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

Comparing trunk/ssdePaper/nptSSD.tex (file contents):
Revision 1029 by gezelter, Thu Feb 5 20:47:50 2004 UTC vs.
Revision 1040 by chrisfen, Mon Feb 9 14:42:27 2004 UTC

# Line 1 | Line 1
1   %\documentclass[prb,aps,times,twocolumn,tabularx]{revtex4}
2 + %\documentclass[preprint,aps,endfloat]{revtex4}
3   \documentclass[11pt]{article}
4   \usepackage{endfloat}
5   \usepackage{amsmath}
# Line 8 | Line 9
9   \usepackage{tabularx}
10   \usepackage{graphicx}
11   \usepackage[ref]{overcite}
11 %\usepackage{berkeley}
12 %\usepackage{curves}
12   \pagestyle{plain}
13   \pagenumbering{arabic}
14   \oddsidemargin 0.0cm \evensidemargin 0.0cm
15   \topmargin -21pt \headsep 10pt
16   \textheight 9.0in \textwidth 6.5in
17   \brokenpenalty=10000
19 \renewcommand{\baselinestretch}{1.2}
20 \renewcommand\citemid{\ } % no comma in optional reference note
18  
19 + %\renewcommand\citemid{\ } % no comma in optional reference note
20 +
21   \begin{document}
22  
23   \title{On the structural and transport properties of the soft sticky
24 < dipole ({\sc ssd}) and related single point water models}
24 > dipole (SSD) and related single point water models}
25  
26 < \author{Christopher J. Fennell and J. Daniel Gezelter\footnote{Corresponding author. \ Electronic mail: gezelter@nd.edu} \\
27 < Department of Chemistry and Biochemistry\\ University of Notre Dame\\
26 > \author{Christopher J. Fennell and J. Daniel
27 > Gezelter\footnote{Corresponding author. \ Electronic mail:
28 > gezelter@nd.edu} \\ Department of Chemistry and Biochemistry\\ University of Notre Dame\\
29   Notre Dame, Indiana 46556}
30  
31   \date{\today}
32  
33   \maketitle
34 + \doublespacing
35  
36   \begin{abstract}
37   The density maximum and temperature dependence of the self-diffusion
38 < constant were investigated for the soft sticky dipole ({\sc ssd}) water
39 < model and two related re-parameterizations of this single-point model.
38 > constant were investigated for the soft sticky dipole (SSD) water
39 > model and two related reparameterizations of this single-point model.
40   A combination of microcanonical and isobaric-isothermal molecular
41   dynamics simulations were used to calculate these properties, both
42   with and without the use of reaction field to handle long-range
43   electrostatics.  The isobaric-isothermal (NPT) simulations of the
44   melting of both ice-$I_h$ and ice-$I_c$ showed a density maximum near
45 < 260 K.  In most cases, the use of the reaction field resulted in
45 > 260~K.  In most cases, the use of the reaction field resulted in
46   calculated densities which were were significantly lower than
47   experimental densities.  Analysis of self-diffusion constants shows
48 < that the original {\sc ssd} model captures the transport properties of
49 < experimental water very well in both the normal and super-cooled
50 < liquid regimes.  We also present our re-parameterized versions of {\sc ssd}
48 > that the original SSD model captures the transport properties of
49 > experimental water very well in both the normal and supercooled
50 > liquid regimes.  We also present our reparameterized versions of SSD
51   for use both with the reaction field or without any long-range
52 < electrostatic corrections.  These are called the {\sc ssd/rf} and {\sc ssd/e}
52 > electrostatic corrections.  These are called the SSD/RF and SSD/E
53   models respectively.  These modified models were shown to maintain or
54   improve upon the experimental agreement with the structural and
55 < transport properties that can be obtained with either the original {\sc ssd}
56 < or the density corrected version of the original model ({\sc ssd1}).
55 > transport properties that can be obtained with either the original SSD
56 > or the density corrected version of the original model (SSD1).
57   Additionally, a novel low-density ice structure is presented
58 < which appears to be the most stable ice structure for the entire {\sc ssd}
58 > which appears to be the most stable ice structure for the entire SSD
59   family.
60   \end{abstract}
61  
# Line 62 | Line 63 | family.
63  
64   %\narrowtext
65  
65
66   %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
67   %                              BODY OF TEXT
68   %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
# Line 89 | Line 89 | cost is the Soft Sticky Dipole ({\sc ssd}) water
89  
90   One recently developed model that largely succeeds in retaining the
91   accuracy of bulk properties while greatly reducing the computational
92 < cost is the Soft Sticky Dipole ({\sc ssd}) water
93 < model.\cite{Ichiye96,Ichiye96b,Ichiye99,Ichiye03} The {\sc ssd} model was
94 < developed by Ichiye \emph{et al.} as a modified form of the
92 > cost is the Soft Sticky Dipole (SSD) water
93 > model.\cite{Ichiye96,Ichiye96b,Ichiye99,Ichiye03} The SSD model
94 > was developed by Ichiye \emph{et al.} as a modified form of the
95   hard-sphere water model proposed by Bratko, Blum, and
96 < Luzar.\cite{Bratko85,Bratko95} {\sc ssd} is a {\it single point} model which
97 < has an interaction site that is both a point dipole along with a
96 > Luzar.\cite{Bratko85,Bratko95} SSD is a {\it single point} model
97 > which has an interaction site that is both a point dipole and a
98   Lennard-Jones core.  However, since the normal aligned and
99   anti-aligned geometries favored by point dipoles are poor mimics of
100   local structure in liquid water, a short ranged ``sticky'' potential
101   is also added.  The sticky potential directs the molecules to assume
102 < the proper hydrogen bond orientation in the first solvation
103 < shell.  
102 > the proper hydrogen bond orientation in the first solvation shell.
