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 1457 by chrisfen, Tue Sep 14 23:03:53 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[preprint,aps,endfloats]{revtex4}
3 < %\documentclass[11pt]{article}
4 < %\usepackage{endfloat}
2 > \documentclass[11pt]{article}
3 > \usepackage{endfloat}
4   \usepackage{amsmath}
5   \usepackage{epsf}
6   \usepackage{berkeley}
7 < %\usepackage{setspace}
8 < %\usepackage{tabularx}
7 > \usepackage{setspace}
8 > \usepackage{tabularx}
9   \usepackage{graphicx}
10 < %\usepackage[ref]{overcite}
11 < %\pagestyle{plain}
12 < %\pagenumbering{arabic}
13 < %\oddsidemargin 0.0cm \evensidemargin 0.0cm
14 < %\topmargin -21pt \headsep 10pt
15 < %\textheight 9.0in \textwidth 6.5in
16 < %\brokenpenalty=10000
10 > \usepackage[ref]{overcite}
11 > \pagestyle{plain}
12 > \pagenumbering{arabic}
13 > \oddsidemargin 0.0cm \evensidemargin 0.0cm
14 > \topmargin -21pt \headsep 10pt
15 > \textheight 9.0in \textwidth 6.5in
16 > \brokenpenalty=10000
17 > \renewcommand{\baselinestretch}{1.2}
18 > \renewcommand\citemid{\ } % no comma in optional reference note
19  
19 %\renewcommand\citemid{\ } % no comma in optional reference note
20
20   \begin{document}
21  
22 < \title{A Free Energy Study of Low Temperature and Anomolous Ice}
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{\thefootnote}
26 < \footnote[1]{Corresponding author. \ Electronic mail: gezelter@nd.edu}}
27 <
28 < \address{Department of Chemistry and Biochemistry\\ University of Notre Dame\\
25 > \author{Christopher J. Fennell and J. Daniel Gezelter \\
26 > Department of Chemistry and Biochemistry\\
27 > University of Notre Dame\\
28   Notre Dame, Indiana 46556}
29  
30   \date{\today}
31  
32 < %\maketitle
32 > \maketitle
33   %\doublespacing
34  
35   \begin{abstract}
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  
39 \maketitle
40
41 \newpage
42
48   %\narrowtext
49  
50   %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
# Line 48 | Line 53 | Notre Dame, Indiana 46556}
53  
54   \section{Introduction}
55  
56 + Water has proven to be a challenging substance to depict in
57 + simulations, and a variety of models have been developed to describe
58 + its behavior under varying simulation
59 + conditions.\cite{Stillinger74,Rahman75,Berendsen81,Jorgensen83,Bratko85,Berendsen87,Caldwell95,Liu96,Berendsen98,Dill00,Mahoney00,Fennell04}
60 + These models have been used to investigate important physical
61 + phenomena like phase transitions, transport properties, and 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
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.  
77 +
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{(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{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
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 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 (Object-Oriented Parallel Simulation Engine)
154 < molecular mechanics package. All molecules were treated as rigid
155 < bodies, with orientational motion propogated using the symplectic DLM
156 < integration method. Details about the implementation of these
157 < techniques can be found in a recent publication.\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
156 > the implementation of this technique can be found in a recent
157 > publication.\cite{Dullweber1997}
158  
159 < Thermodynamic integration was utilized to calculate the free energy of
160 < several ice crystals at 200 K using the TIP3P, TIP4P, TIP5P, SPC/E,
161 < SSD/RF, and SSD/E water models. Liquid state free energies at 300 and
162 < 400 K for all of these water models were also determined using this
163 < same technique, in order to determine melting points and generate
164 < phase diagrams. All simulations were carried out at densities
165 < resulting in a pressure of approximately 1 atm at their respective
166 < temperatures.
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 > A = A_0 + \int_0^1 \left\langle \Delta V \right\rangle_\lambda d\lambda.
