ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/iceiPaper/iceiPaper.tex
(Generate patch)

Comparing trunk/iceiPaper/iceiPaper.tex (file contents):
Revision 1488 by chrisfen, Tue Sep 21 20:17:12 2004 UTC vs.
Revision 1908 by chrisfen, Fri Jan 7 18:42:41 2005 UTC

# Line 1 | Line 1
1   %\documentclass[prb,aps,twocolumn,tabularx]{revtex4}
2   \documentclass[11pt]{article}
3 %\documentclass[11pt]{article}
3   \usepackage{endfloat}
4   \usepackage{amsmath}
5   \usepackage{epsf}
# Line 20 | Line 19
19  
20   \begin{document}
21  
22 < \title{Ice-{\it i}: a simulation-predicted ice polymorph which is more
23 < stable than Ice $I_h$ for point-charge and point-dipole water models}
22 > \title{Computational free energy studies of a new ice polymorph which
23 > exhibits greater stability than Ice $I_h$}
24  
25   \author{Christopher J. Fennell and J. Daniel Gezelter \\
26 < Department of Chemistry and Biochemistry\\ University of Notre Dame\\
26 > Department of Chemistry and Biochemistry\\
27 > University of Notre Dame\\
28   Notre Dame, Indiana 46556}
29  
30   \date{\today}
# Line 33 | Line 33 | The absolute free energies of several ice polymorphs w
33   %\doublespacing
34  
35   \begin{abstract}
36 < The absolute free energies of several ice polymorphs which are stable
37 < at low pressures were calculated using thermodynamic integration to a
38 < reference system (the Einstein crystal).  These integrations were
39 < performed for most of the common water models (SPC/E, TIP3P, TIP4P,
40 < TIP5P, SSD/E, SSD/RF). Ice-{\it i}, a structure we recently observed
41 < crystallizing at room temperature for one of the single-point water
42 < models, was determined to be the stable crystalline state (at 1 atm)
43 < for {\it all} the water models investigated.  Phase diagrams were
44 < generated, and phase coexistence lines were determined for all of the
45 < known low-pressure ice structures under all of these water models.
46 < Additionally, potential truncation was shown to have an effect on the
47 < calculated free energies, and can result in altered free energy
48 < landscapes.  Structure factor predictions for the new crystal were
49 < generated and we await experimental confirmation of the existence of
50 < this new polymorph.
36 > The absolute free energies of several ice polymorphs were calculated
37 > using thermodynamic integration.  These polymorphs are predicted by
38 > computer simulations using a variety of common water models to be
39 > stable at low pressures.  A recently discovered ice polymorph that has
40 > as yet {\it only} been observed in computer simulations (Ice-{\it i}),
41 > was determined to be the stable crystalline state for {\it all} the
42 > water models investigated.  Phase diagrams were generated, and phase
43 > coexistence lines were determined for all of the known low-pressure
44 > ice structures.  Additionally, potential truncation was shown to play
45 > a role in the resulting shape of the free energy landscape.
46   \end{abstract}
47  
48   %\narrowtext
# Line 67 | Line 62 | the limitations of each of the
62   hydrophobic effect.\cite{Yamada02,Marrink94,Gallagher03} With the
63   choice of models available, it is only natural to compare the models
64   under interesting thermodynamic conditions in an attempt to clarify
65 < the limitations of each of the
66 < models.\cite{Jorgensen83,Jorgensen98b,Clancy94,Mahoney01} Two
67 < important properties to quantify are the Gibbs and Helmholtz free
68 < energies, particularly for the solid forms of water.  Difficulty in
69 < these types of studies typically arises from the assortment of
70 < possible crystalline polymorphs that water adopts over a wide range of
71 < pressures and temperatures.  There are currently 13 recognized forms
72 < of ice, and it is a challenging task to investigate the entire free
73 < energy landscape.\cite{Sanz04} Ideally, research is focused on the
65 > the limitations of
66 > each.\cite{Jorgensen83,Jorgensen98b,Clancy94,Mahoney01} Two important
67 > properties to quantify are the Gibbs and Helmholtz free energies,
68 > particularly for the solid forms of water as these predict the
69 > thermodynamic stability of the various phases.  Water has a
70 > particularly rich phase diagram and takes on a number of different and
71 > stable crystalline structures as the temperature and pressure are
72 > varied.  It is a challenging task to investigate the entire free
73 > energy landscape\cite{Sanz04}; and ideally, research is focused on the
74   phases having the lowest free energy at a given state point, because
75   these phases will dictate the relevant transition temperatures and
76 < pressures for the model.
76 > pressures for the model.  
77  
78 < In this paper, standard reference state methods were applied to known
79 < crystalline water polymorphs in the low pressure regime.  This work is
80 < unique in that one of the crystal lattices was arrived at through
81 < crystallization of a computationally efficient water model under
82 < constant pressure and temperature conditions. Crystallization events
83 < are interesting in and of themselves;\cite{Matsumoto02,Yamada02}
84 < however, the crystal structure obtained in this case is different from
85 < any previously observed ice polymorphs in experiment or
86 < simulation.\cite{Fennell04} We have named this structure Ice-{\it i}
87 < to indicate its origin in computational simulation. The unit cell
88 < (Fig. \ref{iceiCell}A) consists of eight water molecules that stack in
89 < rows of interlocking water tetramers. Proton ordering can be
90 < accomplished by orienting two of the molecules so that both of their
91 < donated hydrogen bonds are internal to their tetramer
92 < (Fig. \ref{protOrder}). As expected in an ice crystal constructed of
93 < water tetramers, the hydrogen bonds are not as linear as those
94 < observed in ice $I_h$, however the interlocking of these subunits
95 < appears to provide significant stabilization to the overall
96 < crystal. The arrangement of these tetramers results in surrounding
97 < open octagonal cavities that are typically greater than 6.3 \AA\ in
98 < diameter. This relatively open overall structure leads to crystals
99 < that are 0.07 g/cm$^3$ less dense on average than ice $I_h$.