103  
104 < The interaction between two {\sc ssd} water molecules \emph{i} and \emph{j}
104 > The interaction between two SSD water molecules \emph{i} and \emph{j}
105   is given by the potential
106   \begin{equation}
107   u_{ij} = u_{ij}^{LJ} (r_{ij})\ + u_{ij}^{dp}
# Line 159 | Line 158 | can be found in the original {\sc ssd}
158   enhances the tetrahedral geometry for hydrogen bonded structures),
159   while $w^\prime$ is a purely empirical function.  A more detailed
160   description of the functional parts and variables in this potential
161 < can be found in the original {\sc ssd}
161 > can be found in the original SSD
162   articles.\cite{Ichiye96,Ichiye96b,Ichiye99,Ichiye03}
163  
164 < Since {\sc ssd} is a single-point {\it dipolar} model, the force
164 > Since SSD is a single-point {\it dipolar} model, the force
165   calculations are simplified significantly relative to the standard
166   {\it charged} multi-point models. In the original Monte Carlo
167 < simulations using this model, Ichiye {\it et al.} reported that using
168 < {\sc ssd} decreased computer time by a factor of 6-7 compared to other
167 > simulations using this model, Liu and Ichiye reported that using SSD
168 > decreased computer time by a factor of 6-7 compared to other
169   models.\cite{Ichiye96} What is most impressive is that this savings
170   did not come at the expense of accurate depiction of the liquid state
171 < properties.  Indeed, {\sc ssd} maintains reasonable agreement with the Soper
171 > properties.  Indeed, SSD maintains reasonable agreement with the Soper
172   data for the structural features of liquid
173   water.\cite{Soper86,Ichiye96} Additionally, the dynamical properties
174 < exhibited by {\sc ssd} agree with experiment better than those of more
174 > exhibited by SSD agree with experiment better than those of more
175   computationally expensive models (like TIP3P and
176   SPC/E).\cite{Ichiye99} The combination of speed and accurate depiction
177 < of solvent properties makes {\sc ssd} a very attractive model for the
177 > of solvent properties makes SSD a very attractive model for the
178   simulation of large scale biochemical simulations.
179  
180 < One feature of the {\sc ssd} model is that it was parameterized for use with
181 < the Ewald sum to handle long-range interactions.  This would normally
182 < be the best way of handling long-range interactions in systems that
183 < contain other point charges.  However, our group has recently become
184 < interested in systems with point dipoles as mimics for neutral, but
185 < polarized regions on molecules (e.g. the zwitterionic head group
186 < regions of phospholipids).  If the system of interest does not contain
187 < point charges, the Ewald sum and even particle-mesh Ewald become
188 < computational bottlenecks.  Their respective ideal $N^\frac{3}{2}$ and
189 < $N\log N$ calculation scaling orders for $N$ particles can become
190 < prohibitive when $N$ becomes large.\cite{Darden99} In applying this
191 < water model in these types of systems, it would be useful to know its
192 < properties and behavior under the more computationally efficient
193 < reaction field (RF) technique, or even with a simple cutoff. This
194 < study addresses these issues by looking at the structural and
195 < transport behavior of {\sc ssd} over a variety of temperatures with the
196 < purpose of utilizing the RF correction technique.  We then suggest
197 < modifications to the parameters that result in more realistic bulk
198 < phase behavior.  It should be noted that in a recent publication, some
199 < of the original investigators of the {\sc ssd} water model have suggested
200 < adjustments to the {\sc ssd} water model to address abnormal density
201 < behavior (also observed here), calling the corrected model
202 < {\sc ssd1}.\cite{Ichiye03} In what follows, we compare our
203 < reparamaterization of {\sc ssd} with both the original {\sc ssd} and {\sc ssd1} models
204 < with the goal of improving the bulk phase behavior of an {\sc ssd}-derived
205 < model in simulations utilizing the Reaction Field.
180 > One feature of the SSD model is that it was parameterized for
181 > use with the Ewald sum to handle long-range interactions.  This would
182 > normally be the best way of handling long-range interactions in
183 > systems that contain other point charges.  However, our group has
184 > recently become interested in systems with point dipoles as mimics for
185 > neutral, but polarized regions on molecules (e.g. the zwitterionic
186 > head group regions of phospholipids).  If the system of interest does
187 > not contain point charges, the Ewald sum and even particle-mesh Ewald
188 > become computational bottlenecks.  Their respective ideal
189 > $N^\frac{3}{2}$ and $N\log N$ calculation scaling orders for $N$
190 > particles can become prohibitive when $N$ becomes
191 > large.\cite{Darden99} In applying this water model in these types of
192 > systems, it would be useful to know its properties and behavior under
193 > the more computationally efficient reaction field (RF) technique, or
194 > even with a simple cutoff. This study addresses these issues by
195 > looking at the structural and transport behavior of SSD over a
196 > variety of temperatures with the purpose of utilizing the RF
197 > correction technique.  We then suggest modifications to the parameters
198 > that result in more realistic bulk phase behavior.  It should be noted
199 > that in a recent publication, some of the original investigators of
200 > the SSD water model have suggested adjustments to the SSD
201 > water model to address abnormal density behavior (also observed here),
202 > calling the corrected model SSD1.\cite{Ichiye03} In what
203 > follows, we compare our reparamaterization of SSD with both the
204 > original SSD and SSD1 models with the goal of improving
205 > the bulk phase behavior of an SSD-derived model in simulations
206 > utilizing the reaction field.
207  
208   \section{Methods}
209  
210   Long-range dipole-dipole interactions were accounted for in this study
211 < by using either the reaction field method or by resorting to a simple
212 < cubic switching function at a cutoff radius.  The reaction field
213 < method was actually first used in Monte Carlo simulations of liquid
214 < water.\cite{Barker73} Under this method, the magnitude of the reaction
215 < field acting on dipole $i$ is
211 > by using either the reaction field technique or by resorting to a
212 > simple cubic switching function at a cutoff radius.  One of the early
213 > applications of a reaction field was actually in Monte Carlo
214 > simulations of liquid water.\cite{Barker73} Under this method, the
215 > magnitude of the reaction field acting on dipole $i$ is
216   \begin{equation}
217   \mathcal{E}_{i} = \frac{2(\varepsilon_{s} - 1)}{2\varepsilon_{s} + 1}
218   \frac{1}{r_{c}^{3}} \sum_{j\in{\mathcal{R}}} {\bf \mu}_{j} s(r_{ij}),
# Line 226 | Line 226 | field is known to alter the bulk orientational propert
226   total energy by particle $i$ is given by $-\frac{1}{2}{\bf
227   \mu}_{i}\cdot\mathcal{E}_{i}$ and the torque on dipole $i$ by ${\bf
228   \mu}_{i}\times\mathcal{E}_{i}$.\cite{AllenTildesley}  Use of the reaction
229 < field is known to alter the bulk orientational properties, such as the
230 < dielectric relaxation time.  There is particular sensitivity of this
231 < property on changes in the length of the cutoff
232 < radius.\cite{Berendsen98} This variable behavior makes reaction field
233 < a less attractive method than the Ewald sum.  However, for very large
234 < systems, the computational benefit of reaction field is dramatic.
