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 1474 by chrisfen, Fri Sep 17 14:56:05 2004 UTC vs.
Revision 1834 by chrisfen, Thu Dec 2 18:58:25 2004 UTC

# Line 20 | Line 20
20  
21   \begin{document}
22  
23 < \title{Ice-{\it i}: a simulation-predicted ice polymorph which is more
24 < stable than Ice $I_h$ for point-charge and point-dipole water models}
23 > \title{Free Energy Analysis of Simulated Ice Polymorphs Using Simple
24 > Dipolar and Charge Based Water Models}
25  
26   \author{Christopher J. Fennell and J. Daniel Gezelter \\
27   Department of Chemistry and Biochemistry\\ University of Notre Dame\\
# Line 33 | Line 33 | The free energies of several ice polymorphs in the low
33   %\doublespacing
34  
35   \begin{abstract}
36 < The free energies of several ice polymorphs in the low pressure regime
37 < were calculated using thermodynamic integration.  These integrations
38 < were done for most of the common water models. Ice-{\it i}, a
39 < structure we recently observed to be stable in one of the single-point
40 < water models, was determined to be the stable crystalline state (at 1
41 < atm) for {\it all} the water models investigated.  Phase diagrams were
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 the common water
46 < models.  Additionally, potential truncation was shown to have an
47 < effect on the calculated free energies, and can result in altered free
48 < energy landscapes.
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.
51   \end{abstract}
52  
53   %\narrowtext
# Line 54 | Line 58 | Computer simulations are a valuable tool for studying
58  
59   \section{Introduction}
60  
57 Computer simulations are a valuable tool for studying the phase
58 behavior of systems ranging from small or simple molecules to complex
59 biological species.\cite{Matsumoto02,Li96,Marrink01} Useful techniques
60 have been developed to investigate the thermodynamic properites of
61 model substances, providing both qualitative and quantitative
62 comparisons between simulations and
63 experiment.\cite{Widom63,Frenkel84} Investigation of these properties
64 leads to the development of new and more accurate models, leading to
65 better understanding and depiction of physical processes and intricate
66 molecular systems.
67
61   Water has proven to be a challenging substance to depict in
62   simulations, and a variety of models have been developed to describe
63   its behavior under varying simulation
64 < conditions.\cite{Rahman75,Berendsen81,Jorgensen83,Bratko85,Berendsen87,Liu96,Berendsen98,Mahoney00,Fennell04}
64 > conditions.\cite{Stillinger74,Rahman75,Berendsen81,Jorgensen83,Bratko85,Berendsen87,Caldwell95,Liu96,Berendsen98,Dill00,Mahoney00,Fennell04}
65   These models have been used to investigate important physical
66 < phenomena like phase transitions, molecule transport, and the
66 > phenomena like phase transitions, transport properties, and the
67   hydrophobic effect.\cite{Yamada02,Marrink94,Gallagher03} With the
68   choice of models available, it is only natural to compare the models
69   under interesting thermodynamic conditions in an attempt to clarify
70   the limitations of each of the
71   models.\cite{Jorgensen83,Jorgensen98b,Clancy94,Mahoney01} Two
72 < important property to quantify are the Gibbs and Helmholtz free
72 > important properties to quantify are the Gibbs and Helmholtz free
73   energies, particularly for the solid forms of water.  Difficulty in
74   these types of studies typically arises from the assortment of
75   possible crystalline polymorphs that water adopts over a wide range of
# Line 84 | Line 77 | these phases will dictate the true transition temperat
77   of ice, and it is a challenging task to investigate the entire free
78   energy landscape.\cite{Sanz04} Ideally, research is focused on the
79   phases having the lowest free energy at a given state point, because
80 < these phases will dictate the true transition temperatures and
80 > these phases will dictate the relevant transition temperatures and
81   pressures for the model.
82  
83   In this paper, standard reference state methods were applied to known
84   crystalline water polymorphs in the low pressure regime.  This work is
85 < unique in the fact that one of the crystal lattices was arrived at
86 < through crystallization of a computationally efficient water model
87 < under constant pressure and temperature conditions. Crystallization
88 < events are interesting in and of
89 < themselves;\cite{Matsumoto02,Yamada02} however, the crystal structure
90 < obtained in this case is different from any previously observed ice
91 < polymorphs in experiment or simulation.\cite{Fennell04} We have named
92 < this structure Ice-{\it i} to indicate its origin in computational
93 < simulation. The unit cell (Fig. \ref{iceiCell}A) consists of eight
94 < water molecules that stack in rows of interlocking water
95 < tetramers. Proton ordering can be accomplished by orienting two of the
96 < molecules so that both of their donated hydrogen bonds are internal to
97 < their tetramer (Fig. \ref{protOrder}). As expected in an ice crystal
98 < constructed of water tetramers, the hydrogen bonds are not as linear
99 < as those observed in ice $I_h$, however the interlocking of these
100 < subunits appears to provide significant stabilization to the overall
101 < crystal. The arrangement of these tetramers results in surrounding
102 < open octagonal cavities that are typically greater than 6.3 \AA\ in
103 < diameter. This relatively open overall structure leads to crystals
85 > unique in that one of the crystal lattices was arrived at through
86 > crystallization of a computationally efficient water model under
87 > constant pressure and temperature conditions.  Crystallization events
88 > are interesting in and of themselves;\cite{Matsumoto02,Yamada02}
89 > however, the crystal structure obtained in this case is different from
90 > any previously observed ice polymorphs in experiment or
91 > simulation.\cite{Fennell04} We have named this structure Ice-{\it i}
92 > to indicate its origin in computational simulation. The unit cell
93 > (Fig. \ref{iceiCell}A) consists of eight water molecules that stack in
94 > rows of interlocking water tetramers.  This crystal structure has a
95 > limited resemblence to a recent two-dimensional ice tessellation
96 > simulated on a silica surface.\cite{Yang04} Proton ordering can be
97 > accomplished by orienting two of the molecules so that both of their
98 > donated hydrogen bonds are internal to their tetramer
99 > (Fig. \ref{protOrder}).  As expected in an ice crystal constructed of
100 > water tetramers, the hydrogen bonds are not as linear as those
101 > observed in ice $I_h$, however the interlocking of these subunits
102 > appears to provide significant stabilization to the overall crystal.