174 > \end{equation}
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 is chosen as the reference state that the system is converted
183 < to over the course of the simulation. In an Einstein Crystal, the
184 < molecules are harmonically restrained at their ideal lattice locations
185 < and orientations. The partition function for a molecular crystal
186 < restrained in this fashion has been evaluated, and the Helmholtz Free
187 < Energy ({\it A}) is given by
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},
188 > \end{equation}
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.  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}
202   A = E_m\ -\ kT\ln \left (\frac{kT}{h\nu}\right )^3&-&kT\ln \left
203   [\pi^\frac{1}{2}\left (\frac{8\pi^2I_\mathrm{A}kT}{h^2}\right
# Line 84 | Line 209 | where $2\pi\nu = (K_\mathrm{v}/m)^{1/2}$.\cite{Baez95a
209   )^\frac{1}{2}}\exp(t^2)\mathrm{d}t\right ],
210   \label{ecFreeEnergy}
211   \end{eqnarray}
212 < where $2\pi\nu = (K_\mathrm{v}/m)^{1/2}$.\cite{Baez95a} In equation
213 < \ref{ecFreeEnergy}, $K_\mathrm{v}$, $K_\mathrm{\theta}$, and
214 < $K_\mathrm{\omega}$ are the spring constants restraining translational
90 < motion and deflection of and rotation around the principle axis of the
91 < molecule respectively (Fig. \ref{waterSpring}), and $E_m$ is the
92 < minimum potential energy of the ideal crystal. In the case of
93 < molecular liquids, the ideal vapor is chosen as the target reference
94 < state.
212 > where $2\pi\nu = (K_\mathrm{v}/m)^{1/2}$, and $E_m$ is the minimum
213 > potential energy of the ideal crystal.\cite{Baez95a}
214 >
215   \begin{figure}
216 < \includegraphics[scale=1.0]{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}
225   \end{figure}
226  
227 < Charge, dipole, and Lennard-Jones interactions were modified by a
228 < cubic switching between 100\% and 85\% of the cutoff value (9 \AA ). By
229 < applying this function, these interactions are smoothly truncated,
230 < thereby avoiding poor energy conserving dynamics resulting from
231 < harsher truncation schemes. The effect of a long-range correction was
232 < also investigated on select model systems in a variety of manners. For
233 < the SSD/RF model, a reaction field with a fixed dielectric constant of
234 < 80 was applied in all simulations.\cite{Onsager36} For a series of the
235 < least computationally expensive models (SSD/E, SSD/RF, and TIP3P),
236 < simulations were performed with longer cutoffs of 12 and 15 \AA\ to
237 < compare with the 9 \AA\ cutoff results. Finally, results from the use
238 < of an Ewald summation were estimated for TIP3P and SPC/E by performing
118 < calculations with Particle-Mesh Ewald (PME) in the TINKER molecular
119 < mechanics software package. TINKER was chosen because it can also
120 < propogate the motion of rigid-bodies, and provides the most direct
121 < comparison to the results from OOPSE. The calculated energy difference
122 < in the presence and absence of PME was applied to the previous results
123 < in order to predict changes in the free energy landscape.
227 > In the case of molecular liquids, the ideal vapor is chosen as the
228 > target reference state.  There are several examples of liquid state
229 > free energy calculations of water models present in the
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 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 < \section{Results and discussion}
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
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, 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 < The free energy of proton ordered Ice-{\it i} was calculated and
128 < compared with the free energies of proton ordered variants of the
129 < experimentally observed low-density ice polymorphs, $I_h$ and $I_c$,
130 < as well as the higher density ice B, observed by B\`{a}ez and Clancy
131 < and thought to be the minimum free energy structure for the SPC/E
132 < model at ambient conditions (Table \ref{freeEnergy}).\cite{Baez95b}
133 < Ice XI, the experimentally observed proton ordered variant of ice
134 < $I_h$, was investigated initially, but it was found not to be as
135 < stable as antiferroelectric variants of proton ordered or even proton
136 < disordered ice$I_h$.\cite{Davidson84} The proton ordered variant of
137 < ice $I_h$ used here is a simple antiferroelectric version that has an
138 < 8 molecule unit cell. The crystals contained 648 or 1728 molecules for
139 < ice B, 1024 or 1280 molecules for ice $I_h$, 1000 molecules for ice
140 < $I_c$, or 1024 molecules for Ice-{\it i}. The larger crystal sizes
141 < were necessary for simulations involving larger cutoff values.