229 > field is known to alter the bulk orientational properties of simulated
230 > water, and there is particular sensitivity of these properties on
231 > changes in the length of the cutoff radius.\cite{Berendsen98} This
232 > variable behavior makes reaction field a less attractive method than
233 > the Ewald sum.  However, for very large systems, the computational
234 > benefit of reaction field is dramatic.
235  
236   We have also performed a companion set of simulations {\it without} a
237   surrounding dielectric (i.e. using a simple cubic switching function
238   at the cutoff radius), and as a result we have two reparamaterizations
239 < of {\sc ssd} which could be used either with or without the reaction field
240 < turned on.
239 > of SSD which could be used either with or without the reaction
240 > field turned on.
241  
242   Simulations to obtain the preferred densities of the models were
243   performed in the isobaric-isothermal (NPT) ensemble, while all
# Line 247 | Line 247 | simulations of {\sc ssd}, because there are no explici
247   implemented using an integral thermostat and barostat as outlined by
248   Hoover.\cite{Hoover85,Hoover86} All molecules were treated as
249   non-linear rigid bodies. Vibrational constraints are not necessary in
250 < simulations of {\sc ssd}, because there are no explicit hydrogen atoms, and
251 < thus no molecular vibrational modes need to be considered.
250 > simulations of SSD, because there are no explicit hydrogen
251 > atoms, and thus no molecular vibrational modes need to be considered.
252  
253   Integration of the equations of motion was carried out using the
254   symplectic splitting method proposed by Dullweber, Leimkuhler, and
255 < McLachlan ({\sc dlm}).\cite{Dullweber1997} Our reason for selecting this
256 < integrator centers on poor energy conservation of rigid body dynamics
257 < using traditional quaternion integration.\cite{Evans77,Evans77b} In
258 < typical microcanonical ensemble simulations, the energy drift when
259 < using quaternions was substantially greater than when using the {\sc dlm}
260 < method (fig. \ref{timestep}).  This steady drift in the total energy
261 < has also been observed by Kol {\it et al.}\cite{Laird97}
255 > McLachlan ({\sc dlm}).\cite{Dullweber1997} Our reason for selecting
256 > this integrator centers on poor energy conservation of rigid body
257 > dynamics using traditional quaternion
258 > integration.\cite{Evans77,Evans77b} In typical microcanonical ensemble
259 > simulations, the energy drift when using quaternions was substantially
260 > greater than when using the {\sc dlm} method (fig. \ref{timestep}).
261 > This steady drift in the total energy has also been observed by Kol
262 > {\it et al.}\cite{Laird97}
263  
264   The key difference in the integration method proposed by Dullweber
265   \emph{et al.} is that the entire rotation matrix is propagated from
# Line 273 | Line 274 | step, a 1000 {\sc ssd} particle simulation shows an av
274   of matrix evaluations to update the rotation
275   matrix.\cite{Dullweber1997} These matrix rotations are more costly
276   than the simpler arithmetic quaternion propagation. With the same time
277 < step, a 1000 {\sc ssd} particle simulation shows an average 7\% increase in
278 < computation time using the {\sc dlm} method in place of quaternions. The
279 < additional expense per step is justified when one considers the
280 < ability to use time steps that are nearly twice as large under {\sc dlm}
281 < than would be usable under quaternion dynamics.  The energy
282 < conservation of the two methods using a number of different time steps
283 < is illustrated in figure
277 > step, a 1000 SSD particle simulation shows an average 7\%
278 > increase in computation time using the {\sc dlm} method in place of
279 > quaternions. The additional expense per step is justified when one
280 > considers the ability to use time steps that are nearly twice as large
281 > under {\sc dlm} than would be usable under quaternion dynamics.  The
282 > energy conservation of the two methods using a number of different
283 > time steps is illustrated in figure
284   \ref{timestep}.
285  
286   \begin{figure}
287   \begin{center}
288   \epsfxsize=6in
289   \epsfbox{timeStep.epsi}
290 < \caption{Energy conservation using both quaternion-based integration and
291 < the {\sc dlm} method with increasing time step. The larger time step plots
292 < are shifted from the true energy baseline (that of $\Delta t$ = 0.1
293 < fs) for clarity.}
290 > \caption{Energy conservation using both quaternion-based integration and the
291 > {\sc dlm} method with increasing time step. The larger time step plots
292 > are shifted from the true energy baseline (that of $\Delta t$ =
293 > 0.1~fs) for clarity.}
294   \label{timestep}
295   \end{center}
296   \end{figure}
297  
298   In figure \ref{timestep}, the resulting energy drift at various time
299 < steps for both the {\sc dlm} and quaternion integration schemes is compared.
300 < All of the 1000 {\sc ssd} particle simulations started with the same
301 < configuration, and the only difference was the method used to handle
302 < orientational motion. At time steps of 0.1 and 0.5 fs, both methods
303 < for propagating the orientational degrees of freedom conserve energy
304 < fairly well, with the quaternion method showing a slight energy drift
305 < over time in the 0.5 fs time step simulation. At time steps of 1 and 2
306 < fs, the energy conservation benefits of the {\sc dlm} method are clearly
307 < demonstrated. Thus, while maintaining the same degree of energy
308 < conservation, one can take considerably longer time steps, leading to
309 < an overall reduction in computation time.
299 > steps for both the {\sc dlm} and quaternion integration schemes is
300 > compared.  All of the 1000 SSD particle simulations started with
301 > the same configuration, and the only difference was the method used to
302 > handle orientational motion. At time steps of 0.1 and 0.5~fs, both
303 > methods for propagating the orientational degrees of freedom conserve
304 > energy fairly well, with the quaternion method showing a slight energy
305 > drift over time in the 0.5~fs time step simulation. At time steps of 1
306 > and 2~fs, the energy conservation benefits of the {\sc dlm} method are
307 > clearly demonstrated. Thus, while maintaining the same degree of
308 > energy conservation, one can take considerably longer time steps,
309 > leading to an overall reduction in computation time.