78 > The high-pressure phases of water (ice II - ice X as well as ice XII)
79 > have been studied extensively both experimentally and
80 > computationally. In this paper, standard reference state methods were
81 > applied in the {\it low} pressure regime to evaluate the free energies
82 > for a few known crystalline water polymorphs that might be stable at
83 > these pressures.  This work is unique in that one of the crystal
84 > lattices was arrived at through crystallization of a computationally
85 > efficient water model under constant pressure and temperature
86 > conditions.  Crystallization events are interesting in and of
87 > themselves\cite{Matsumoto02,Yamada02}; however, the crystal structure
88 > obtained in this case is different from any previously observed ice
89 > polymorphs in experiment or simulation.\cite{Fennell04} We have named
90 > this structure Ice-{\it i} to indicate its origin in computational
91 > simulation. The unit cell of Ice-{\it i} and an axially-elongated
92 > variant named Ice-{\it i}$^\prime$ both consist of eight water
93 > molecules that stack in rows of interlocking water tetramers as
94 > illustrated in figures \ref{unitcell}A and \ref{unitcell}B.  These
95 > tetramers form a crystal structure similar in appearance to a recent
96 > two-dimensional surface tessellation simulated on silica.\cite{Yang04}
97 > As expected in an ice crystal constructed of water tetramers, the
98 > hydrogen bonds are not as linear as those observed in ice $I_h$,
99 > however the interlocking of these subunits appears to provide
100 > significant stabilization to the overall crystal.  The arrangement of
101 > these tetramers results in octagonal cavities that are typically
102 > greater than 6.3 \AA\ in diameter (Fig. \ref{iCrystal}).  This open
103 > structure leads to crystals that are typically 0.07 g/cm$^3$ less
104 > dense than ice $I_h$.
105  
106   \begin{figure}
107 + \centering
108   \includegraphics[width=\linewidth]{unitCell.eps}
109 < \caption{Unit cells for (A) Ice-{\it i} and (B) Ice-{\it i}$^\prime$,
110 < the elongated variant of Ice-{\it i}.  The spheres represent the
111 < center-of-mass locations of the water molecules.  The $a$ to $c$
112 < ratios for Ice-{\it i} and Ice-{\it i}$^\prime$ are given by
113 < $a:2.1214c$ and $a:1.7850c$ respectively.}
113 < \label{iceiCell}
109 > \caption{(A) Unit cells for Ice-{\it i} and (B) Ice-{\it i}$^\prime$.  
110 > The spheres represent the center-of-mass locations of the water
111 > molecules.  The $a$ to $c$ ratios for Ice-{\it i} and Ice-{\it
112 > i}$^\prime$ are given by 2.1214 and 1.785 respectively.}
113 > \label{unitcell}
114   \end{figure}
115  
116   \begin{figure}
117 + \centering
118   \includegraphics[width=\linewidth]{orderedIcei.eps}
119 < \caption{Image of a proton ordered crystal of Ice-{\it i} looking
120 < down the (001) crystal face. The rows of water tetramers surrounded by
121 < octagonal pores leads to a crystal structure that is significantly
122 < less dense than ice $I_h$.}
122 < \label{protOrder}
119 > \caption{A rendering of a proton ordered crystal of Ice-{\it i} looking
120 > down the (001) crystal face.  The presence of large octagonal pores
121 > leads to a polymorph that is less dense than ice $I_h$.}
122 > \label{iCrystal}
123   \end{figure}
124  
125   Results from our previous study indicated that Ice-{\it i} is the
126 < minimum energy crystal structure for the single point water models we
127 < had investigated (for discussions on these single point dipole models,
128 < see our previous work and related
129 < articles).\cite{Fennell04,Liu96,Bratko85} Those results only
130 < considered energetic stabilization and neglected entropic
131 < contributions to the overall free energy. To address this issue, we
126 > minimum energy crystal structure for the single point water models
127 > investigated (for discussions on these single point dipole models, see
128 > our previous work and related
129 > articles).\cite{Fennell04,Liu96,Bratko85} Our earlier results
130 > considered only energetic stabilization and neglected entropic
131 > contributions to the overall free energy.  To address this issue, we
132   have calculated the absolute free energy of this crystal using
133 < thermodynamic integration and compared to the free energies of cubic
134 < and hexagonal ice $I$ (the experimental low density ice polymorphs)
135 < and ice B (a higher density, but very stable crystal structure
136 < observed by B\`{a}ez and Clancy in free energy studies of
137 < SPC/E).\cite{Baez95b} This work includes results for the water model
138 < from which Ice-{\it i} was crystallized (SSD/E) in addition to several
139 < common water models (TIP3P, TIP4P, TIP5P, and SPC/E) and a reaction
140 < field parametrized single point dipole water model (SSD/RF). It should
141 < be noted that a second version of Ice-{\it i} (Ice-{\it i}$^\prime$)
142 < was used in calculations involving SPC/E, TIP4P, and TIP5P. The unit
143 < cell of this crystal (Fig. \ref{iceiCell}B) is similar to the Ice-{\it
144 < i} unit it is extended in the direction of the (001) face and
145 < compressed along the other two faces.  There is typically a small
146 < distortion of proton ordered Ice-{\it i}$^\prime$ that converts the
147 < normally square tetramer into a rhombus with alternating approximately
148 < 85 and 95 degree angles.  The degree of this distortion is model
149 < dependent and significant enough to split the tetramer diagonal
150 < location peak in the radial distribution function.