259 > \section{Results and Discussion}
260  
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}
145 \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. *Ice $I_c$ is unstable at 200 K using SSD/RF.}
291 < \begin{tabular}{ l  c  c  c  c }
292 < \hline \\[-7mm]
293 < \ \quad \ Water Model\ \ & \ \quad \ \ \ \ $I_h$ \ \ & \ \quad \ \ \ \ $I_c$ \ \  & \ \quad \ \ \ \ B \ \  & \ \quad \ \ \ Ice-{\it i} \ \quad \ \\
294 < \hline \\[-3mm]
295 < \ \quad \ TIP3P  & \ \quad \ -11.41 & \ \quad \ -11.23 & \ \quad \ -11.82 & \quad -12.30\\
296 < \ \quad \ TIP4P  & \ \quad \ -11.84 & \ \quad \ -12.04 & \ \quad \ -12.08 & \quad -12.33\\
297 < \ \quad \ TIP5P  & \ \quad \ -11.85 & \ \quad \ -11.86 & \ \quad \ -11.96 & \quad -12.29\\
298 < \ \quad \ SPC/E  & \ \quad \ -12.67 & \ \quad \ -12.96 & \ \quad \ -13.25 & \quad -13.55\\
299 < \ \quad \ SSD/E  & \ \quad \ -11.27 & \ \quad \ -11.19 & \ \quad \ -12.09 & \quad -12.54\\
300 < \ \quad \ SSD/RF & \ \quad \ -11.51 & \ \quad \ NA* & \ \quad \ -12.08 & \quad -12.29\\
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 > 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  
167 The free energy values computed for the studied polymorphs indicate
168 that Ice-{\it i} is the most stable state for all of the common water
169 models studied. With the free energy at these state points, the
170 temperature and pressure dependence of the free energy was used to
171 project to other state points and build phase diagrams. Figures
172 \ref{tp3phasedia} and \ref{ssdrfphasedia} are example diagrams built
173 from the free energy results. All other models have similar structure,
174 only the crossing points between these phases exist at different
175 temperatures and pressures. It is interesting to note that ice $I$
176 does not exist in either cubic or hexagonal form in any of the phase
177 diagrams for any of the models. For purposes of this study, ice B is
178 representative of the dense ice polymorphs. A recent study by Sanz
179 {\it et al.} goes into detail on the phase diagrams for SPC/E and
180 TIP4P in the high pressure regime.\cite{Sanz04}
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 + 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 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]{ssdrfPhaseDia.eps}
346 < \caption{Phase diagram for the SSD/RF water model in the low pressure
347 < regime. Calculations producing these results were done under an
348 < applied reaction field. It is interesting to note that this
349 < computationally efficient model (over 3 times more efficient than
350 < TIP3P) exhibits phase behavior similar to the less computationally
351 < conservative charge based models.}
352 < \label{ssdrfphasedia}
345 > \includegraphics[width=\linewidth]{cutoffChange.eps}
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 < \begin{table*}
358 < \begin{minipage}{\linewidth}
359 < \renewcommand{\thefootnote}{\thempfootnote}
360 < \begin{center}
361 < \caption{Melting ($T_m$), boiling ($T_b$), and sublimation ($T_s$)
362 < temperatures of several common water models compared with experiment.}
363 < \begin{tabular}{ l  c  c  c  c  c  c  c }
364 < \hline \\[-7mm]
365 < \ \ Equilibria Point\ \ & \ \ \ \ \ TIP3P \ \ & \ \ \ \ \ TIP4P \ \ & \ \quad \ \ \ \ TIP5P \ \ & \ \ \ \ \ SPC/E \ \ & \ \ \ \ \ SSD/E \ \ & \ \ \ \ \ SSD/RF \ \ & \ \ \ \ \ Exp. \ \ \\
366 < \hline \\[-3mm]
367 < \ \ $T_m$ (K)  & \ \ 269 & \ \ 265 & \ \ 271 &  297 & \ \ - & \ \ 278 & \ \ 273\\
368 < \ \ $T_b$ (K)  & \ \ 357 & \ \ 354 & \ \ 337 &  396 & \ \ - & \ \ 349 & \ \ 373\\
369 < \ \ $T_s$ (K)  & \ \ - & \ \ - & \ \ - &  - & \ \ 355 & \ \ - & \ \ -\\
370 < \end{tabular}
371 < \label{meltandboil}
372 < \end{center}
373 < \end{minipage}
374 < \end{table*}
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 < Table \ref{meltandboil} lists the melting and boiling temperatures
221 < calculated from this work. Surprisingly, most of these models have
222 < melting points that compare quite favorably with experiment. The
223 < unfortunate aspect of this result is that this phase change occurs
224 < between Ice-{\it i} and the liquid state rather than ice $I_h$ and the
225 < liquid state. These results are actually not contrary to previous
226 < studies in the literature. Earlier free energy studies of ice $I$
227 < using TIP4P predict a $T_m$ ranging from 214 to 238 K (differences
228 < being attributed to choice of interaction truncation and different
229 < ordered and disordered molecular arrangements). If the presence of ice
230 < B and Ice-{\it i} were omitted, a $T_m$ value around 210 K would be
231 < predicted from this work. However, the $T_m$ from Ice-{\it i} is
232 < calculated at 265 K, significantly higher in temperature than the
233 < previous studies. Also of interest in these results is that SSD/E does
234 < not exhibit a melting point at 1 atm, but it shows a sublimation point
235 < at 355 K. This is due to the significant stability of Ice-{\it i} over
236 < all other polymorphs for this particular model under these
237 < conditions. While troubling, this behavior turned out to be
238 < advantagious in that it facilitated the spontaneous crystallization of
239 < Ice-{\it i}. These observations provide a warning that simulations of
240 < SSD/E as a ``liquid'' near 300 K are actually metastable and run the
241 < risk of spontaneous crystallization. However, this risk changes when
242 < applying a longer cutoff.