310  
311 < Energy drift in the simulations using {\sc dlm} integration was unnoticeable
312 < for time steps up to 3 fs. A slight energy drift on the order of 0.012
313 < kcal/mol per nanosecond was observed at a time step of 4 fs, and as
314 < expected, this drift increases dramatically with increasing time
315 < step. To insure accuracy in our microcanonical simulations, time steps
316 < were set at 2 fs and kept at this value for constant pressure
317 < simulations as well.
311 > Energy drift in the simulations using {\sc dlm} integration was
312 > unnoticeable for time steps up to 3~fs. A slight energy drift on the
313 > order of 0.012~kcal/mol per nanosecond was observed at a time step of
314 > 4~fs, and as expected, this drift increases dramatically with
315 > increasing time step. To insure accuracy in our microcanonical
316 > simulations, time steps were set at 2~fs and kept at this value for
317 > constant pressure simulations as well.
318  
319   Proton-disordered ice crystals in both the $I_h$ and $I_c$ lattices
320   were generated as starting points for all simulations. The $I_h$
321 < crystals were formed by first arranging the centers of mass of the {\sc ssd}
321 > crystals were formed by first arranging the centers of mass of the SSD
322   particles into a ``hexagonal'' ice lattice of 1024 particles. Because
323   of the crystal structure of $I_h$ ice, the simulation box assumed an
324   orthorhombic shape with an edge length ratio of approximately
325   1.00$\times$1.06$\times$1.23. The particles were then allowed to
326   orient freely about fixed positions with angular momenta randomized at
327 < 400 K for varying times. The rotational temperature was then scaled
328 < down in stages to slowly cool the crystals to 25 K. The particles were
327 > 400~K for varying times. The rotational temperature was then scaled
328 > down in stages to slowly cool the crystals to 25~K. The particles were
329   then allowed to translate with fixed orientations at a constant
330 < pressure of 1 atm for 50 ps at 25 K. Finally, all constraints were
331 < removed and the ice crystals were allowed to equilibrate for 50 ps at
332 < 25 K and a constant pressure of 1 atm.  This procedure resulted in
330 > pressure of 1 atm for 50~ps at 25~K. Finally, all constraints were
331 > removed and the ice crystals were allowed to equilibrate for 50~ps at
332 > 25~K and a constant pressure of 1~atm.  This procedure resulted in
333   structurally stable $I_h$ ice crystals that obey the Bernal-Fowler
334   rules.\cite{Bernal33,Rahman72} This method was also utilized in the
335   making of diamond lattice $I_c$ ice crystals, with each cubic
# Line 346 | Line 347 | for 100 ps prior to a 200 ps data collection run at ea
347   supercooled regime. An ensemble average from five separate melting
348   simulations was acquired, each starting from different ice crystals
349   generated as described previously. All simulations were equilibrated
350 < for 100 ps prior to a 200 ps data collection run at each temperature
351 < setting. The temperature range of study spanned from 25 to 400 K, with
352 < a maximum degree increment of 25 K. For regions of interest along this
353 < stepwise progression, the temperature increment was decreased from 25
354 < K to 10 and 5 K.  The above equilibration and production times were
350 > for 100~ps prior to a 200~ps data collection run at each temperature
351 > setting. The temperature range of study spanned from 25 to 400~K, with
352 > a maximum degree increment of 25~K. For regions of interest along this
353 > stepwise progression, the temperature increment was decreased from
354 > 25~K to 10 and 5~K.  The above equilibration and production times were
355   sufficient in that fluctuations in the volume autocorrelation function
356 < were damped out in all simulations in under 20 ps.
356 > were damped out in all simulations in under 20~ps.
357  
358   \subsection{Density Behavior}
359  
360 < Our initial simulations focused on the original {\sc ssd} water model, and
361 < an average density versus temperature plot is shown in figure
360 > Our initial simulations focused on the original SSD water model,
361 > and an average density versus temperature plot is shown in figure
362   \ref{dense1}. Note that the density maximum when using a reaction
363 < field appears between 255 and 265 K.  There were smaller fluctuations
364 < in the density at 260 K than at either 255 or 265, so we report this
363 > field appears between 255 and 265~K.  There were smaller fluctuations
364 > in the density at 260~K than at either 255 or 265~K, so we report this
365   value as the location of the density maximum. Figure \ref{dense1} was
366   constructed using ice $I_h$ crystals for the initial configuration;
367   though not pictured, the simulations starting from ice $I_c$ crystal
368   configurations showed similar results, with a liquid-phase density
369 < maximum in this same region (between 255 and 260 K).
369 > maximum in this same region (between 255 and 260~K).
370  
371   \begin{figure}
372   \begin{center}
373   \epsfxsize=6in
374 < \epsfbox{denseSSD.eps}
375 < \caption{Density versus temperature for TIP4P [Ref. \citen{Jorgensen98b}],
376 < TIP3P [Ref. \citen{Jorgensen98b}], SPC/E [Ref. \citen{Clancy94}], {\sc ssd}
377 < without Reaction Field, {\sc ssd}, and experiment [Ref. \citen{CRC80}]. The
374 > \epsfbox{denseSSDnew.eps}
375 > \caption{ Density versus temperature for TIP4P [Ref. \citen{Jorgensen98b}],
376 > TIP3P [Ref. \citen{Jorgensen98b}], SPC/E [Ref. \citen{Clancy94}], SSD
377 > without Reaction Field, SSD, and experiment [Ref. \citen{CRC80}]. The
378   arrows indicate the change in densities observed when turning off the
379 < reaction field. The the lower than expected densities for the {\sc ssd}
380 < model were what prompted the original reparameterization of {\sc ssd1}
379 > reaction field. The the lower than expected densities for the SSD
380 > model were what prompted the original reparameterization of SSD1
381   [Ref. \citen{Ichiye03}].}
382   \label{dense1}
383   \end{center}
384   \end{figure}
385  
386 < The density maximum for {\sc ssd} compares quite favorably to other simple
387 < water models. Figure \ref{dense1} also shows calculated densities of
388 < several other models and experiment obtained from other
386 > The density maximum for SSD compares quite favorably to other
387 > simple water models. Figure \ref{dense1} also shows calculated
388 > densities of several other models and experiment obtained from other
389   sources.\cite{Jorgensen98b,Clancy94,CRC80} Of the listed simple water
390 < models, {\sc ssd} has a temperature closest to the experimentally observed
391 < density maximum. Of the {\it charge-based} models in
390 > models, SSD has a temperature closest to the experimentally
391 > observed density maximum. Of the {\it charge-based} models in
392   Fig. \ref{dense1}, TIP4P has a density maximum behavior most like that
393 < seen in {\sc ssd}. Though not included in this plot, it is useful
394 < to note that TIP5P has a density maximum nearly identical to the
393 > seen in SSD. Though not included in this plot, it is useful to
394 > note that TIP5P has a density maximum nearly identical to the
395   experimentally measured temperature.