103 > The arrangement of these tetramers results in surrounding open
104 > octagonal cavities that are typically greater than 6.3 \AA\ in
105 > diameter.  This relatively open overall structure leads to crystals
106   that are 0.07 g/cm$^3$ less dense on average than ice $I_h$.
107  
108   \begin{figure}
109   \includegraphics[width=\linewidth]{unitCell.eps}
110 < \caption{Unit cells for (A) Ice-{\it i} and (B) Ice-$i^\prime$, the
111 < elongated variant of Ice-{\it i}.  The spheres represent the
110 > \caption{Unit cells for (A) Ice-{\it i} and (B) Ice-{\it i}$^\prime$,
111 > the elongated variant of Ice-{\it i}.  The spheres represent the
112   center-of-mass locations of the water molecules.  The $a$ to $c$
113   ratios for Ice-{\it i} and Ice-{\it i}$^\prime$ are given by
114   $a:2.1214c$ and $a:1.7850c$ respectively.}
# Line 123 | Line 118 | down the (001) crystal face. The rows of water tetrame
118   \begin{figure}
119   \includegraphics[width=\linewidth]{orderedIcei.eps}
120   \caption{Image of a proton ordered crystal of Ice-{\it i} looking
121 < down the (001) crystal face. The rows of water tetramers surrounded by
122 < octagonal pores leads to a crystal structure that is significantly
121 > down the (001) crystal face.  The rows of water tetramers surrounded
122 > by octagonal pores leads to a crystal structure that is significantly
123   less dense than ice $I_h$.}
124   \label{protOrder}
125   \end{figure}
126  
127   Results from our previous study indicated that Ice-{\it i} is the
128   minimum energy crystal structure for the single point water models we
129 < investigated (for discussions on these single point dipole models, see
130 < our previous work and related
129 > had investigated (for discussions on these single point dipole models,
130 > see our previous work and related
131   articles).\cite{Fennell04,Liu96,Bratko85} Those results only
132   considered energetic stabilization and neglected entropic
133 < contributions to the overall free energy. To address this issue, the
134 < absolute free energy of this crystal was calculated using
133 > contributions to the overall free energy.  To address this issue, we
134 > have calculated the absolute free energy of this crystal using
135   thermodynamic integration and compared to the free energies of cubic
136   and hexagonal ice $I$ (the experimental low density ice polymorphs)
137   and ice B (a higher density, but very stable crystal structure
# Line 145 | Line 140 | be noted that a second version of Ice-{\it i} (Ice-$i^
140   from which Ice-{\it i} was crystallized (SSD/E) in addition to several
141   common water models (TIP3P, TIP4P, TIP5P, and SPC/E) and a reaction
142   field parametrized single point dipole water model (SSD/RF). It should
143 < be noted that a second version of Ice-{\it i} (Ice-$i^\prime$) was used
144 < in calculations involving SPC/E, TIP4P, and TIP5P. The unit cell of
145 < this crystal (Fig. \ref{iceiCell}B) is similar to the Ice-{\it i} unit
146 < it is extended in the direction of the (001) face and compressed along
147 < the other two faces.
143 > be noted that a second version of Ice-{\it i} (Ice-{\it i}$^\prime$)
144 > was used in calculations involving SPC/E, TIP4P, and TIP5P.  The unit
145 > cell of this crystal (Fig. \ref{iceiCell}B) is similar to the Ice-{\it
146 > i} unit it is extended in the direction of the (001) face and
147 > compressed along the other two faces.  There is typically a small
148 > distortion of proton ordered Ice-{\it i}$^\prime$ that converts the
149 > normally square tetramer into a rhombus with alternating approximately
150 > 85 and 95 degree angles.  The degree of this distortion is model
151 > dependent and significant enough to split the tetramer diagonal
152 > location peak in the radial distribution function.
153  
154   \section{Methods}
155  
156   Canonical ensemble (NVT) molecular dynamics calculations were
157   performed using the OOPSE molecular mechanics package.\cite{Meineke05}
158   All molecules were treated as rigid bodies, with orientational motion
159 < propagated using the symplectic DLM integration method. Details about
159 > propagated using the symplectic DLM integration method.  Details about
160   the implementation of this technique can be found in a recent
161   publication.\cite{Dullweber1997}
162  
# Line 168 | Line 168 | this same technique in order to determine melting poin
168   crystals at 200 K using the TIP3P, TIP4P, TIP5P, SPC/E, SSD/RF, and
169   SSD/E water models.\cite{Baez95a} Liquid state free energies at 300
170   and 400 K for all of these water models were also determined using
171 < this same technique in order to determine melting points and generate
172 < phase diagrams. All simulations were carried out at densities
173 < resulting in a pressure of approximately 1 atm at their respective
174 < temperatures.
