ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/chrisDissertation/Ice.tex
Revision: 3019
Committed: Fri Sep 22 13:45:24 2006 UTC (17 years, 11 months ago) by chrisfen
Content type: application/x-tex
File size: 31086 byte(s)
Log Message:
more changes and an abstract included

File Contents

# User Rev Content
1 chrisfen 3001 \chapter{\label{chap:ice}PHASE BEHAVIOR OF WATER IN COMPUTER \\ SIMULATIONS}
2 chrisfen 2987
3     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     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     property to quantify are the Gibbs and Helmholtz free energies,
15     particularly for the solid forms of water, as these predict the
16     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     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     The high-pressure phases of water (ice II-ice X as well as ice XII)
26     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     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     \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    
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     \label{fig:iceiCell}
65     \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     less dense than ice I$_\textrm{h}$.}
74     \label{fig:protOrder}
75     \end{figure}
76    
77     Results from our initial studies indicated that Ice-{\it i} is the
78     minimum energy crystal structure for the single point water models
79     investigated (for discussions on these single point dipole models, see
80     the previous work and related
81     articles\cite{Fennell04,Liu96,Bratko85}). These earlier results only
82     considered energetic stabilization and neglected entropic
83     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     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     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    
102     \section{Methods and Thermodynamic Integration}
103    
104     Canonical ensemble ({\it NVT}) molecular dynamics calculations were
105     performed using the OOPSE molecular mechanics package.\cite{Meineke05}
106     The densities chosen for the simulations were taken from
107     isobaric-isothermal ({\it NPT}) simulations performed at 1 atm and
108     200~K. Each model (and each crystal structure) was allowed to relax for
109     300~ps 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     \ref{sec:IntroIntegrate}.
114    
115    
116     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     materials.\cite{Frenkel84,Hermens88,Meijer90,Baez95a,Vlot99}. This
122     method uses a sequence of simulations over which the system of
123     interest is converted into a reference system for which the free
124     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     \begin{equation}
129     A = A_0 + \int_0^1 \left\langle \Delta V \right\rangle_\lambda d\lambda.
130     \end{equation}
131     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    
149     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, $K_\mathrm{v} =
163     4.29$~kcal~mol$^{-1}$~\AA$^{-2}$, $K_\theta\ =
164     13.88$~kcal~mol$^{-1}$~rad$^{-2}$, and $K_\omega\ =
165     17.75$~kcal~mol$^{-1}$~rad$^{-2}$. It is clear from
166     Fig. \ref{fig:waterSpring} that the values of $\theta$ range from $0$
167     to $\pi$, while $\omega$ ranges from $-\pi$ to $\pi$. The partition
168     function for a molecular crystal restrained in this fashion can be
169     evaluated analytically, and the Helmholtz Free Energy ({\it A}) is
170     given by
171     \begin{equation}
172     \begin{split}
173     A = E_m &- kT\ln\left(\frac{kT}{h\nu}\right)^3 \\
174     &- kT\ln\left[\pi^\frac{1}{2}\left(
175     \frac{8\pi^2I_\mathrm{A}kT}{h^2}\right)^\frac{1}{2}
176     \left(\frac{8\pi^2I_\mathrm{B}kT}{h^2}\right)^\frac{1}{2}
177     \left(\frac{8\pi^2I_\mathrm{C}kT}{h^2}\right)^\frac{1}{2}
178     \right] \\
179     &- kT\ln\left[\frac{kT}{2(\pi K_\omega K_\theta)^{\frac{1}{2}}}
180     \exp\left(-\frac{kT}{2K_\theta}\right)
181     \int_0^{\left(\frac{kT}{2K_\theta}\right)^\frac{1}{2}}
182     \exp(t^2)\mathrm{d}t\right],
183     \end{split}
184     \label{eq:ecFreeEnergy}
185     \end{equation}
186     where $2\pi\nu = (K_\mathrm{v}/m)^{1/2}$, and $E_m$ is the minimum
187     potential energy of the ideal crystal.\cite{Baez95a} The choice of an
188     Einstein crystal reference state is somewhat arbitrary. Any ideal
189     system for which the partition function is known exactly could be used
190     as a reference point as long as the system does not undergo a phase
191     transition during the integration path between the real and ideal
192     systems. Nada and van der Eerden have shown that the use of different
193     force constants in the Einstein crystal does not affect the total
194     free energy, and Gao {\it et al.} have shown that free energies
195     computed with the Debye crystal reference state differ from the
196     Einstein crystal by only a few tenths of a
197     kJ~mol$^{-1}$.\cite{Nada03,Gao00} These free energy differences can
198     lead to some uncertainty in the computed melting point of the solids.