396  
397   It has been observed that liquid state densities in water are
398   dependent on the cutoff radius used both with and without the use of
399   reaction field.\cite{Berendsen98} In order to address the possible
400   effect of cutoff radius, simulations were performed with a dipolar
401 < cutoff radius of 12.0 \AA\ to complement the previous {\sc ssd} simulations,
402 < all performed with a cutoff of 9.0 \AA. All of the resulting densities
403 < overlapped within error and showed no significant trend toward lower
404 < or higher densities as a function of cutoff radius, for simulations
405 < both with and without reaction field. These results indicate that
406 < there is no major benefit in choosing a longer cutoff radius in
407 < simulations using {\sc ssd}. This is advantageous in that the use of a
408 < longer cutoff radius results in a significant increase in the time
409 < required to obtain a single trajectory.
401 > cutoff radius of 12.0~\AA\ to complement the previous SSD
402 > simulations, all performed with a cutoff of 9.0~\AA. All of the
403 > resulting densities overlapped within error and showed no significant
404 > trend toward lower or higher densities as a function of cutoff radius,
405 > for simulations both with and without reaction field. These results
406 > indicate that there is no major benefit in choosing a longer cutoff
407 > radius in simulations using SSD. This is advantageous in that
408 > the use of a longer cutoff radius results in a significant increase in
409 > the time required to obtain a single trajectory.
410  
411   The key feature to recognize in figure \ref{dense1} is the density
412 < scaling of {\sc ssd} relative to other common models at any given
413 < temperature. {\sc ssd} assumes a lower density than any of the other listed
414 < models at the same pressure, behavior which is especially apparent at
415 < temperatures greater than 300 K. Lower than expected densities have
416 < been observed for other systems using a reaction field for long-range
417 < electrostatic interactions, so the most likely reason for the
418 < significantly lower densities seen in these simulations is the
412 > scaling of SSD relative to other common models at any given
413 > temperature. SSD assumes a lower density than any of the other
414 > listed models at the same pressure, behavior which is especially
415 > apparent at temperatures greater than 300~K. Lower than expected
416 > densities have been observed for other systems using a reaction field
417 > for long-range electrostatic interactions, so the most likely reason
418 > for the significantly lower densities seen in these simulations is the
419   presence of the reaction field.\cite{Berendsen98,Nezbeda02} In order
420   to test the effect of the reaction field on the density of the
421   systems, the simulations were repeated without a reaction field
# Line 422 | Line 423 | the curve produced from {\sc ssd} simulations using re
423   \ref{dense1}. Without the reaction field, the densities increase
424   to more experimentally reasonable values, especially around the
425   freezing point of liquid water. The shape of the curve is similar to
426 < the curve produced from {\sc ssd} simulations using reaction field,
426 > the curve produced from SSD simulations using reaction field,
427   specifically the rapidly decreasing densities at higher temperatures;
428 < however, a shift in the density maximum location, down to 245 K, is
428 > however, a shift in the density maximum location, down to 245~K, is
429   observed. This is a more accurate comparison to the other listed water
430   models, in that no long range corrections were applied in those
431   simulations.\cite{Clancy94,Jorgensen98b} However, even without the
432 < reaction field, the density around 300 K is still significantly lower
432 > reaction field, the density around 300~K is still significantly lower
433   than experiment and comparable water models. This anomalous behavior
434   was what lead Tan {\it et al.} to recently reparameterize
435 < {\sc ssd}.\cite{Ichiye03} Throughout the remainder of the paper our
436 < reparamaterizations of {\sc ssd} will be compared with their newer {\sc ssd1}
435 > SSD.\cite{Ichiye03} Throughout the remainder of the paper our
436 > reparamaterizations of SSD will be compared with their newer SSD1
437   model.
438  
439   \subsection{Transport Behavior}
# Line 443 | Line 444 | underwent 50 ps of temperature scaling and 50 ps of co
444   constant energy (NVE) simulations were performed at the average
445   density obtained by the NPT simulations at an identical target
446   temperature. Simulations started with randomized velocities and
447 < underwent 50 ps of temperature scaling and 50 ps of constant energy
448 < equilibration before a 200 ps data collection run. Diffusion constants
447 > underwent 50~ps of temperature scaling and 50~ps of constant energy
448 > equilibration before a 200~ps data collection run. Diffusion constants
449   were calculated via linear fits to the long-time behavior of the
450   mean-square displacement as a function of time. The averaged results
451   from five sets of NVE simulations are displayed in figure
# Line 455 | Line 456 | results.\cite{Gillen72,Holz00,Clancy94,Jorgensen01}
456   \begin{center}
457   \epsfxsize=6in
458   \epsfbox{betterDiffuse.epsi}
459 < \caption{Average self-diffusion constant as a function of temperature for
460 < {\sc ssd}, SPC/E [Ref. \citen{Clancy94}], and TIP5P
459 > \caption{ Average self-diffusion constant as a function of temperature for
460 > SSD, SPC/E [Ref. \citen{Clancy94}], and TIP5P
461   [Ref. \citen{Jorgensen01}] compared with experimental data
462   [Refs. \citen{Gillen72} and \citen{Holz00}]. Of the three water models
463 < shown, {\sc ssd} has the least deviation from the experimental values. The
464 < rapidly increasing diffusion constants for TIP5P and {\sc ssd} correspond to
463 > shown, SSD has the least deviation from the experimental values. The
464 > rapidly increasing diffusion constants for TIP5P and SSD correspond to
465   significant decreases in density at the higher temperatures.}
466   \label{diffuse}
467   \end{center}
468   \end{figure}
469  
470   The observed values for the diffusion constant point out one of the
471 < strengths of the {\sc ssd} model. Of the three models shown, the {\sc ssd} model
471 > strengths of the SSD model. Of the three models shown, the SSD model
472   has the most accurate depiction of self-diffusion in both the
473   supercooled and liquid regimes.  SPC/E does a respectable job by
474 < reproducing values similar to experiment around 290 K; however, it
474 > reproducing values similar to experiment around 290~K; however, it
475   deviates at both higher and lower temperatures, failing to predict the
476 < correct thermal trend. TIP5P and {\sc ssd} both start off low at colder
476 > correct thermal trend. TIP5P and SSD both start off low at colder
477   temperatures and tend to diffuse too rapidly at higher temperatures.