171 > this same technique in order to determine melting points and to
172 > generate phase diagrams.  All simulations were carried out at
173 > densities which correspond to a pressure of approximately 1 atm at
174 > their respective temperatures.
175  
176 < A single thermodynamic integration involves a sequence of simulations
177 < over which the system of interest is converted into a reference system
178 < for which the free energy is known analytically. This transformation
179 < path is then integrated in order to determine the free energy
180 < difference between the two states:
176 > Thermodynamic integration involves a sequence of simulations during
177 > which the system of interest is converted into a reference system for
178 > which the free energy is known analytically.  This transformation path
179 > is then integrated in order to determine the free energy difference
180 > between the two states:
181   \begin{equation}
182   \Delta A = \int_0^1\left\langle\frac{\partial V(\lambda
183   )}{\partial\lambda}\right\rangle_\lambda d\lambda,
184   \end{equation}
185   where $V$ is the interaction potential and $\lambda$ is the
186 < transformation parameter that scales the overall
187 < potential. Simulations are distributed strategically along this path
188 < in order to sufficiently sample the regions of greatest change in the
189 < potential. Typical integrations in this study consisted of $\sim$25
190 < simulations ranging from 300 ps (for the unaltered system) to 75 ps
191 < (near the reference state) in length.
186 > transformation parameter that scales the overall potential.
187 > Simulations are distributed strategically along this path in order to
188 > sufficiently sample the regions of greatest change in the potential.
189 > Typical integrations in this study consisted of $\sim$25 simulations
190 > ranging from 300 ps (for the unaltered system) to 75 ps (near the
191 > reference state) in length.
192  
193   For the thermodynamic integration of molecular crystals, the Einstein
194 < crystal was chosen as the reference system. In an Einstein crystal,
194 > crystal was chosen as the reference system.  In an Einstein crystal,
195   the molecules are restrained at their ideal lattice locations and
196   orientations. Using harmonic restraints, as applied by B\`{a}ez and
197   Clancy, the total potential for this reference crystal
# Line 203 | Line 203 | respectively.  It is clear from Fig. \ref{waterSpring}
203   where $K_\mathrm{v}$, $K_\mathrm{\theta}$, and $K_\mathrm{\omega}$ are
204   the spring constants restraining translational motion and deflection
205   of and rotation around the principle axis of the molecule
206 < respectively.  It is clear from Fig. \ref{waterSpring} that the values
207 < of $\theta$ range from $0$ to $\pi$, while $\omega$ ranges from
208 < $-\pi$ to $\pi$.  The partition function for a molecular crystal
206 > respectively.  These spring constants are typically calculated from
207 > the mean-square displacements of water molecules in an unrestrained
208 > ice crystal at 200 K.  For these studies, $K_\mathrm{r} = 4.29$ kcal
209 > mol$^{-1}$, $K_\theta\ = 13.88$ kcal mol$^{-1}$, and $K_\omega\ =
210 > 17.75$ kcal mol$^{-1}$.  It is clear from Fig. \ref{waterSpring} that
211 > the values of $\theta$ range from $0$ to $\pi$, while $\omega$ ranges
212 > from $-\pi$ to $\pi$.  The partition function for a molecular crystal
213   restrained in this fashion can be evaluated analytically, and the
214   Helmholtz Free Energy ({\it A}) is given by
215   \begin{eqnarray}
# Line 227 | Line 231 | body-frame {\it z}-axis. $K_\theta$ and $K_\omega$ are
231   \caption{Possible orientational motions for a restrained molecule.
232   $\theta$ angles correspond to displacement from the body-frame {\it
233   z}-axis, while $\omega$ angles correspond to rotation about the
234 < body-frame {\it z}-axis. $K_\theta$ and $K_\omega$ are spring
234 > body-frame {\it z}-axis.  $K_\theta$ and $K_\omega$ are spring
235   constants for the harmonic springs restraining motion in the $\theta$
236   and $\omega$ directions.}
237   \label{waterSpring}
# Line 239 | Line 243 | molecules.  In this study, we apply of one of the most
243   literature.\cite{Hermens88,Quintana92,Mezei92,Baez95b} These methods
244   typically differ in regard to the path taken for switching off the
245   interaction potential to convert the system to an ideal gas of water
246 < molecules.  In this study, we apply of one of the most convenient
247 < methods and integrate over the $\lambda^4$ path, where all interaction
248 < parameters are scaled equally by this transformation parameter.  This
249 < method has been shown to be reversible and provide results in
250 < excellent agreement with other established methods.\cite{Baez95b}
246 > molecules.  In this study, we applied of one of the most convenient
247 > methods and integrated over the $\lambda^4$ path, where all
248 > interaction parameters are scaled equally by this transformation
249 > parameter.  This method has been shown to be reversible and provide
250 > results in excellent agreement with other established
251 > methods.\cite{Baez95b}
252  
253   Charge, dipole, and Lennard-Jones interactions were modified by a
254 < cubic switching between 100\% and 85\% of the cutoff value (9 \AA
255 < ). By applying this function, these interactions are smoothly
256 < truncated, thereby avoiding the poor energy conservation which results
257 < from harsher truncation schemes. The effect of a long-range correction
258 < was also investigated on select model systems in a variety of
259 < manners. For the SSD/RF model, a reaction field with a fixed
260 < dielectric constant of 80 was applied in all
261 < simulations.\cite{Onsager36} For a series of the least computationally
262 < expensive models (SSD/E, SSD/RF, and TIP3P), simulations were
263 < performed with longer cutoffs of 12 and 15 \AA\ to compare with the 9
264 < \AA\ cutoff results. Finally, results from the use of an Ewald
265 < summation were estimated for TIP3P and SPC/E by performing
266 < calculations with Particle-Mesh Ewald (PME) in the TINKER molecular
267 < mechanics software package.\cite{Tinker} The calculated energy
268 < difference in the presence and absence of PME was applied to the
269 < previous results in order to predict changes to the free energy
265 < landscape.