133 > thermodynamic integration and compared it to the free energies of ice
134 > $I_c$ and ice $I_h$ (the common low density ice polymorphs) and ice B
135 > (a higher density, but very stable crystal structure observed by
136 > B\`{a}ez and Clancy in free energy studies of SPC/E).\cite{Baez95b}
137 > This work includes results for the water model from which Ice-{\it i}
138 > was crystallized (SSD/E) in addition to several common water models
139 > (TIP3P, TIP4P, TIP5P, and SPC/E) and a reaction field parametrized
140 > single point dipole water model (SSD/RF).  The axially-elongated
141 > variant, Ice-{\it i}$^\prime$, was used in calculations involving
142 > SPC/E, TIP4P, and TIP5P.  The square tetramers in Ice-{\it i} distort
143 > in Ice-{\it i}$^\prime$ to form a rhombus with alternating 85 and 95
144 > degree angles.  Under SPC/E, TIP4P, and TIP5P, this geometry is better
145 > at forming favorable hydrogen bonds.  The degree of rhomboid
146 > distortion depends on the water model used, but is significant enough
147 > to split a peak in the radial distribution function which corresponds
148 > to diagonal sites in the tetramers.
149  
150   \section{Methods}
151  
152   Canonical ensemble (NVT) molecular dynamics calculations were
153 < performed using the OOPSE molecular mechanics package.\cite{Meineke05}
153 > performed using the OOPSE molecular mechanics program.\cite{Meineke05}
154   All molecules were treated as rigid bodies, with orientational motion
155 < propagated using the symplectic DLM integration method. Details about
155 > propagated using the symplectic DLM integration method.  Details about
156   the implementation of this technique can be found in a recent
157   publication.\cite{Dullweber1997}
158  
159 < Thermodynamic integration is an established technique for
160 < determination of free energies of condensed phases of
161 < materials.\cite{Frenkel84,Hermens88,Meijer90,Baez95a,Vlot99}. This
162 < method, implemented in the same manner illustrated by B\`{a}ez and
163 < Clancy, was utilized to calculate the free energy of several ice
164 < crystals at 200 K using the TIP3P, TIP4P, TIP5P, SPC/E, SSD/RF, and
165 < SSD/E water models.\cite{Baez95a} Liquid state free energies at 300
166 < and 400 K for all of these water models were also determined using
167 < this same technique in order to determine melting points and to
168 < generate phase diagrams. All simulations were carried out at densities
169 < which correspond to a pressure of approximately 1 atm at their
170 < respective temperatures.
171 <
174 < Thermodynamic integration involves a sequence of simulations during
175 < which the system of interest is converted into a reference system for
176 < which the free energy is known analytically. This transformation path
177 < is then integrated in order to determine the free energy difference
178 < between the two states:
159 > Thermodynamic integration was utilized to calculate the Helmholtz free
160 > energies ($A$) of the listed water models at various state points
161 > using the OOPSE molecular dynamics program.\cite{Meineke05}
162 > Thermodynamic integration is an established technique that has been
163 > used extensively in the calculation of free energies for condensed
164 > phases of
165 > materials.\cite{Frenkel84,Hermens88,Meijer90,Baez95a,Vlot99}.  This
166 > method uses a sequence of simulations during which the system of
167 > interest is converted into a reference system for which the free
168 > energy is known analytically ($A_0$).  The difference in potential
169 > energy between the reference system and the system of interest
170 > ($\Delta V$) is then integrated in order to determine the free energy
171 > difference between the two states:
172   \begin{equation}
173 < \Delta A = \int_0^1\left\langle\frac{\partial V(\lambda
181 < )}{\partial\lambda}\right\rangle_\lambda d\lambda,
173 > A = A_0 + \int_0^1 \left\langle \Delta V \right\rangle_\lambda d\lambda.
174   \end{equation}
175 < where $V$ is the interaction potential and $\lambda$ is the
176 < transformation parameter that scales the overall
177 < potential. Simulations are distributed strategically along this path
178 < in order to sufficiently sample the regions of greatest change in the
179 < potential. Typical integrations in this study consisted of $\sim$25
188 < simulations ranging from 300 ps (for the unaltered system) to 75 ps
189 < (near the reference state) in length.
175 > Here, $\lambda$ is the parameter that governs the transformation
176 > between the reference system and the system of interest.  For
177 > crystalline phases, an harmonically-restrained (Einsten) crystal is
178 > chosen as the reference state, while for liquid phases, the ideal gas
179 > is taken as the reference state.  
180  
181 < For the thermodynamic integration of molecular crystals, the Einstein
182 < crystal was chosen as the reference system. In an Einstein crystal,
183 < the molecules are restrained at their ideal lattice locations and
184 < orientations. Using harmonic restraints, as applied by B\`{a}ez and
195 < Clancy, the total potential for this reference crystal
196 < ($V_\mathrm{EC}$) is the sum of all the harmonic restraints,
181 > In an Einstein crystal, the molecules are restrained at their ideal
182 > lattice locations and orientations. Using harmonic restraints, as
183 > applied by B\`{a}ez and Clancy, the total potential for this reference
184 > crystal ($V_\mathrm{EC}$) is the sum of all the harmonic restraints,
185   \begin{equation}
186   V_\mathrm{EC} = \frac{K_\mathrm{v}r^2}{2} + \frac{K_\theta\theta^2}{2} +
187   \frac{K_\omega\omega^2}{2},
# Line 201 | Line 189 | respectively.  It is clear from Fig. \ref{waterSpring}
189   where $K_\mathrm{v}$, $K_\mathrm{\theta}$, and $K_\mathrm{\omega}$ are
190   the spring constants restraining translational motion and deflection
191   of and rotation around the principle axis of the molecule
192 < respectively.  It is clear from Fig. \ref{waterSpring} that the values
193 < of $\theta$ range from $0$ to $\pi$, while $\omega$ ranges from
194 < $-\pi$ to $\pi$.  The partition function for a molecular crystal
192 > respectively.  These spring constants are typically calculated from
193 > the mean-square displacements of water molecules in an unrestrained
194 > ice crystal at 200 K.  For these studies, $K_\mathrm{r} = 4.29$ kcal
195 > mol$^{-1}$, $K_\theta\ = 13.88$ kcal mol$^{-1}$, and $K_\omega\ =
196 > 17.75$ kcal mol$^{-1}$.  It is clear from Fig. \ref{waterSpring} that
197 > the values of $\theta$ range from $0$ to $\pi$, while $\omega$ ranges
198 > from $-\pi$ to $\pi$.  The partition function for a molecular crystal
199   restrained in this fashion can be evaluated analytically, and the
200   Helmholtz Free Energy ({\it A}) is given by
201   \begin{eqnarray}
# Line 221 | Line 213 | potential energy of the ideal crystal.\cite{Baez95a}
213   potential energy of the ideal crystal.\cite{Baez95a}
214  
215   \begin{figure}
216 < \includegraphics[width=\linewidth]{rotSpring.eps}
216 > \centering
217 > \includegraphics[width=4in]{rotSpring.eps}
218   \caption{Possible orientational motions for a restrained molecule.