391 > \section{Conclusions}
392  
393 < Increasing the cutoff radius in simulations of the more
394 < computationally efficient water models was done in order to evaluate
395 < the trend in free energy values when moving to systems that do not
396 < involve potential truncation. As seen in Fig. \ref{incCutoff}, the
397 < free energy of all the ice polymorphs show a substantial dependence on
398 < cutoff radius. In general, there is a narrowing of the free energy
399 < differences while moving to greater cutoff radius. This trend is much
400 < more subtle in the case of SSD/RF, indicating that the free energies
252 < calculated with a reaction field present provide a more accurate
253 < picture of the free energy landscape in the absence of potential
254 < truncation.
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 < To further study the changes resulting to the inclusion of a
403 < long-range interaction correction, the effect of an Ewald summation
404 < was estimated by applying the potential energy difference do to its
405 < inclusion in systems in the presence and absence of the
406 < correction. This was accomplished by calculation of the potential
407 < energy of identical crystals with and without PME using TINKER. The
408 < free energies for the investigated polymorphs using the TIP3P and
409 < SPC/E water models are shown in Table \ref{pmeShift}. TIP4P and TIP5P
410 < are not fully supported in TINKER, so the results for these models
411 < could not be estimated. The same trend pointed out through increase of
412 < cutoff radius is observed in these results. Ice-{\it i} is the
413 < preferred polymorph at ambient conditions for both the TIP3P and SPC/E
414 < water models; however, there is a narrowing of the free energy
269 < differences between the various solid forms. In the case of SPC/E this
270 < narrowing is significant enough that it becomes less clear cut that
271 < Ice-{\it i} is the most stable polymorph, and is possibly metastable
272 < with respect to ice B and possibly ice $I_c$. However, these results
273 < do not significantly alter the finding that the Ice-{\it i} polymorph
274 < is a stable crystal structure that should be considered when studying
275 < the phase behavior of water models.
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 < \begin{table*}
417 < \begin{minipage}{\linewidth}
418 < \renewcommand{\thefootnote}{\thempfootnote}
419 < \begin{center}
420 < \caption{The free energy of the studied ice polymorphs after applying the energy difference attributed to the inclusion of the PME long-range interaction correction. Units are kcal/mol.}
421 < \begin{tabular}{ l  c  c  c  c }
283 < \hline \\[-7mm]
284 < \ \ Water Model \ \ & \ \ \ \ \ $I_h$ \ \ & \ \ \ \ \ $I_c$ \ \ & \ \quad \ \ \ \ B \ \ & \ \ \ \ \ Ice-{\it i} \ \ \\
285 < \hline \\[-3mm]
286 < \ \ TIP3P  & \ \ -11.53 & \ \ -11.24 & \ \ -11.51 & \ \ -11.67\\
287 < \ \ SPC/E  & \ \ -12.77 & \ \ -12.92 & \ \ -12.96 & \ \ -13.02\\
288 < \end{tabular}
289 < \label{pmeShift}
290 < \end{center}
291 < \end{minipage}
292 < \end{table*}
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 < \section{Conclusions}
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.  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 $I_h$, $I_c$, 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
455 + been convoluted with a gaussian instrument function (0.075 \AA$^{-1}$
456 + width) to compensate for the trunction effects in our finite size
457 + simulations.}
458 + \label{fig:sofq}
459 + \end{figure}
460 +
461   \section{Acknowledgments}
462   Support for this project was provided by the National Science
463   Foundation under grant CHE-0134881. Computation time was provided by
464 < the Notre Dame Bunch-of-Boxes (B.o.B) computer cluster under NSF grant
465 < DMR-0079647.
464 > the Notre Dame High Performance Computing Cluster and the Notre Dame
465 > Bunch-of-Boxes (B.o.B) computer cluster (NSF grant DMR-0079647).
466  
467   \newpage
468  

Diff Legend

Removed lines
+ Added lines
< Changed lines
> Changed lines