254 > cubic switching between 100\% and 85\% of the cutoff value (9 \AA).
255 > By applying this function, these interactions are smoothly truncated,
256 > thereby avoiding the poor energy conservation which results from
257 > harsher truncation schemes.  The effect of a long-range correction was
258 > also investigated on select model systems in a variety of manners.
259 > For the SSD/RF model, a reaction field with a fixed dielectric
260 > constant of 80 was applied in all simulations.\cite{Onsager36} For a
261 > series of the least computationally expensive models (SSD/E, SSD/RF,
262 > and TIP3P), simulations were performed with longer cutoffs of 12 and
263 > 15 \AA\ to compare with the 9 \AA\ cutoff results.  Finally, the
264 > effects of utilizing an Ewald summation were estimated for TIP3P and
265 > SPC/E by performing single configuration calculations with
266 > Particle-Mesh Ewald (PME) in the TINKER molecular mechanics software
267 > package.\cite{Tinker} The calculated energy difference in the presence
268 > and absence of PME was applied to the previous results in order to
269 > predict changes to the free energy landscape.
270  
271   \section{Results and discussion}
272  
273 < The free energy of proton ordered Ice-{\it i} was calculated and
273 > The free energy of proton-ordered Ice-{\it i} was calculated and
274   compared with the free energies of proton ordered variants of the
275   experimentally observed low-density ice polymorphs, $I_h$ and $I_c$,
276   as well as the higher density ice B, observed by B\`{a}ez and Clancy
# Line 274 | Line 278 | as proton disordered or antiferroelectric variants of
278   model at ambient conditions (Table \ref{freeEnergy}).\cite{Baez95b}
279   Ice XI, the experimentally-observed proton-ordered variant of ice
280   $I_h$, was investigated initially, but was found to be not as stable
281 < as proton disordered or antiferroelectric variants of ice $I_h$. The
281 > as proton disordered or antiferroelectric variants of ice $I_h$.  The
282   proton ordered variant of ice $I_h$ used here is a simple
283 < antiferroelectric version that we divised, and it has an 8 molecule
283 > antiferroelectric version that we devised, and it has an 8 molecule
284   unit cell similar to other predicted antiferroelectric $I_h$
285   crystals.\cite{Davidson84} The crystals contained 648 or 1728
286   molecules for ice B, 1024 or 1280 molecules for ice $I_h$, 1000
287 < molecules for ice $I_c$, or 1024 molecules for Ice-{\it i}. The larger
288 < crystal sizes were necessary for simulations involving larger cutoff
289 < values.
287 > molecules for ice $I_c$, or 1024 molecules for Ice-{\it i}.  The
288 > larger crystal sizes were necessary for simulations involving larger
289 > cutoff values.
290  
291   \begin{table*}
292   \begin{minipage}{\linewidth}
289 \renewcommand{\thefootnote}{\thempfootnote}
293   \begin{center}
294 +
295   \caption{Calculated free energies for several ice polymorphs with a
296 < variety of common water models. All calculations used a cutoff radius
297 < of 9 \AA\ and were performed at 200 K and $\sim$1 atm. Units are
298 < kcal/mol. Calculated error of the final digits is in parentheses. *Ice
299 < $I_c$ rapidly converts to a liquid at 200 K with the SSD/RF model.}
300 < \begin{tabular}{ l  c  c  c  c }
296 > variety of common water models.  All calculations used a cutoff radius
297 > of 9.0 \AA\ and were performed at 200 K and $\sim$1 atm.  Units are
298 > kcal/mol.  Calculated error of the final digits is in parentheses.}
299 >
300 > \begin{tabular}{lccccc}
301   \hline
302 < Water Model & $I_h$ & $I_c$ & B & Ice-{\it i}\\
302 > Water Model & $I_h$ & $I_c$ & B & Ice-{\it i} & Ice-{\it i}$^\prime$\\
303   \hline
304 < TIP3P & -11.41(2) & -11.23(3) & -11.82(3) & -12.30(3)\\
305 < TIP4P & -11.84(3) & -12.04(2) & -12.08(3) & -12.33(3)\\
306 < TIP5P & -11.85(3) & -11.86(2) & -11.96(2) & -12.29(2)\\
307 < SPC/E & -12.67(2) & -12.96(2) & -13.25(3) & -13.55(2)\\
308 < SSD/E & -11.27(2) & -11.19(4) & -12.09(2) & -12.54(2)\\
309 < SSD/RF & -11.51(2) & NA* & -12.08(3) & -12.29(2)\\
304 > TIP3P & -11.41(2) & -11.23(3) & -11.82(3) & -12.30(3) & -\\
305 > TIP4P & -11.84(3) & -12.04(2) & -12.08(3) & - & -12.33(3)\\
306 > TIP5P & -11.85(3) & -11.86(2) & -11.96(2) & - & -12.29(2)\\
307 > SPC/E & -12.87(2) & -13.05(2) & -13.26(3) & - & -13.55(2)\\
308 > SSD/E & -11.27(2) & -11.19(4) & -12.09(2) & -12.54(2) & -\\
309 > SSD/RF & -11.51(2) & -11.47(2) & -12.08(3) & -12.29(2) & -\\
310   \end{tabular}
311   \label{freeEnergy}
312   \end{center}
# Line 311 | Line 315 | models studied. With the free energy at these state po
315  
316   The free energy values computed for the studied polymorphs indicate
317   that Ice-{\it i} is the most stable state for all of the common water
318 < models studied. With the free energy at these state points, the
319 < Gibbs-Helmholtz equation was used to project to other state points and
320 < to build phase diagrams.  Figures