199     \begin{figure}
200     \centering
201     \includegraphics[width=3.5in]{./figures/rotSpring.pdf}
202     \caption{Possible orientational motions for a restrained molecule.
203     $\theta$ angles correspond to displacement from the body-frame {\it
204     z}-axis, while $\omega$ angles correspond to rotation about the
205     body-frame {\it z}-axis. $K_\theta$ and $K_\omega$ are spring
206     constants for the harmonic springs restraining motion in the $\theta$
207     and $\omega$ directions.}
208     \label{fig:waterSpring}
209     \end{figure}
210    
211     In the case of molecular liquids, the ideal vapor is chosen as the
212     target reference state. There are several examples of liquid state
213     free energy calculations of water models present in the
214     literature.\cite{Hermens88,Quintana92,Mezei92,Baez95b} These methods
215     typically differ in regard to the path taken for switching off the
216     interaction potential to convert the system to an ideal gas of water
217     molecules. In this study, we applied one of the most convenient
218     methods and integrated over the $\lambda^4$ path, where all
219     interaction parameters are scaled equally by this transformation
220     parameter. This method has been shown to be reversible and provide
221     results in excellent agreement with other established
222     methods.\cite{Baez95b}
223    
224 chrisfen 3004 The Helmholtz free energy error was determined in the same manner in
225     both the solid and the liquid free energy calculations . At each point
226     along the integration path, we calculated the standard deviation of
227     the potential energy difference. Addition or subtraction of these
228     values to each of their respective points and integrating the curve
229     again provides the upper and lower bounds of the uncertainty in the
230     Helmholtz free energy.
231    
232 chrisfen 2987 Near the cutoff radius ($0.85 * r_{cut}$), charge, dipole, and
233     Lennard-Jones interactions were gradually reduced by a cubic switching
234     function. By applying this function, these interactions are smoothly
235     truncated, thereby avoiding the poor energy conservation which results
236     from harsher truncation schemes. The effect of a long-range
237     correction was also investigated on select model systems in a variety
238     of manners. For the SSD/RF model, a reaction field with a fixed
239     dielectric constant of 80 was applied in all
240     simulations.\cite{Onsager36} For a series of the least computationally
241     expensive models (SSD/E, SSD/RF, TIP3P, and SPC/E), simulations were
242     performed with longer cutoffs of 10.5, 12, 13.5, and 15~\AA\ to
243     compare with the 9~\AA\ cutoff results. Finally, the effects of using
244     the Ewald summation were estimated for TIP3P and SPC/E by performing
245     single configuration Particle-Mesh Ewald (PME) calculations for each
246     of the ice polymorphs.\cite{Ponder87} The calculated energy difference
247     in the presence and absence of PME was applied to the previous results
248     in order to predict changes to the free energy landscape.
249    
250 chrisfen 3019 In addition to the above procedures, we also tested how the inclusion
251     of the Lennard-Jones long-range correction affects the free energy
252     results. The correction for the Lennard-Jones trucation was included
253     by integration of the equation discussed in section
254     \ref{sec:LJCorrections}. Rather than discuss its affect alongside the
255     free energy results, we will just mention that while the correction
256     does lower the free energy of the higher density states more than the
257     lower density states, the effect is so small that it is entirely
258     overwelmed by the error in the free energy calculation. Since its
259     inclusion does not influence the results, the Lennard-Jones correction
260     was omitted from all the calculations below.