219   $\theta$ angles correspond to displacement from the body-frame {\it
220   z}-axis, while $\omega$ angles correspond to rotation about the
221 < body-frame {\it z}-axis. $K_\theta$ and $K_\omega$ are spring
221 > body-frame {\it z}-axis.  $K_\theta$ and $K_\omega$ are spring
222   constants for the harmonic springs restraining motion in the $\theta$
223   and $\omega$ directions.}
224   \label{waterSpring}
# Line 237 | Line 230 | molecules.  In this study, we applied of one of the mo
230   literature.\cite{Hermens88,Quintana92,Mezei92,Baez95b} These methods
231   typically differ in regard to the path taken for switching off the
232   interaction potential to convert the system to an ideal gas of water
233 < molecules.  In this study, we applied of one of the most convenient
233 > molecules.  In this study, we applied one of the most convenient
234   methods and integrated over the $\lambda^4$ path, where all
235   interaction parameters are scaled equally by this transformation
236   parameter.  This method has been shown to be reversible and provide
237   results in excellent agreement with other established
238   methods.\cite{Baez95b}
239  
240 < Charge, dipole, and Lennard-Jones interactions were modified by a
241 < cubic switching between 100\% and 85\% of the cutoff value (9 \AA
242 < ). By applying this function, these interactions are smoothly
240 > Near the cutoff radius ($0.85 * r_{cut}$), charge, dipole, and
241 > Lennard-Jones interactions were gradually reduced by a cubic switching
242 > function.  By applying this function, these interactions are smoothly
243   truncated, thereby avoiding the poor energy conservation which results
244 < from harsher truncation schemes. The effect of a long-range correction
245 < was also investigated on select model systems in a variety of
246 < manners. For the SSD/RF model, a reaction field with a fixed
244 > from harsher truncation schemes.  The effect of a long-range
245 > correction was also investigated on select model systems in a variety
246 > of manners.  For the SSD/RF model, a reaction field with a fixed
247   dielectric constant of 80 was applied in all
248   simulations.\cite{Onsager36} For a series of the least computationally
249 < expensive models (SSD/E, SSD/RF, and TIP3P), simulations were
250 < performed with longer cutoffs of 12 and 15 \AA\ to compare with the 9
251 < \AA\ cutoff results. Finally, the effects of utilizing an Ewald
252 < summation were estimated for TIP3P and SPC/E by performing single
253 < configuration calculations with Particle-Mesh Ewald (PME) in the
254 < TINKER molecular mechanics software package.\cite{Tinker} The
249 > expensive models (SSD/E, SSD/RF, TIP3P, and SPC/E), simulations were
250 > performed with longer cutoffs of 10.5, 12, 13.5, and 15 \AA\ to
251 > compare with the 9 \AA\ cutoff results.  Finally, the effects of using
252 > the Ewald summation were estimated for TIP3P and SPC/E by performing
253 > single configuration Particle-Mesh Ewald (PME)
254 > calculations~\cite{Tinker} for each of the ice polymorphs.  The
255   calculated energy difference in the presence and absence of PME was
256   applied to the previous results in order to predict changes to the
257   free energy landscape.
258  
259 < \section{Results and discussion}
259 > \section{Results and Discussion}
260  
261 < The free energy of proton-ordered Ice-{\it i} was calculated and
262 < compared with the free energies of proton ordered variants of the
263 < experimentally observed low-density ice polymorphs, $I_h$ and $I_c$,
264 < as well as the higher density ice B, observed by B\`{a}ez and Clancy
265 < and thought to be the minimum free energy structure for the SPC/E
266 < model at ambient conditions (Table \ref{freeEnergy}).\cite{Baez95b}
267 < Ice XI, the experimentally-observed proton-ordered variant of ice
268 < $I_h$, was investigated initially, but was found to be not as stable
269 < as proton disordered or antiferroelectric variants of ice $I_h$. The
270 < proton ordered variant of ice $I_h$ used here is a simple
271 < antiferroelectric version that we devised, and it has an 8 molecule
272 < unit cell similar to other predicted antiferroelectric $I_h$
273 < crystals.\cite{Davidson84} The crystals contained 648 or 1728
274 < molecules for ice B, 1024 or 1280 molecules for ice $I_h$, 1000
275 < molecules for ice $I_c$, or 1024 molecules for Ice-{\it i}. The larger
276 < crystal sizes were necessary for simulations involving larger cutoff
277 < values.