321 < \ref{tp3phasedia} and \ref{ssdrfphasedia} are example diagrams built
322 < from the free energy results. All other models have similar structure,
323 < although the crossing points between the phases exist at slightly
324 < different temperatures and pressures. It is interesting to note that
325 < ice $I$ does not exist in either cubic or hexagonal form in any of the
326 < phase diagrams for any of the models. For purposes of this study, ice
327 < B is representative of the dense ice polymorphs. A recent study by
328 < Sanz {\it et al.} goes into detail on the phase diagrams for SPC/E and
329 < TIP4P in the high pressure regime.\cite{Sanz04}
318 > models studied.  With the calculated free energy at these state
319 > points, the Gibbs-Helmholtz equation was used to project to other
320 > state points and to build phase diagrams.  Figures \ref{tp3phasedia}
321 > and \ref{ssdrfphasedia} are example diagrams built from the free
322 > energy results.  All other models have similar structure, although the
323 > crossing points between the phases move to slightly different
324 > temperatures and pressures.  It is interesting to note that ice $I$
325 > does not exist in either cubic or hexagonal form in any of the phase
326 > diagrams for any of the models.  For purposes of this study, ice B is
327 > representative of the dense ice polymorphs.  A recent study by Sanz
328 > {\it et al.} goes into detail on the phase diagrams for SPC/E and
329 > TIP4P at higher pressures than those studied here.\cite{Sanz04}
330  
331   \begin{figure}
332   \includegraphics[width=\linewidth]{tp3PhaseDia.eps}
333   \caption{Phase diagram for the TIP3P water model in the low pressure
334 < regime. The displayed $T_m$ and $T_b$ values are good predictions of
334 > regime.  The displayed $T_m$ and $T_b$ values are good predictions of
335   the experimental values; however, the solid phases shown are not the
336 < experimentally observed forms. Both cubic and hexagonal ice $I$ are
336 > experimentally observed forms.  Both cubic and hexagonal ice $I$ are
337   higher in energy and don't appear in the phase diagram.}
338   \label{tp3phasedia}
339   \end{figure}
# Line 337 | Line 341 | regime. Calculations producing these results were done
341   \begin{figure}
342   \includegraphics[width=\linewidth]{ssdrfPhaseDia.eps}
343   \caption{Phase diagram for the SSD/RF water model in the low pressure
344 < regime. Calculations producing these results were done under an
345 < applied reaction field. It is interesting to note that this
344 > regime.  Calculations producing these results were done under an
345 > applied reaction field.  It is interesting to note that this
346   computationally efficient model (over 3 times more efficient than
347   TIP3P) exhibits phase behavior similar to the less computationally
348   conservative charge based models.}
# Line 347 | Line 351 | conservative charge based models.}
351  
352   \begin{table*}
353   \begin{minipage}{\linewidth}
350 \renewcommand{\thefootnote}{\thempfootnote}
354   \begin{center}
355 +
356   \caption{Melting ($T_m$), boiling ($T_b$), and sublimation ($T_s$)
357   temperatures at 1 atm for several common water models compared with
358 < experiment. The $T_m$ and $T_s$ values from simulation correspond to a
359 < transition between Ice-{\it i} (or Ice-{\it i}$^\prime$) and the
358 > experiment.  The $T_m$ and $T_s$ values from simulation correspond to
359 > a transition between Ice-{\it i} (or Ice-{\it i}$^\prime$) and the
360   liquid or gas state.}
361 < \begin{tabular}{ l  c  c  c  c  c  c  c }
361 >
362 > \begin{tabular}{lccccccc}
363   \hline
364 < Equilibria Point & TIP3P & TIP4P & TIP5P & SPC/E & SSD/E & SSD/RF & Exp.\\
364 > Equilibrium Point & TIP3P & TIP4P & TIP5P & SPC/E & SSD/E & SSD/RF & Exp.\\
365   \hline
366   $T_m$ (K)  & 269(4) & 266(5) & 271(4) & 296(3) & - & 278(4) & 273\\
367   $T_b$ (K)  & 357(2) & 354(2) & 337(2) & 396(2) & - & 348(2) & 373\\
# Line 368 | Line 373 | calculated from this work. Surprisingly, most of these
373   \end{table*}
374  
375   Table \ref{meltandboil} lists the melting and boiling temperatures
376 < calculated from this work. Surprisingly, most of these models have
377 < melting points that compare quite favorably with experiment. The
376 > calculated from this work.  Surprisingly, most of these models have
377 > melting points that compare quite favorably with experiment.  The
378   unfortunate aspect of this result is that this phase change occurs
379   between Ice-{\it i} and the liquid state rather than ice $I_h$ and the
380 < liquid state. These results are actually not contrary to previous
381 < studies in the literature. Earlier free energy studies of ice $I$
382 < using TIP4P predict a $T_m$ ranging from 214 to 238 K (differences
383 < being attributed to choice of interaction truncation and different
379 < ordered and disordered molecular
380 > liquid state.  These results are actually not contrary to other
381 > studies.  Studies of ice $I_h$ using TIP4P predict a $T_m$ ranging