261    
262 chrisfen 2987 \section{Initial Free Energy Results}
263    
264     The calculated free energies of proton-ordered variants of three low
265     density polymorphs (I$_\textrm{h}$, I$_\textrm{c}$, and Ice-{\it i} or
266     Ice-$i^\prime$) and the stable higher density ice B are listed in
267     table \ref{tab:freeEnergy}. Ice B was included because it has been
268     shown to be a minimum free energy structure for SPC/E at ambient
269     conditions.\cite{Baez95b} In addition to the free energies, the
270     relevant transition temperatures at standard pressure are also
271     displayed in table \ref{tab:freeEnergy}. These free energy values
272     indicate that Ice-{\it i} is the most stable state for all of the
273     investigated water models. With the free energy at these state
274     points, the Gibbs-Helmholtz equation was used to project to other
275     state points and to build phase diagrams. Figures
276     \ref{fig:tp3PhaseDia} and \ref{fig:ssdrfPhaseDia} are example diagrams
277     built from the results for the TIP3P and SSD/RF water models. All
278     other models have similar structure, although the crossing points
279     between the phases move to different temperatures and pressures as
280     indicated from the transition temperatures in table
281     \ref{tab:freeEnergy}. It is interesting to note that ice
282     I$_\textrm{h}$ (and ice I$_\textrm{c}$ for that matter) do not appear
283     in any of the phase diagrams for any of the models. For purposes of
284     this study, ice B is representative of the dense ice polymorphs. A
285     recent study by Sanz {\it et al.} provides details on the phase
286     diagrams for SPC/E and TIP4P at higher pressures than those studied
287     here.\cite{Sanz04}
288     \begin{table}
289     \centering
290     \caption{HELMHOLTZ FREE ENERGIES AND TRANSITION TEMPERATURES AT 1
291     ATMOSPHERE FOR SEVERAL WATER MODELS}
292    
293     \footnotesize
294     \begin{tabular}{lccccccc}
295     \toprule
296     \toprule
297     Water Model & I$_\textrm{h}$ & I$_\textrm{c}$ & B & Ice-{\it i} & Ice-$i^\prime$ & $T_\textrm{m}$ (*$T_\textrm{s}$) & $T_\textrm{b}$\\
298     \cmidrule(lr){2-6}
299     \cmidrule(l){7-8}
300     & \multicolumn{5}{c}{(kcal mol$^{-1}$)} & \multicolumn{2}{c}{(K)}\\
301     \midrule
302     TIP5P & -11.85(3) & -11.86(2) & -11.96(2) & - & -12.29(2) & 266(7) & 337(4)\\
303     TIP4P & -11.84(3) & -12.04(2) & -12.08(3) & - & -12.33(3) & 262(6) & 354(4)\\
304     TIP3P & -11.41(2) & -11.23(3) & -11.82(3) & -12.30(3) & - & 269(7) & 357(4)\\
305     SPC/E & -12.87(2) & -13.05(2) & -13.26(3) & - & -13.55(2) & 299(6) & 396(4)\\
306     SSD/E & -11.27(2) & -11.19(4) & -12.09(2) & -12.54(2) & - & *355(4) & -\\
307     SSD/RF & -11.96(2) & -11.60(2) & -12.53(3) & -12.79(2) & - & 278(7) & 382(4)\\
308     \bottomrule
309     \end{tabular}
310     \label{tab:freeEnergy}
311     \end{table}
312    
313     \begin{figure}
314     \centering
315     \includegraphics[width=\linewidth]{./figures/tp3PhaseDia.pdf}
316     \caption{Phase diagram for the TIP3P water model in the low pressure
317     regime. The displayed $T_m$ and $T_b$ values are good predictions of
318     the experimental values; however, the solid phases shown are not the
319     experimentally observed forms. Both cubic and hexagonal ice $I$ are
320     higher in energy and don't appear in the phase diagram.}
321     \label{fig:tp3PhaseDia}
322     \end{figure}
323    
324     \begin{figure}
325     \centering
326     \includegraphics[width=\linewidth]{./figures/ssdrfPhaseDia.pdf}
327     \caption{Phase diagram for the SSD/RF water model in the low pressure
328     regime. Calculations producing these results were done under an
329     applied reaction field. It is interesting to note that this
330     computationally efficient model (over 3 times more efficient than
331     TIP3P) exhibits phase behavior similar to the less computationally
332     conservative charge based models.}
333     \label{fig:ssdrfPhaseDia}
334     \end{figure}
335    
336     We note that all of the crystals investigated in this study are ideal
337     proton-ordered antiferroelectric structures. All of the structures
338     obey the Bernal-Fowler rules and should be able to form stable
339     proton-{\it disordered} crystals which have the traditional
340     $k_\textrm{B}$ln(3/2) residual entropy at 0~K.\cite{Bernal33,Pauling35}
341     Simulations of proton-disordered structures are relatively unstable
342     with all but the most expensive water models.\cite{Nada03} Our
343     simulations have therefore been performed with the ordered
344     antiferroelectric structures which do not require the residual entropy
345     term to be accounted for in the free energies. This may result in some
346     discrepancies when comparing our melting temperatures to the melting
347     temperatures that have been calculated via thermodynamic integrations