261 > The calculated free energies of proton-ordered variants of three low
262 > density polymorphs ($I_h$, $I_c$, and Ice-{\it i} or Ice-{\it
263 > i}$^\prime$) and the stable higher density ice B are listed in Table
264 > \ref{freeEnergy}.  Ice B was included because it has been
265 > shown to be a minimum free energy structure for SPC/E at ambient
266 > conditions.\cite{Baez95b} In addition to the free energies, the
267 > relevant transition temperatures at standard pressure are also
268 > displayed in Table \ref{freeEnergy}.  These free energy values
269 > indicate that Ice-{\it i} is the most stable state for all of the
270 > investigated water models.  With the free energy at these state
271 > points, the Gibbs-Helmholtz equation was used to project to other
272 > state points and to build phase diagrams.  Figure \ref{tp3PhaseDia} is
273 > an example diagram built from the results for the TIP3P water model.
274 > All other models have similar structure, although the crossing points
275 > between the phases move to different temperatures and pressures as
276 > indicated from the transition temperatures in Table \ref{freeEnergy}.
277 > It is interesting to note that ice $I_h$ (and ice $I_c$ for that
278 > matter) do not appear in any of the phase diagrams for any of the
279 > models.  For purposes of this study, ice B is representative of the
280 > dense ice polymorphs.  A recent study by Sanz {\it et al.} provides
281 > details on the phase diagrams for SPC/E and TIP4P at higher pressures
282 > than those studied here.\cite{Sanz04}
283  
284   \begin{table*}
285   \begin{minipage}{\linewidth}
288 \renewcommand{\thefootnote}{\thempfootnote}
286   \begin{center}
287 < \caption{Calculated free energies for several ice polymorphs with a
288 < variety of common water models. All calculations used a cutoff radius
289 < of 9 \AA\ and were performed at 200 K and $\sim$1 atm. Units are
290 < kcal/mol. Calculated error of the final digits is in parentheses. *Ice
291 < $I_c$ rapidly converts to a liquid at 200 K with the SSD/RF model.}
292 < \begin{tabular}{ l  c  c  c  c }
293 < \hline
297 < Water Model & $I_h$ & $I_c$ & B & Ice-{\it i}\\
287 > \caption{Calculated free energies for several ice polymorphs along
288 > with the calculated melting (or sublimation) and boiling points for
289 > the investigated water models.  All free energy calculations used a
290 > cutoff radius of 9.0 \AA\ and were performed at 200 K and $\sim$1 atm.
291 > Units of free energy are kcal/mol, while transition temperature are in
292 > Kelvin.  Calculated error of the final digits is in parentheses.}
293 > \begin{tabular}{lccccccc}
294   \hline
295 < TIP3P & -11.41(2) & -11.23(3) & -11.82(3) & -12.30(3)\\
296 < TIP4P & -11.84(3) & -12.04(2) & -12.08(3) & -12.33(3)\\
297 < TIP5P & -11.85(3) & -11.86(2) & -11.96(2) & -12.29(2)\\
298 < SPC/E & -12.67(2) & -12.96(2) & -13.25(3) & -13.55(2)\\
299 < SSD/E & -11.27(2) & -11.19(4) & -12.09(2) & -12.54(2)\\
300 < SSD/RF & -11.51(2) & NA* & -12.08(3) & -12.29(2)\\
295 > Water Model & $I_h$ & $I_c$ & B & Ice-{\it i} & Ice-{\it i}$^\prime$ & $T_m$ (*$T_s$) & $T_b$\\
296 > \hline
297 > TIP3P & -11.41(2) & -11.23(3) & -11.82(3) & -12.30(3) & - & 269(4) & 357(2)\\
298 > TIP4P & -11.84(3) & -12.04(2) & -12.08(3) & - & -12.33(3) & 266(5) & 354(2)\\
299 > TIP5P & -11.85(3) & -11.86(2) & -11.96(2) & - & -12.29(2) & 271(4) & 337(2)\\
300 > SPC/E & -12.87(2) & -13.05(2) & -13.26(3) & - & -13.55(2) & 296(3) & 396(2)\\
301 > SSD/E & -11.27(2) & -11.19(4) & -12.09(2) & -12.54(2) & - & *355(2) & -\\
302 > SSD/RF & -11.51(2) & -11.47(2) & -12.08(3) & -12.29(2) & - & 278(4) & 349(2)\\
303   \end{tabular}
304   \label{freeEnergy}
305   \end{center}
306   \end{minipage}
307   \end{table*}
308  
311 The free energy values computed for the studied polymorphs indicate
312 that Ice-{\it i} is the most stable state for all of the common water
313 models studied. With the calculated free energy at these state points,
314 the Gibbs-Helmholtz equation was used to project to other state points
315 and to build phase diagrams.  Figures
316 \ref{tp3phasedia} and \ref{ssdrfphasedia} are example diagrams built
317 from the free energy results. All other models have similar structure,
318 although the crossing points between the phases move to slightly
319 different temperatures and pressures. It is interesting to note that
320 ice $I$ does not exist in either cubic or hexagonal form in any of the
321 phase diagrams for any of the models. For purposes of this study, ice
322 B is representative of the dense ice polymorphs. A recent study by
323 Sanz {\it et al.} goes into detail on the phase diagrams for SPC/E and
324 TIP4P at higher pressures than those studied here.\cite{Sanz04}
325
309   \begin{figure}
310   \includegraphics[width=\linewidth]{tp3PhaseDia.eps}
311   \caption{Phase diagram for the TIP3P water model in the low pressure
312 < regime. The displayed $T_m$ and $T_b$ values are good predictions of
312 > regime.  The displayed $T_m$ and $T_b$ values are good predictions of