382 > from 214 to 238 K (differences being attributed to choice of
383 > interaction truncation and different ordered and disordered molecular
384   arrangements).\cite{Vlot99,Gao00,Sanz04} If the presence of ice B and
385   Ice-{\it i} were omitted, a $T_m$ value around 210 K would be
386 < predicted from this work. However, the $T_m$ from Ice-{\it i} is
387 < calculated at 265 K, significantly higher in temperature than the
388 < previous studies. Also of interest in these results is that SSD/E does
386 > predicted from this work.  However, the $T_m$ from Ice-{\it i} is
387 > calculated to be 265 K, indicating that these simulation based
388 > structures ought to be included in studies probing phase transitions
389 > with this model.  Also of interest in these results is that SSD/E does
390   not exhibit a melting point at 1 atm, but it shows a sublimation point
391 < at 355 K. This is due to the significant stability of Ice-{\it i} over
392 < all other polymorphs for this particular model under these
393 < conditions. While troubling, this behavior turned out to be
394 < advantageous in that it facilitated the spontaneous crystallization of
395 < Ice-{\it i}. These observations provide a warning that simulations of
391 > at 355 K.  This is due to the significant stability of Ice-{\it i}
392 > over all other polymorphs for this particular model under these
393 > conditions.  While troubling, this behavior resulted in spontaneous
394 > crystallization of Ice-{\it i} and led us to investigate this
395 > structure.  These observations provide a warning that simulations of
396   SSD/E as a ``liquid'' near 300 K are actually metastable and run the
397 < risk of spontaneous crystallization. However, this risk changes when
397 > risk of spontaneous crystallization.  However, this risk lessens when
398   applying a longer cutoff.
399  
400   \begin{figure}
401   \includegraphics[width=\linewidth]{cutoffChange.eps}
402 < \caption{Free energy as a function of cutoff radius for (A) SSD/E, (B)
403 < TIP3P, and (C) SSD/RF. Data points omitted include SSD/E: $I_c$ 12
404 < \AA\, TIP3P: $I_c$ 12 \AA\ and B 12 \AA\, and SSD/RF: $I_c$ 9
405 < \AA . These crystals are unstable at 200 K and rapidly convert into
406 < liquids. The connecting lines are qualitative visual aid.}
402 > \caption{Free energy as a function of cutoff radius for SSD/E, TIP3P,
403 > SPC/E, SSD/RF with a reaction field, and the TIP3P and SPC/E models
404 > with an added Ewald correction term.  Calculations performed without a
405 > long-range correction show noticable free energy dependence on the
406 > cutoff radius and show some degree of converge at large cutoff radii.
407 > Inclusion of a long-range correction reduces the cutoff radius
408 > dependence of the free energy for all the models.  Error for the
409 > larger cutoff points is equivalent to that observed at 9.0 \AA\ (see
410 > Table \ref{freeEnergy}).  Data for ice I$_c$ with TIP3P using both 12
411 > and 13.5 \AA\ cutoffs were omitted because the crystal was prone to
412 > distortion and melting at 200 K.  Ice-{\it i}$^\prime$ is the form of
413 > Ice-{\it i} used in the SPC/E simulations.}
414   \label{incCutoff}
415   \end{figure}
416  
417   Increasing the cutoff radius in simulations of the more
418   computationally efficient water models was done in order to evaluate
419   the trend in free energy values when moving to systems that do not
420 < involve potential truncation. As seen in Fig. \ref{incCutoff}, the
421 < free energy of all the ice polymorphs show a substantial dependence on
422 < cutoff radius. In general, there is a narrowing of the free energy
423 < differences while moving to greater cutoff radius. Interestingly, by
424 < increasing the cutoff radius, the free energy gap was narrowed enough
425 < in the SSD/E model that the liquid state is preferred under standard
426 < simulation conditions (298 K and 1 atm). Thus, it is recommended that
420 > involve potential truncation.  As seen in Fig. \ref{incCutoff}, the
421 > free energy of the ice polymorphs with water models lacking a
422 > long-range correction show a significant cutoff radius dependence.  In
423 > general, there is a narrowing of the free energy differences while
424 > moving to greater cutoff radii.  As the free energies for the
425 > polymorphs converge, the stability advantage that Ice-{\it i} exhibits
426 > is reduced.  Interestingly, increasing the cutoff radius a mere 1.5
427 > \AA\ with the SSD/E model destabilizes the Ice-{\it i} polymorph
428 > enough that the liquid state is preferred under standard simulation
429 > conditions (298 K and 1 atm).  Thus, it is recommended that
430   simulations using this model choose interaction truncation radii
431 < greater than 9 \AA\ . This narrowing trend is much more subtle in the
432 < case of SSD/RF, indicating that the free energies calculated with a
433 < reaction field present provide a more accurate picture of the free
434 < energy landscape in the absence of potential truncation.