348     of the disordered structures.\cite{Sanz04}
349    
350     Most of the water models have melting points that compare quite
351     favorably with the experimental value of 273~K. The unfortunate
352     aspect of this result is that this phase change occurs between
353     Ice-{\it i} and the liquid state rather than ice I$_h$ and the liquid
354     state. These results do not contradict other studies. Studies of ice
355     I$_h$ using TIP4P predict a $T_m$ ranging from 191 to 238~K
356     (differences being attributed to choice of interaction truncation and
357     different ordered and disordered molecular
358     arrangements).\cite{Nada03,Vlot99,Gao00,Sanz04} If the presence of ice
359     B and Ice-{\it i} were omitted, a $T_\textrm{m}$ value around 200~K
360     would be predicted from this work. However, the $T_\textrm{m}$ from
361     Ice-{\it i} is calculated to be 262~K, indicating that these
362     simulation based structures ought to be included in studies probing
363     phase transitions with this model. Also of interest in these results
364     is that SSD/E does not exhibit a melting point at 1 atm but does
365     sublime at 355~K. This is due to the significant stability of
366     Ice-{\it i} over all other polymorphs for this particular model under
367     these conditions. While troubling, this behavior resulted in the
368     spontaneous crystallization of Ice-{\it i} which led us to investigate
369     this structure. These observations provide a warning that simulations
370     of SSD/E as a ``liquid'' near 300~K are actually metastable and run
371     the risk of spontaneous crystallization. However, when a longer
372     cutoff radius is used, SSD/E prefers the liquid state under standard
373     temperature and pressure.
374    
375     \section{Effects of Potential Truncation}
376    
377     \begin{figure}
378     \includegraphics[width=\linewidth]{./figures/cutoffChange.pdf}
379     \caption{Free energy as a function of cutoff radius for SSD/E, TIP3P,
380     SPC/E, SSD/RF with a reaction field, and the TIP3P and SPC/E models
381     with an added Ewald correction term. Error for the larger cutoff
382     points is equivalent to that observed at 9.0~\AA\ (see Table
383     \ref{tab:freeEnergy}). Data for ice I$_\textrm{c}$ with TIP3P using
384     both 12 and 13.5~\AA\ cutoffs were omitted because the crystal was
385     prone to distortion and melting at 200~K. Ice-$i^\prime$ is the
386     form of Ice-{\it i} used in the SPC/E simulations.}
387     \label{fig:incCutoff}
388     \end{figure}
389    
390     For the more computationally efficient water models, we have also
391     investigated the effect of potential truncation on the computed free
392     energies as a function of the cutoff radius. As seen in
393     Fig. \ref{fig:incCutoff}, the free energies of the ice polymorphs with
394     water models lacking a long-range correction show significant cutoff
395     dependence. In general, there is a narrowing of the free energy
396     differences while moving to greater cutoff radii. As the free
397     energies for the polymorphs converge, the stability advantage that
398     Ice-{\it i} exhibits is reduced. Adjacent to each of these plots are
399     results for systems with applied or estimated long-range corrections.
400     SSD/RF was parametrized for use with a reaction field, and the benefit
401     provided by this computationally inexpensive correction is apparent.
402     The free energies are largely independent of the size of the reaction
403     field cavity in this model, so small cutoff radii mimic bulk
404     calculations quite well under SSD/RF.
405    
406     Although TIP3P was parametrized for use without the Ewald summation,
407     we have estimated the effect of this method for computing long-range
408     electrostatics for both TIP3P and SPC/E. This was accomplished by
409     calculating the potential energy of identical crystals both with and
410     without particle mesh Ewald (PME). Similar behavior to that observed
411     with reaction field is seen for both of these models. The free
412     energies show reduced dependence on cutoff radius and span a narrower
413     range for the various polymorphs. Like the dipolar water models,
414     TIP3P displays a relatively constant preference for the Ice-{\it i}
415     polymorph. Crystal preference is much more difficult to determine for
416     SPC/E. Without a long-range correction, each of the polymorphs
417     studied assumes the role of the preferred polymorph under different
418     cutoff radii. The inclusion of the Ewald correction flattens and
419     narrows the gap in free energies such that the polymorphs are
420     isoenergetic within statistical uncertainty. This suggests that other
421     conditions, such as the density in fixed-volume simulations, can
422     influence the polymorph expressed upon crystallization.