478   This behavior at higher temperatures is not particularly surprising
479 < since the densities of both TIP5P and {\sc ssd} are lower than experimental
479 > since the densities of both TIP5P and SSD are lower than experimental
480   water densities at higher temperatures.  When calculating the
481 < diffusion coefficients for {\sc ssd} at experimental densities (instead of
482 < the densities from the NPT simulations), the resulting values fall
483 < more in line with experiment at these temperatures.
481 > diffusion coefficients for SSD at experimental densities
482 > (instead of the densities from the NPT simulations), the resulting
483 > values fall more in line with experiment at these temperatures.
484  
485   \subsection{Structural Changes and Characterization}
486  
# Line 489 | Line 490 | at 245 K, indicating a first order phase transition fo
490   capacity (C$_\text{p}$) was monitored to locate the melting transition
491   in each of the simulations. In the melting simulations of the 1024
492   particle ice $I_h$ simulations, a large spike in C$_\text{p}$ occurs
493 < at 245 K, indicating a first order phase transition for the melting of
493 > at 245~K, indicating a first order phase transition for the melting of
494   these ice crystals. When the reaction field is turned off, the melting
495 < transition occurs at 235 K.  These melting transitions are
495 > transition occurs at 235~K.  These melting transitions are
496   considerably lower than the experimental value.
497  
498   \begin{figure}
498 \begin{center}
499 \epsfxsize=6in
500 \epsfbox{corrDiag.eps}
501 \caption{An illustration of angles involved in the correlations observed in Fig. \ref{contour}.}
502 \label{corrAngle}
503 \end{center}
504 \end{figure}
505
506 \begin{figure}
499   \begin{center}
500   \epsfxsize=6in
501   \epsfbox{fullContours.eps}
502 < \caption{Contour plots of 2D angular pair correlation functions for
503 < 512 {\sc ssd} molecules at 100 K (A \& B) and 300 K (C \& D). Dark areas
502 > \caption{ Contour plots of 2D angular pair correlation functions for
503 > 512 SSD molecules at 100~K (A \& B) and 300~K (C \& D). Dark areas
504   signify regions of enhanced density while light areas signify
505   depletion relative to the bulk density. White areas have pair
506   correlation values below 0.5 and black areas have values above 1.5.}
# Line 516 | Line 508 | Additional analysis of the melting process was perform
508   \end{center}
509   \end{figure}
510  
511 + \begin{figure}
512 + \begin{center}
513 + \epsfxsize=6in
514 + \epsfbox{corrDiag.eps}
515 + \caption{ An illustration of angles involved in the correlations observed in Fig. \ref{contour}.}
516 + \label{corrAngle}
517 + \end{center}
518 + \end{figure}
519 +
520   Additional analysis of the melting process was performed using
521   two-dimensional structure and dipole angle correlations. Expressions
522   for these correlations are as follows:
# Line 554 | Line 555 | this split character alters to show the leading 4 \AA\
555   $g_\mathrm{OO}(r)$.\cite{Ichiye96} At low temperatures, the second
556   solvation shell peak appears to have two distinct components that
557   blend together to form one observable peak. At higher temperatures,
558 < this split character alters to show the leading 4 \AA\ peak dominated
558 > this split character alters to show the leading 4~\AA\ peak dominated
559   by equatorial anti-parallel dipole orientations. There is also a
560   tightly bunched group of axially arranged dipoles that most likely
561   consist of the smaller fraction of aligned dipole pairs. The trailing
562 < component of the split peak at 5 \AA\ is dominated by aligned dipoles
562 > component of the split peak at 5~\AA\ is dominated by aligned dipoles
563   that assume hydrogen bond arrangements similar to those seen in the
564   first solvation shell. This evidence indicates that the dipole pair
565   interaction begins to dominate outside of the range of the dipolar
566   repulsion term.  The energetically favorable dipole arrangements
567   populate the region immediately outside this repulsion region (around
568 < 4 \AA), while arrangements that seek to satisfy both the sticky and
568 > 4~\AA), while arrangements that seek to satisfy both the sticky and
569   dipole forces locate themselves just beyond this initial buildup
570 < (around 5 \AA).
570 > (around 5~\AA).
571  
572   From these findings, the split second peak is primarily the product of
573   the dipolar repulsion term of the sticky potential. In fact, the inner
574   peak can be pushed out and merged with the outer split peak just by
575   extending the switching function ($s^\prime(r_{ij})$) from its normal
576 < 4.0 \AA\ cutoff to values of 4.5 or even 5 \AA. This type of
576 > 4.0~\AA\ cutoff to values of 4.5 or even 5~\AA. This type of
577   correction is not recommended for improving the liquid structure,
578   since the second solvation shell would still be shifted too far
579   out. In addition, this would have an even more detrimental effect on
580   the system densities, leading to a liquid with a more open structure
581 < and a density considerably lower than the already low {\sc ssd} density.  A
582 < better correction would be to include the quadrupole-quadrupole
583 < interactions for the water particles outside of the first solvation
584 < shell, but this would remove the simplicity and speed advantage of
585 < {\sc ssd}.
581 > and a density considerably lower than the already low SSD
582 > density.  A better correction would be to include the
583 > quadrupole-quadrupole interactions for the water particles outside of
584 > the first solvation shell, but this would remove the simplicity and
585 > speed advantage of SSD.