431 > greater than 9 \AA.  Considering the stabilization of Ice-{\it i} with
432 > smaller cutoffs, it is not surprising that crystallization was
433 > observed with SSD/E.  The choice of a 9 \AA\ cutoff in the previous
434 > simulations gives the Ice-{\it i} polymorph a greater than 1 kcal/mol
435 > lower free energy than the ice $I_\textrm{h}$ starting configurations.
436 > Additionally, it should be noted that ice $I_c$ is not stable with
437 > cutoff radii of 12 and 13.5 \AA\ using the TIP3P water model.  These
438 > simulations showed bulk distortions of the simulation cell that
439 > rapidly deteriorated crystal integrity.
440  
441 < To further study the changes resulting to the inclusion of a
442 < long-range interaction correction, the effect of an Ewald summation
443 < was estimated by applying the potential energy difference do to its
444 < inclusion in systems in the presence and absence of the
445 < correction. This was accomplished by calculation of the potential
446 < energy of identical crystals with and without PME using TINKER. The
447 < free energies for the investigated polymorphs using the TIP3P and
448 < SPC/E water models are shown in Table \ref{pmeShift}. The same trend
449 < pointed out through increase of cutoff radius is observed in these PME
450 < results. Ice-{\it i} is the preferred polymorph at ambient conditions
451 < for both the TIP3P and SPC/E water models; however, the narrowing of
452 < the free energy differences between the various solid forms is
453 < significant enough that it becomes less clear that it is the most
454 < stable polymorph with the SPC/E model.  The free energies of Ice-{\it
455 < i} and ice B nearly overlap within error, with ice $I_c$ just outside
456 < as well, indicating that Ice-{\it i} might be metastable with respect
457 < to ice B and possibly ice $I_c$ with SPC/E. However, these results do
458 < not significantly alter the finding that the Ice-{\it i} polymorph is
459 < a stable crystal structure that should be considered when studying the
460 < phase behavior of water models.
441 > Adjacent to each of these model plots is a system with an applied or
442 > estimated long-range correction.  SSD/RF was parametrized for use with
443 > a reaction field, and the benefit provided by this computationally
444 > inexpensive correction is apparent.  Due to the relative independence
445 > of the resultant free energies, calculations performed with a small
446 > cutoff radius provide resultant properties similar to what one would
447 > expect for the bulk material.  In the cases of TIP3P and SPC/E, the
448 > effect of an Ewald summation was estimated by applying the potential
449 > energy difference do to its inclusion in systems in the presence and
450 > absence of the correction.  This was accomplished by calculation of
451 > the potential energy of identical crystals both with and without
452 > particle mesh Ewald (PME).  Similar behavior to that observed with
453 > reaction field is seen for both of these models.  The free energies
454 > show less dependence on cutoff radius and span a more narrowed range
455 > for the various polymorphs.  Like the dipolar water models, TIP3P
456 > displays a relatively constant preference for the Ice-{\it i}
457 > polymorph.  Crystal preference is much more difficult to determine for
458 > SPC/E.  Without a long-range correction, each of the polymorphs
459 > studied assumes the role of the preferred polymorph under different
460 > cutoff conditions.  The inclusion of the Ewald correction flattens and
461 > narrows the sequences of free energies so much that they often overlap
462 > within error, indicating that other conditions, such as cell volume in
463 > microcanonical simulations, can influence the chosen polymorph upon
464 > crystallization.  All of these results support the finding that the
465 > Ice-{\it i} polymorph is a stable crystal structure that should be
466 > considered when studying the phase behavior of water models.
467  
442 \begin{table*}
443 \begin{minipage}{\linewidth}
444 \renewcommand{\thefootnote}{\thempfootnote}
445 \begin{center}
446 \caption{The free energy of the studied ice polymorphs after applying
447 the energy difference attributed to the inclusion of the PME
448 long-range interaction correction. Units are kcal/mol.}
449 \begin{tabular}{ l  c  c  c  c }
450 \hline
451 \ \ Water Model \ \ & \ \ \ \ \ $I_h$ \ \ & \ \ \ \ \ $I_c$ \ \ & \ \quad \ \ \ \ B \ \ & \ \ \ \ \ Ice-{\it i} \ \ \\
452 \hline
453 TIP3P  & -11.53(2) & -11.24(3) & -11.51(3) & -11.67(3)\\
454 SPC/E  & -12.77(2) & -12.92(2) & -12.96(3) & -13.02(2)\\
455 \end{tabular}
456 \label{pmeShift}
457 \end{center}
458 \end{minipage}
459 \end{table*}
460
468   \section{Conclusions}
469  
470   The free energy for proton ordered variants of hexagonal and cubic ice
471 < $I$, ice B, and recently discovered Ice-{\it i} were calculated under
472 < standard conditions for several common water models via thermodynamic
473 < integration. All the water models studied show Ice-{\it i} to be the
474 < minimum free energy crystal structure in the with a 9 \AA\ switching
475 < function cutoff. Calculated melting and boiling points show
476 < surprisingly good agreement with the experimental values; however, the
477 < solid phase at 1 atm is Ice-{\it i}, not ice $I_h$. The effect of
478 < interaction truncation was investigated through variation of the
479 < cutoff radius, use of a reaction field parameterized model, and
480 < estimation of the results in the presence of the Ewald
481 < summation. Interaction truncation has a significant effect on the
482 < computed free energy values, and may significantly alter the free
483 < energy landscape for the more complex multipoint water models. Despite
484 < these effects, these results show Ice-{\it i} to be an important ice
485 < polymorph that should be considered in simulation studies.