423    
424     \section{Expanded Results Using Damped Shifted Force Electrostatics}
425    
426     In chapter \ref{chap:electrostatics}, we discussed in detail a
427     pairwise method for handling electrostatics (shifted force, {\sc sf})
428     that can be used as a simple and efficient replacement for the Ewald
429     summation. Answering the question of the free energies of these ice
430     polymorphs with varying water models would be an interesting
431     application of this technique. To this end, we set up thermodynamic
432     integrations of all of the previously discussed ice polymorphs using
433     the {\sc sf} technique with a cutoff radius of 12~\AA\ and an $\alpha$
434     of 0.2125~\AA . These calculations were performed on TIP5P-E and
435     TIP4P-Ew (variants of the root models optimized for the Ewald
436     summation) as well as SPC/E, SSD/RF, and TRED (see section
437     \ref{sec:tredWater}).
438    
439     \begin{table}
440     \centering
441     \caption{HELMHOLTZ FREE ENERGIES OF ICE POLYMORPHS USING THE DAMPED
442     SHIFTED FORCE CORRECTION}
443     \begin{tabular}{ lccccc }
444     \toprule
445     \toprule
446     Model & I$_\textrm{h}$ & I$_\textrm{c}$ & B & Ice-$i$ & Ice-$i^\prime$ \\
447     \cmidrule(lr){2-6}
448     & \multicolumn{5}{c}{(kcal mol$^{-1}$)} \\
449     \midrule
450 chrisfen 3019 TIP5P-E & -11.98(4) & -11.96(4) & -11.87(3) & - & -11.95(3) \\
451 chrisfen 3004 TIP4P-Ew & -13.11(3) & -13.09(3) & -12.97(3) & - & -12.98(3) \\
452     SPC/E & -12.99(3) & -13.00(3) & -13.03(3) & - & -12.99(3) \\
453 chrisfen 3001 SSD/RF & -11.83(3) & -11.66(4) & -12.32(3) & -12.39(3) & - \\
454     TRED & -12.61(3) & -12.43(3) & -12.89(3) & -13.12(3) & - \\
455 chrisfen 2987 \end{tabular}
456     \label{tab:dampedFreeEnergy}
457     \end{table}
458 chrisfen 3001 The results of these calculations in table \ref{tab:dampedFreeEnergy}
459     show similar behavior to the Ewald results in figure
460     \ref{fig:incCutoff}, at least for SSD/RF and SPC/E which are present
461     in both. The ice polymorph Helmholtz free energies for SSD/RF order in
462     the same fashion; however Ice-$i$ and ice B are quite a bit closer in
463     free energy (nearly isoenergetic). The free energy differences between
464     ice polymorphs for TRED water parallel SSD/RF, with the exception that
465 chrisfen 3004 ice B is destabilized such that it is not very close to Ice-$i$. The
466     SPC/E results really show the near isoenergetic behavior when using
467     the electrostatics correction. Ice B has the lowest Helmholtz free
468     energy; however, all the polymorph results overlap within error.
469 chrisfen 2987
470 chrisfen 3004 The most interesting results from these calculations come from the
471     more expensive TIP4P-Ew and TIP5P-E results. Both of these models were
472     optimized for use with an electrostatic correction and are
473     geometrically arranged to mimic water following two different
474     ideas. In TIP5P-E, the primary location for the negative charge in the
475     molecule is assigned to the lone-pairs of the oxygen, while TIP4P-Ew
476     places the negative charge near the center-of-mass along the H-O-H
477     bisector. There is some debate as to which is the proper choice for
478     the negative charge location, and this has in part led to a six-site
479     water model that balances both of these options.\cite{Vega05,Nada03}
480     The limited results in table \ref{tab:dampedFreeEnergy} support the
481     results of Vega {\it et al.}, which indicate the TIP4P charge location
482     geometry is more physically valid.\cite{Vega05} With the TIP4P-Ew
483     water model, the experimentally observed polymorph (ice
484     I$_\textrm{h}$) is the preferred form with ice I$_\textrm{c}$ slightly
485     higher in energy, though overlapping within error, and the less
486     realistic ice B and Ice-$i^\prime$ are destabilized relative to these
487     polymorphs. TIP5P-E shows similar behavior to SPC/E, where there is no
488 chrisfen 3019 real free energy distinction between the various polymorphs because
489     many overlap within error. While ice B is close in free energy to the
490     other polymorphs, these results fail to support the findings of other
491     researchers indicating the preferred form of TIP5P at 1~atm is a
492     structure similar to ice B.\cite{Yamada02,Vega05,Abascal05} It should
493     be noted that we are looking at TIP5P-E rather than TIP5P, and the
494     differences in the Lennard-Jones parameters could be a reason for this
495     dissimilarity. Overall, these results indicate that TIP4P-Ew is a
496     better mimic of real water than these other models when studying
497     crystallization and solid forms of water.