313   the experimental values; however, the solid phases shown are not the
314 < experimentally observed forms. Both cubic and hexagonal ice $I$ are
314 > experimentally observed forms.  Both cubic and hexagonal ice $I$ are
315   higher in energy and don't appear in the phase diagram.}
316 < \label{tp3phasedia}
316 > \label{tp3PhaseDia}
317   \end{figure}
318  
319 < \begin{figure}
320 < \includegraphics[width=\linewidth]{ssdrfPhaseDia.eps}
321 < \caption{Phase diagram for the SSD/RF water model in the low pressure
322 < regime. Calculations producing these results were done under an
323 < applied reaction field. It is interesting to note that this
324 < computationally efficient model (over 3 times more efficient than
325 < TIP3P) exhibits phase behavior similar to the less computationally
326 < conservative charge based models.}
344 < \label{ssdrfphasedia}
345 < \end{figure}
346 <
347 < \begin{table*}
348 < \begin{minipage}{\linewidth}
349 < \renewcommand{\thefootnote}{\thempfootnote}
350 < \begin{center}
351 < \caption{Melting ($T_m$), boiling ($T_b$), and sublimation ($T_s$)
352 < temperatures at 1 atm for several common water models compared with
353 < experiment. The $T_m$ and $T_s$ values from simulation correspond to a
354 < transition between Ice-{\it i} (or Ice-{\it i}$^\prime$) and the
355 < liquid or gas state.}
356 < \begin{tabular}{ l  c  c  c  c  c  c  c }
357 < \hline
358 < Equilibria Point & TIP3P & TIP4P & TIP5P & SPC/E & SSD/E & SSD/RF & Exp.\\
359 < \hline
360 < $T_m$ (K)  & 269(4) & 266(5) & 271(4) & 296(3) & - & 278(4) & 273\\
361 < $T_b$ (K)  & 357(2) & 354(2) & 337(2) & 396(2) & - & 348(2) & 373\\
362 < $T_s$ (K)  & - & - & - & - & 355(2) & - & -\\
363 < \end{tabular}
364 < \label{meltandboil}
365 < \end{center}
366 < \end{minipage}
367 < \end{table*}
368 <
369 < Table \ref{meltandboil} lists the melting and boiling temperatures
370 < calculated from this work. Surprisingly, most of these models have
371 < melting points that compare quite favorably with experiment. The
372 < unfortunate aspect of this result is that this phase change occurs
373 < between Ice-{\it i} and the liquid state rather than ice $I_h$ and the
374 < liquid state. These results are actually not contrary to previous
375 < studies in the literature. Earlier free energy studies of ice $I$
376 < using TIP4P predict a $T_m$ ranging from 214 to 238 K (differences
377 < being attributed to choice of interaction truncation and different
378 < ordered and disordered molecular
319 > Most of the water models have melting points that compare quite
320 > favorably with the experimental value of 273 K.  The unfortunate
321 > aspect of this result is that this phase change occurs between
322 > Ice-{\it i} and the liquid state rather than ice $I_h$ and the liquid
323 > state.  These results do not contradict other studies.  Studies of ice
324 > $I_h$ using TIP4P predict a $T_m$ ranging from 214 to 238 K
325 > (differences being attributed to choice of interaction truncation and
326 > different ordered and disordered molecular
327   arrangements).\cite{Vlot99,Gao00,Sanz04} If the presence of ice B and
328   Ice-{\it i} were omitted, a $T_m$ value around 210 K would be
329 < predicted from this work. However, the $T_m$ from Ice-{\it i} is
330 < calculated at 265 K, significantly higher in temperature than the
331 < previous studies. Also of interest in these results is that SSD/E does
332 < not exhibit a melting point at 1 atm, but it shows a sublimation point
333 < at 355 K. This is due to the significant stability of Ice-{\it i} over
334 < all other polymorphs for this particular model under these
335 < conditions. While troubling, this behavior resulted in spontaneous
336 < crystallization of Ice-{\it i} and led us to investigate this
337 < structure. These observations provide a warning that simulations of
338 < SSD/E as a ``liquid'' near 300 K are actually metastable and run the
339 < risk of spontaneous crystallization. However, this risk lessens when
340 < applying a longer cutoff.
329 > predicted from this work.  However, the $T_m$ from Ice-{\it i} is
330 > calculated to be 265 K, indicating that these simulation based
331 > structures ought to be included in studies probing phase transitions
332 > with this model.  Also of interest in these results is that SSD/E does
333 > not exhibit a melting point at 1 atm but does sublime at 355 K.  This
334 > is due to the significant stability of Ice-{\it i} over all other
335 > polymorphs for this particular model under these conditions.  While
336 > troubling, this behavior resulted in the spontaneous crystallization
337 > of Ice-{\it i} which led us to investigate this structure.  These
338 > observations provide a warning that simulations of SSD/E as a
339 > ``liquid'' near 300 K are actually metastable and run the risk of
340 > spontaneous crystallization.  However, when a longer cutoff radius is
341 > used, SSD/E prefers the liquid state under standard temperature and
342 > pressure.