471 > $I$, ice B, and our recently discovered Ice-{\it i} structure were
472 > calculated under standard conditions for several common water models
473 > via thermodynamic integration.  All the water models studied show
474 > Ice-{\it i} to be the minimum free energy crystal structure with a 9
475 > \AA\ switching function cutoff.  Calculated melting and boiling points
476 > show surprisingly good agreement with the experimental values;
477 > however, the solid phase at 1 atm is Ice-{\it i}, not ice $I_h$.  The
478 > effect of interaction truncation was investigated through variation of
479 > the cutoff radius, use of a reaction field parameterized model, and
480 > estimation of the results in the presence of the Ewald summation.
481 > Interaction truncation has a significant effect on the computed free
482 > energy values, and may significantly alter the free energy landscape
483 > for the more complex multipoint water models.  Despite these effects,
484 > these results show Ice-{\it i} to be an important ice polymorph that
485 > should be considered in simulation studies.
486  
487 < Due to this relative stability of Ice-{\it i} in all manner of
488 < investigated simulation examples, the question arises as to possible
487 > Due to this relative stability of Ice-{\it i} in all of the
488 > investigated simulation conditions, the question arises as to possible
489   experimental observation of this polymorph.  The rather extensive past
490   and current experimental investigation of water in the low pressure
491   regime makes us hesitant to ascribe any relevance of this work outside
# Line 490 | Line 497 | and the structure factor ($S(\vec{q})$ for ice $I_c$ a
497   deposition environments, and in clathrate structures involving small
498   non-polar molecules.  Figs. \ref{fig:gofr} and \ref{fig:sofq} contain
499   our predictions for both the pair distribution function ($g_{OO}(r)$)
500 < and the structure factor ($S(\vec{q})$ for ice $I_c$ and for ice-{\it
501 < i} at a temperature of 77K.  In a quick comparison of the predicted
502 < S(q) for Ice-{\it i} and experimental studies of amorphous solid
503 < water, it is possible that some of the ``spurious'' peaks that could
504 < not be assigned in HDA could correspond to peaks labeled in this
505 < S(q).\cite{Bizid87} It should be noted that there is typically poor
506 < agreement on crystal densities between simulation and experiment, so
507 < such peak comparisons should be made with caution.  We will leave it
501 < to our experimental colleagues to determine whether this ice polymorph
502 < is named appropriately or if it should be promoted to Ice-0.
500 > and the structure factor ($S(\vec{q})$ for ice $I_h$, $I_c$, and for
501 > ice-{\it i} at a temperature of 77K.  In studies of the high and low
502 > density forms of amorphous ice, ``spurious'' diffraction peaks have
503 > been observed experimentally.\cite{Bizid87} It is possible that a
504 > variant of Ice-{\it i} could explain some of this behavior; however,
505 > we will leave it to our experimental colleagues to make the final
506 > determination on whether this ice polymorph is named appropriately
507 > (i.e. with an imaginary number) or if it can be promoted to Ice-0.
508  
509   \begin{figure}
510   \includegraphics[width=\linewidth]{iceGofr.eps}
511 < \caption{Radial distribution functions of Ice-{\it i} and ice $I_c$
512 < calculated from from simulations of the SSD/RF water model at 77 K.}
511 > \caption{Radial distribution functions of ice $I_h$, $I_c$,
512 > Ice-{\it i}, and Ice-{\it i}$^\prime$ calculated from from simulations
513 > of the SSD/RF water model at 77 K.}
514   \label{fig:gofr}
515   \end{figure}
516  
517   \begin{figure}
518   \includegraphics[width=\linewidth]{sofq.eps}
519 < \caption{Predicted structure factors for Ice-{\it i} and ice $I_c$ at
520 < 77 K.  The raw structure factors have been convoluted with a gaussian
521 < instrument function (0.075 \AA$^{-1}$ width) to compensate for the
522 < trunction effects in our finite size simulations. The labeled peaks
523 < compared favorably with ``spurious'' peaks observed in experimental
518 < studies of amorphous solid water.\cite{Bizid87}}
519 > \caption{Predicted structure factors for ice $I_h$, $I_c$, Ice-{\it i},
520 > and Ice-{\it i}$^\prime$ at 77 K.  The raw structure factors have
521 > been convoluted with a gaussian instrument function (0.075 \AA$^{-1}$
522 > width) to compensate for the trunction effects in our finite size
523 > simulations.}
524   \label{fig:sofq}
525   \end{figure}
526  

Diff Legend

Removed lines
+ Added lines
< Changed lines
> Changed lines