498 chrisfen 3004
499 chrisfen 2987 \section{Conclusions}
500    
501     In this work, thermodynamic integration was used to determine the
502     absolute free energies of several ice polymorphs. The new polymorph,
503 chrisfen 3016 Ice-$i$ was observed to be the stable crystalline state for {\it
504 chrisfen 2987 all} the water models when using a 9.0~\AA\ cutoff. However, the free
505     energy partially depends on simulation conditions (particularly on the
506 chrisfen 3016 choice of long range correction method). Regardless, Ice-$i$ was
507 chrisfen 2987 still observed to be a stable polymorph for all of the studied water
508     models.
509    
510     So what is the preferred solid polymorph for simulated water? As
511     indicated above, the answer appears to be dependent both on the
512     conditions and the model used. In the case of short cutoffs without a
513 chrisfen 3016 long-range interaction correction, Ice-$i$ and Ice-$i^\prime$ have
514 chrisfen 2987 the lowest free energy of the studied polymorphs with all the models.
515     Ideally, crystallization of each model under constant pressure
516     conditions, as was done with SSD/E, would aid in the identification of
517     their respective preferred structures. This work, however, helps
518     illustrate how studies involving one specific model can lead to
519     insight about important behavior of others.
520    
521     We also note that none of the water models used in this study are
522 chrisfen 3016 polarizable or flexible models. It is entirely possible that the
523     polarizability of real water makes Ice-$i$ substantially less stable
524     than ice I$_\textrm{h}$. The dipole moment of the water molecules
525     increases as the system becomes more condensed, and the increasing
526     dipole moment should destabilize the tetramer structures in
527     Ice-$i$. Right now, using TIP4P-Ew with an electrostatic correction
528     gives the proper thermodynamically preferred state, and we recommend
529     this arrangement for study of crystallization processes if the
530     computational cost increase that comes with including polarizability
531     is an issue.
532 chrisfen 2987
533 chrisfen 3016 Finally, due to the stability of Ice-$i$ in the investigated
534 chrisfen 2987 simulation conditions, the question arises as to possible experimental
535     observation of this polymorph. The rather extensive past and current
536     experimental investigation of water in the low pressure regime makes
537     us hesitant to ascribe any relevance to this work outside of the
538     simulation community. It is for this reason that we chose a name for
539     this polymorph which involves an imaginary quantity. That said, there
540     are certain experimental conditions that would provide the most ideal
541     situation for possible observation. These include the negative
542     pressure or stretched solid regime, small clusters in vacuum
543     deposition environments, and in clathrate structures involving small
544     non-polar molecules. For the purpose of comparison with experimental
545     results, we have calculated the oxygen-oxygen pair correlation
546     function, $g_\textrm{OO}(r)$, and the structure factor, $S(\vec{q})$
547     for the two Ice-{\it i} variants (along with example ice
548     I$_\textrm{h}$ and I$_\textrm{c}$ plots) at 77~K, and they are shown in
549     figures \ref{fig:gofr} and \ref{fig:sofq} respectively. It is
550     interesting to note that the structure factors for Ice-$i^\prime$ and
551     Ice-I$_c$ are quite similar. The primary differences are small peaks
552     at 1.125, 2.29, and 2.53~\AA$^{-1}$, so particular attention to these
553     regions would be needed to identify the new $i^\prime$ variant from
554     the I$_\textrm{c}$ polymorph.
555    
556    
557     \begin{figure}
558     \includegraphics[width=\linewidth]{./figures/iceGofr.pdf}
559     \caption{Radial distribution functions of Ice-{\it i} and ice
560     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     \caption{Predicted structure factors for Ice-{\it i} and ice
568     I$_\textrm{c}$ at 77~K. The raw structure factors have been
569     convoluted with a gaussian instrument function (0.075~\AA$^{-1}$
570     width) to compensate for the truncation effects in our finite size
571     simulations. The labeled peaks compared favorably with ``spurious''
572     peaks observed in experimental studies of amorphous solid
573     water.\cite{Bizid87}}
574     \label{fig:sofq}
575     \end{figure}
576