586  
587 < \subsection{Adjusted Potentials: {\sc ssd/rf} and {\sc ssd/e}}
587 > \subsection{Adjusted Potentials: SSD/RF and SSD/E}
588  
589 < The propensity of {\sc ssd} to adopt lower than expected densities under
589 > The propensity of SSD to adopt lower than expected densities under
590   varying conditions is troubling, especially at higher temperatures. In
591   order to correct this model for use with a reaction field, it is
592   necessary to adjust the force field parameters for the primary
# Line 601 | Line 602 | Sticky Dipole / Reaction Field ({\sc ssd/rf} - for use
602   function cutoffs ($r_l$, $r_u$ and $r_l^\prime$, $r_u^\prime$
603   respectively). The results of the reparameterizations are shown in
604   table \ref{params}. We are calling these reparameterizations the Soft
605 < Sticky Dipole / Reaction Field ({\sc ssd/rf} - for use with a reaction
606 < field) and Soft Sticky Dipole Extended ({\sc ssd/e} - an attempt to improve
605 > Sticky Dipole / Reaction Field (SSD/RF - for use with a reaction
606 > field) and Soft Sticky Dipole Extended (SSD/E - an attempt to improve
607   the liquid structure in simulations without a long-range correction).
608  
609   \begin{table}
610   \begin{center}
611 < \caption{Parameters for the original and adjusted models}
611 > \caption{ Parameters for the original and adjusted models}
612   \begin{tabular}{ l  c  c  c  c }
613   \hline \\[-3mm]
614 < \ \ \ Parameters\ \ \  & \ \ \ {\sc ssd} [Ref. \citen{Ichiye96}] \ \ \
615 < & \ {\sc ssd1} [Ref. \citen{Ichiye03}]\ \  & \ {\sc ssd/e}\ \  & \ {\sc ssd/rf} \\
614 > \ \ \ Parameters\ \ \  & \ \ \ SSD [Ref. \citen{Ichiye96}] \ \ \
615 > & \ SSD1 [Ref. \citen{Ichiye03}]\ \  & \ SSD/E\ \  & \ \ SSD/RF \\
616   \hline \\[-3mm]
617   \ \ \ $\sigma$ (\AA)  & 3.051 & 3.016 & 3.035 & 3.019\\
618   \ \ \ $\epsilon$ (kcal/mol) & 0.152 & 0.152 & 0.152 & 0.152\\
# Line 631 | Line 632 | the liquid structure in simulations without a long-ran
632   \begin{center}
633   \epsfxsize=5in
634   \epsfbox{GofRCompare.epsi}
635 < \caption{Plots comparing experiment [Ref. \citen{Head-Gordon00_1}] with {\sc ssd/e}
636 < and {\sc ssd1} without reaction field (top), as well as {\sc ssd/rf} and {\sc ssd1} with
637 < reaction field turned on (bottom). The insets show the respective
638 < first peaks in detail. Note how the changes in parameters have lowered
639 < and broadened the first peak of {\sc ssd/e} and {\sc ssd/rf}.}
635 > \caption{ Plots comparing experiment [Ref. \citen{Head-Gordon00_1}] with
636 > SSD/E and SSD1 without reaction field (top), as well as
637 > SSD/RF and SSD1 with reaction field turned on
638 > (bottom). The insets show the respective first peaks in detail. Note
639 > how the changes in parameters have lowered and broadened the first
640 > peak of SSD/E and SSD/RF.}
641   \label{grcompare}
642   \end{center}
643   \end{figure}
# Line 644 | Line 646 | and broadened the first peak of {\sc ssd/e} and {\sc s
646   \begin{center}
647   \epsfxsize=6in
648   \epsfbox{dualsticky_bw.eps}
649 < \caption{Positive and negative isosurfaces of the sticky potential for
650 < {\sc ssd1} (left) and {\sc ssd/e} \& {\sc ssd/rf} (right). Light areas correspond to the
651 < tetrahedral attractive component, and darker areas correspond to the
652 < dipolar repulsive component.}
649 > \caption{ Positive and negative isosurfaces of the sticky potential for
650 > SSD1 (left) and SSD/E \& SSD/RF (right). Light areas
651 > correspond to the tetrahedral attractive component, and darker areas
652 > correspond to the dipolar repulsive component.}
653   \label{isosurface}
654   \end{center}
655   \end{figure}
656  
657 < In the original paper detailing the development of {\sc ssd}, Liu and Ichiye
657 > In the original paper detailing the development of SSD, Liu and Ichiye
658   placed particular emphasis on an accurate description of the first
659   solvation shell. This resulted in a somewhat tall and narrow first
660   peak in $g(r)$ that integrated to give similar coordination numbers to
661   the experimental data obtained by Soper and
662   Phillips.\cite{Ichiye96,Soper86} New experimental x-ray scattering
663   data from the Head-Gordon lab indicates a slightly lower and shifted
664 < first peak in the g$_\mathrm{OO}(r)$, so our adjustments to {\sc ssd} were
664 > first peak in the g$_\mathrm{OO}(r)$, so our adjustments to SSD were
665   made after taking into consideration the new experimental
666   findings.\cite{Head-Gordon00_1} Figure \ref{grcompare} shows the
667   relocation of the first peak of the oxygen-oxygen $g(r)$ by comparing
668 < the revised {\sc ssd} model ({\sc ssd1}), {\sc ssd/e}, and {\sc ssd/rf} to the new
668 > the revised SSD model (SSD1), SSD/E, and SSD/RF to the new
669   experimental results. Both modified water models have shorter peaks
670   that match more closely to the experimental peak (as seen in the
671   insets of figure \ref{grcompare}).  This structural alteration was
# Line 676 | Line 678 | density for the overall system.  This change in intera
678   see how altering the cutoffs changes the repulsive and attractive
679   character of the particles. With a reduced repulsive surface (darker
680   region), the particles can move closer to one another, increasing the
681 < density for the overall system.  This change in interaction cutoff also
682 < results in a more gradual orientational motion by allowing the
681 > density for the overall system.  This change in interaction cutoff
682 > also results in a more gradual orientational motion by allowing the
683   particles to maintain preferred dipolar arrangements before they begin
684   to feel the pull of the tetrahedral restructuring. As the particles
685   move closer together, the dipolar repulsion term becomes active and
686   excludes unphysical nearest-neighbor arrangements. This compares with
687 < how {\sc ssd} and {\sc ssd1} exclude preferred dipole alignments before the
687 > how SSD and SSD1 exclude preferred dipole alignments before the
688   particles feel the pull of the ``hydrogen bonds''. Aside from
689   improving the shape of the first peak in the g(\emph{r}), this
690   modification improves the densities considerably by allowing the
691 < persistence of full dipolar character below the previous 4.0 \AA\
691 > persistence of full dipolar character below the previous 4.0~\AA\
692   cutoff.