343  
344   \begin{figure}
345   \includegraphics[width=\linewidth]{cutoffChange.eps}
346 < \caption{Free energy as a function of cutoff radius for (A) SSD/E, (B)
347 < TIP3P, and (C) SSD/RF. Data points omitted include SSD/E: $I_c$ 12
348 < \AA\, TIP3P: $I_c$ 12 \AA\ and B 12 \AA\, and SSD/RF: $I_c$ 9
349 < \AA . These crystals are unstable at 200 K and rapidly convert into
350 < liquids. The connecting lines are qualitative visual aid.}
346 > \caption{Free energy as a function of cutoff radius for SSD/E, TIP3P,
347 > SPC/E, SSD/RF with a reaction field, and the TIP3P and SPC/E models
348 > with an added Ewald correction term.  Error for the larger cutoff
349 > points is equivalent to that observed at 9.0\AA\ (see Table
350 > \ref{freeEnergy}).  Data for ice I$_c$ with TIP3P using both 12 and
351 > 13.5 \AA\ cutoffs were omitted because the crystal was prone to
352 > distortion and melting at 200 K.  Ice-{\it i}$^\prime$ is the form of
353 > Ice-{\it i} used in the SPC/E simulations.}
354   \label{incCutoff}
355   \end{figure}
356  
357 < Increasing the cutoff radius in simulations of the more
358 < computationally efficient water models was done in order to evaluate
359 < the trend in free energy values when moving to systems that do not
360 < involve potential truncation. As seen in Fig. \ref{incCutoff}, the
361 < free energy of all the ice polymorphs show a substantial dependence on
362 < cutoff radius. In general, there is a narrowing of the free energy
363 < differences while moving to greater cutoff radius. Interestingly, by
364 < increasing the cutoff radius, the free energy gap was narrowed enough
365 < in the SSD/E model that the liquid state is preferred under standard
366 < simulation conditions (298 K and 1 atm). Thus, it is recommended that
367 < simulations using this model choose interaction truncation radii
368 < greater than 9 \AA\ . This narrowing trend is much more subtle in the
369 < case of SSD/RF, indicating that the free energies calculated with a
370 < reaction field present provide a more accurate picture of the free
371 < energy landscape in the absence of potential truncation.
357 > For the more computationally efficient water models, we have also
358 > investigated the effect of potential trunctaion on the computed free
359 > energies as a function of the cutoff radius.  As seen in
360 > Fig. \ref{incCutoff}, the free energies of the ice polymorphs with
361 > water models lacking a long-range correction show significant cutoff
362 > dependence.  In general, there is a narrowing of the free energy
363 > differences while moving to greater cutoff radii.  As the free
364 > energies for the polymorphs converge, the stability advantage that
365 > Ice-{\it i} exhibits is reduced.  Adjacent to each of these plots are
366 > results for systems with applied or estimated long-range corrections.
367 > SSD/RF was parametrized for use with a reaction field, and the benefit
368 > provided by this computationally inexpensive correction is apparent.
369 > The free energies are largely independent of the size of the reaction
370 > field cavity in this model, so small cutoff radii mimic bulk
371 > calculations quite well under SSD/RF.
372 >
373 > Although TIP3P was paramaterized for use without the Ewald summation,
374 > we have estimated the effect of this method for computing long-range
375 > electrostatics for both TIP3P and SPC/E.  This was accomplished by
376 > calculating the potential energy of identical crystals both with and
377 > without particle mesh Ewald (PME).  Similar behavior to that observed
378 > with reaction field is seen for both of these models.  The free
379 > energies show reduced dependence on cutoff radius and span a narrower
380 > range for the various polymorphs.  Like the dipolar water models,
381 > TIP3P displays a relatively constant preference for the Ice-{\it i}
382 > polymorph.  Crystal preference is much more difficult to determine for
383 > SPC/E.  Without a long-range correction, each of the polymorphs
384 > studied assumes the role of the preferred polymorph under different
385 > cutoff radii.  The inclusion of the Ewald correction flattens and
386 > narrows the gap in free energies such that the polymorphs are
387 > isoenergetic within statistical uncertainty.  This suggests that other
388 > conditions, such as the density in fixed-volume simulations, can
389 > influence the polymorph expressed upon crystallization.
390  
391 < To further study the changes resulting to the inclusion of a
421 < long-range interaction correction, the effect of an Ewald summation
422 < was estimated by applying the potential energy difference do to its
423 < inclusion in systems in the presence and absence of the
424 < correction. This was accomplished by calculation of the potential
425 < energy of identical crystals both with and without PME. The free
426 < energies for the investigated polymorphs using the TIP3P and SPC/E
427 < water models are shown in Table \ref{pmeShift}. The same trend pointed
428 < out through increase of cutoff radius is observed in these PME
429 < results. Ice-{\it i} is the preferred polymorph at ambient conditions
430 < for both the TIP3P and SPC/E water models; however, the narrowing of
431 < the free energy differences between the various solid forms is
432 < significant enough that it becomes less clear that it is the most
433 < stable polymorph with the SPC/E model.  The free energies of Ice-{\it
434 < i} and ice B nearly overlap within error, with ice $I_c$ just outside
435 < as well, indicating that Ice-{\it i} might be metastable with respect
436 < to ice B and possibly ice $I_c$ with SPC/E. However, these results do
437 < not significantly alter the finding that the Ice-{\it i} polymorph is
438 < a stable crystal structure that should be considered when studying the
439 < phase behavior of water models.
391 > \section{Conclusions}
392  
393 < \begin{table*}
394 < \begin{minipage}{\linewidth}
395 < \renewcommand{\thefootnote}{\thempfootnote}
396 < \begin{center}
397 < \caption{The free energy of the studied ice polymorphs after applying
398 < the energy difference attributed to the inclusion of the PME
399 < long-range interaction correction. Units are kcal/mol.}
400 < \begin{tabular}{ l  c  c  c  c }
449 < \hline
450 < \ \ Water Model \ \ & \ \ \ \ \ $I_h$ \ \ & \ \ \ \ \ $I_c$ \ \ & \ \quad \ \ \ \ B \ \ & \ \ \ \ \ Ice-{\it i} \ \ \\
451 < \hline
452 < TIP3P  & -11.53(2) & -11.24(3) & -11.51(3) & -11.67(3)\\
453 < SPC/E  & -12.77(2) & -12.92(2) & -12.96(3) & -13.02(2)\\
454 < \end{tabular}
455 < \label{pmeShift}
456 < \end{center}
457 < \end{minipage}
458 < \end{table*}
393 > In this report, thermodynamic integration was used to determine the
394 > absolute free energies of several ice polymorphs.  Of the studied
395 > crystal forms, Ice-{\it i} was observed to be the stable crystalline
396 > state for {\it all} the water models when using a 9.0 \AA\
397 > intermolecular interaction cutoff.  Through investigation of possible
398 > interaction truncation methods, the free energy was shown to be
399 > partially dependent on simulation conditions; however, Ice-{\it i} was
400 > still observered to be a stable polymorph of the studied water models.
