due to the stability of Ice-{\it i} in the investigated
436     simulation conditions, the question arises as to possible experimental
437     observation of this polymorph. The rather extensive past and current
438     experimental investigation of water in the low pressure regime makes
439     us hesitant to ascribe any relevance to this work outside of the
440     simulation community. It is for this reason that we chose a name for
441     this polymorph which involves an imaginary quantity. That said, there
442     are certain experimental conditions that would provide the most ideal
443     situation for possible observation. These include the negative
444     pressure or stretched solid regime, small clusters in vacuum
445 chrisfen 2977 deposition environments, and in clathrate structures involving small
446 chrisfen 2979 non-polar molecules. For the purpose of comparison with experimental
447     results, we have calculated the oxygen-oxygen pair correlation
448     function, $g_\textrm{OO}(r)$, and the structure factor, $S(\vec{q})$
449     for the two Ice-{\it i} variants (along with example ice
450     I$_\textrm{h}$ and I$_\textrm{c}$ plots) at 77K, and they are shown in
451     figures \ref{fig:gofr} and \ref{fig:sofq} respectively. It is
452     interesting to note that the structure factors for Ice-$i^\prime$ and
453     Ice-I$_c$ are quite similar. The primary differences are small peaks
454     at 1.125, 2.29, and 2.53\AA$^{-1}$, so particular attention to these
455     regions would be needed to identify the new $i^\prime$ variant from
456     the I$_\textrm{c}$ polymorph.
457 chrisfen 2977
458 chrisfen 2979
459 chrisfen 2977 \begin{figure}
460     \includegraphics[width=\linewidth]{./figures/iceGofr.pdf}
461 chrisfen 2978 \caption{Radial distribution functions of Ice-{\it i} and ice
462     I$_\textrm{c}$ calculated from from simulations of the SSD/RF water
463     model at 77 K.}
464 chrisfen 2977 \label{fig:gofr}
465     \end{figure}
466    
467     \begin{figure}
468     \includegraphics[width=\linewidth]{./figures/sofq.pdf}
469 chrisfen 2978 \caption{Predicted structure factors for Ice-{\it i} and ice
470     I$_\textrm{c}$ at 77 K. The raw structure factors have been
471     convoluted with a gaussian instrument function (0.075 \AA$^{-1}$
472     width) to compensate for the trunction effects in our finite size
473     simulations. The labeled peaks compared favorably with ``spurious''
474     peaks observed in experimental studies of amorphous solid
475     water.\cite{Bizid87}}
476 chrisfen 2977 \label{fig:sofq}
477     \end{figure}
478