401  
402 < \section{Conclusions}
402 > So what is the preferred solid polymorph for simulated water?  As
403 > indicated above, the answer appears to be dependent both on the
404 > conditions and the model used.  In the case of short cutoffs without a
405 > long-range interaction correction, Ice-{\it i} and Ice-{\it
406 > i}$^\prime$ have the lowest free energy of the studied polymorphs with
407 > all the models.  Ideally, crystallization of each model under constant
408 > pressure conditions, as was done with SSD/E, would aid in the
409 > identification of their respective preferred structures.  This work,
410 > however, helps illustrate how studies involving one specific model can
411 > lead to insight about important behavior of others.  In general, the
412 > above results support the finding that the Ice-{\it i} polymorph is a
413 > stable crystal structure that should be considered when studying the
414 > phase behavior of water models.
415  
416 < The free energy for proton ordered variants of hexagonal and cubic ice
417 < $I$, ice B, and our recently discovered Ice-{\it i} structure were
418 < calculated under standard conditions for several common water models
419 < via thermodynamic integration. All the water models studied show
420 < Ice-{\it i} to be the minimum free energy crystal structure with a 9
421 < \AA\ switching function cutoff. Calculated melting and boiling points
468 < show surprisingly good agreement with the experimental values;
469 < however, the solid phase at 1 atm is Ice-{\it i}, not ice $I_h$. The
470 < effect of interaction truncation was investigated through variation of
471 < the cutoff radius, use of a reaction field parameterized model, and
472 < estimation of the results in the presence of the Ewald
473 < summation. Interaction truncation has a significant effect on the
474 < computed free energy values, and may significantly alter the free
475 < energy landscape for the more complex multipoint water models. Despite
476 < these effects, these results show Ice-{\it i} to be an important ice
477 < polymorph that should be considered in simulation studies.
416 > We also note that none of the water models used in this study are
417 > polarizable or flexible models.  It is entirely possible that the
418 > polarizability of real water makes Ice-{\it i} substantially less
419 > stable than ice $I_h$.  However, the calculations presented above seem
420 > interesting enough to communicate before the role of polarizability
421 > (or flexibility) has been thoroughly investigated.
422  
423 < Due to this relative stability of Ice-{\it i} in all of the
424 < investigated simulation conditions, the question arises as to possible
425 < experimental observation of this polymorph.  The rather extensive past
426 < and current experimental investigation of water in the low pressure
427 < regime makes us hesitant to ascribe any relevance of this work outside
428 < of the simulation community.  It is for this reason that we chose a
429 < name for this polymorph which involves an imaginary quantity.  That
430 < said, there are certain experimental conditions that would provide the
431 < most ideal situation for possible observation. These include the
432 < negative pressure or stretched solid regime, small clusters in vacuum
423 > Finally, due to the stability of Ice-{\it i} in the investigated
424 > simulation conditions, the question arises as to possible experimental
425 > observation of this polymorph.  The rather extensive past and current
426 > experimental investigation of water in the low pressure regime makes
427 > us hesitant to ascribe any relevance to this work outside of the
428 > simulation community.  It is for this reason that we chose a name for
429 > this polymorph which involves an imaginary quantity.  That said, there
430 > are certain experimental conditions that would provide the most ideal
431 > situation for possible observation. These include the negative
432 > pressure or stretched solid regime, small clusters in vacuum
433   deposition environments, and in clathrate structures involving small
434 < non-polar molecules.  Figs. \ref{fig:gofr} and \ref{fig:sofq} contain
435 < our predictions for both the pair distribution function ($g_{OO}(r)$)
436 < and the structure factor ($S(\vec{q})$ for ice $I_h$, $I_c$, and for
437 < ice-{\it i} at a temperature of 77K.  In studies of the high and low
438 < density forms of amorphous ice, ``spurious'' diffraction peaks have
495 < been observed experimentally.\cite{Bizid87} It is possible that a
496 < variant of Ice-{\it i} could explain some of this behavior; however,
497 < we will leave it to our experimental colleagues to make the final
498 < determination on whether this ice polymorph is named appropriately
499 < (i.e. with an imaginary number) or if it can be promoted to Ice-0.
434 > non-polar molecules.  For experimental comparison purposes, example
435 > $g_{OO}(r)$ and $S(\vec{q})$ plots were generated for the two Ice-{\it
436 > i} variants (along with example ice $I_h$ and $I_c$ plots) at 77K, and
437 > they are shown in figures \ref{fig:gofr} and \ref{fig:sofq}
438 > respectively.
439  
440   \begin{figure}
441 + \centering
442   \includegraphics[width=\linewidth]{iceGofr.eps}
443 < \caption{Radial distribution functions of ice $I_h$, $I_c$,
444 < Ice-{\it i}, and Ice-{\it i}$^\prime$ calculated from from simulations
445 < of the SSD/RF water model at 77 K.}
443 > \caption{Radial distribution functions of ice $I_h$, $I_c$, and
444 > Ice-{\it i} calculated from from simulations of the SSD/RF water model
445 > at 77 K.  The Ice-{\it i} distribution function was obtained from
446 > simulations composed of TIP4P water.}
447   \label{fig:gofr}
448   \end{figure}
449  
450   \begin{figure}
451 + \centering
452   \includegraphics[width=\linewidth]{sofq.eps}
453   \caption{Predicted structure factors for ice $I_h$, $I_c$, Ice-{\it i},
454   and Ice-{\it i}$^\prime$ at 77 K.  The raw structure factors have

Diff Legend

Removed lines
+ Added lines
< Changed lines
> Changed lines