693  
694   While adjusting the location and shape of the first peak of $g(r)$
695   improves the densities, these changes alone are insufficient to bring
696   the system densities up to the values observed experimentally.  To
697   further increase the densities, the dipole moments were increased in
698 < both of our adjusted models. Since {\sc ssd} is a dipole based model, the
698 > both of our adjusted models. Since SSD is a dipole based model, the
699   structure and transport are very sensitive to changes in the dipole
700 < moment. The original {\sc ssd} simply used the dipole moment calculated from
701 < the TIP3P water model, which at 2.35 D is significantly greater than
702 < the experimental gas phase value of 1.84 D. The larger dipole moment
700 > moment. The original SSD simply used the dipole moment calculated from
701 > the TIP3P water model, which at 2.35~D is significantly greater than
702 > the experimental gas phase value of 1.84~D. The larger dipole moment
703   is a more realistic value and improves the dielectric properties of
704   the fluid. Both theoretical and experimental measurements indicate a
705 < liquid phase dipole moment ranging from 2.4 D to values as high as
706 < 3.11 D, providing a substantial range of reasonable values for a
705 > liquid phase dipole moment ranging from 2.4~D to values as high as
706 > 3.11~D, providing a substantial range of reasonable values for a
707   dipole moment.\cite{Sprik91,Kusalik02,Badyal00,Barriol64} Moderately
708 < increasing the dipole moments to 2.42 and 2.48 D for {\sc ssd/e} and {\sc ssd/rf},
708 > increasing the dipole moments to 2.42 and 2.48~D for SSD/E and SSD/RF,
709   respectively, leads to significant changes in the density and
710   transport of the water models.
711  
712   In order to demonstrate the benefits of these reparameterizations, a
713   series of NPT and NVE simulations were performed to probe the density
714   and transport properties of the adapted models and compare the results
715 < to the original {\sc ssd} model. This comparison involved full NPT melting
716 < sequences for both {\sc ssd/e} and {\sc ssd/rf}, as well as NVE transport
715 > to the original SSD model. This comparison involved full NPT melting
716 > sequences for both SSD/E and SSD/RF, as well as NVE transport
717   calculations at the calculated self-consistent densities. Again, the
718   results are obtained from five separate simulations of 1024 particle
719   systems, and the melting sequences were started from different ice
720   $I_h$ crystals constructed as described previously. Each NPT
721 < simulation was equilibrated for 100 ps before a 200 ps data collection
721 > simulation was equilibrated for 100~ps before a 200~ps data collection
722   run at each temperature step, and the final configuration from the
723   previous temperature simulation was used as a starting point. All NVE
724   simulations had the same thermalization, equilibration, and data
# Line 726 | Line 728 | collection times as stated previously.
728   \begin{center}
729   \epsfxsize=6in
730   \epsfbox{ssdeDense.epsi}
731 < \caption{Comparison of densities calculated with {\sc ssd/e} to {\sc ssd1} without a
732 < reaction field, TIP3P [Ref. \citen{Jorgensen98b}], TIP5P
733 < [Ref. \citen{Jorgensen00}], SPC/E [Ref. \citen{Clancy94}] and
731 > \caption{ Comparison of densities calculated with SSD/E to
732 > SSD1 without a reaction field, TIP3P [Ref. \citen{Jorgensen98b}],
733 > TIP5P [Ref. \citen{Jorgensen00}], SPC/E [Ref. \citen{Clancy94}] and
734   experiment [Ref. \citen{CRC80}]. The window shows a expansion around
735   300 K with error bars included to clarify this region of
736 < interest. Note that both {\sc ssd1} and {\sc ssd/e} show good agreement with
736 > interest. Note that both SSD1 and SSD/E show good agreement with
737   experiment when the long-range correction is neglected.}
738   \label{ssdedense}
739   \end{center}
740   \end{figure}
741  
742 < Fig. \ref{ssdedense} shows the density profile for the {\sc ssd/e} model
743 < in comparison to {\sc ssd1} without a reaction field, other common water
744 < models, and experimental results. The calculated densities for both
745 < {\sc ssd/e} and {\sc ssd1} have increased significantly over the original {\sc ssd}
746 < model (see fig. \ref{dense1}) and are in better agreement with the
747 < experimental values. At 298 K, the densities of {\sc ssd/e} and {\sc ssd1} without
742 > Figure \ref{ssdedense} shows the density profile for the SSD/E
743 > model in comparison to SSD1 without a reaction field, other
744 > common water models, and experimental results. The calculated
745 > densities for both SSD/E and SSD1 have increased
746 > significantly over the original SSD model (see
747 > fig. \ref{dense1}) and are in better agreement with the experimental
748 > values. At 298 K, the densities of SSD/E and SSD1 without
749   a long-range correction are 0.996$\pm$0.001 g/cm$^3$ and
750   0.999$\pm$0.001 g/cm$^3$ respectively.  These both compare well with
751   the experimental value of 0.997 g/cm$^3$, and they are considerably
752 < better than the {\sc ssd} value of 0.967$\pm$0.003 g/cm$^3$. The changes to
753 < the dipole moment and sticky switching functions have improved the
754 < structuring of the liquid (as seen in figure \ref{grcompare}, but they
755 < have shifted the density maximum to much lower temperatures. This
756 < comes about via an increase in the liquid disorder through the
757 < weakening of the sticky potential and strengthening of the dipolar
758 < character. However, this increasing disorder in the {\sc ssd/e} model has
759 < little effect on the melting transition. By monitoring $C_p$
760 < throughout these simulations, the melting transition for {\sc ssd/e} was
761 < shown to occur at 235 K.  The same transition temperature observed
762 < with {\sc ssd} and {\sc ssd1}.
752 > better than the SSD value of 0.967$\pm$0.003 g/cm$^3$. The
753 > changes to the dipole moment and sticky switching functions have
754 > improved the structuring of the liquid (as seen in figure
755 > \ref{grcompare}), but they have shifted the density maximum to much
756 > lower temperatures. This comes about via an increase in the liquid
757 > disorder through the weakening of the sticky potential and
758 > strengthening of the dipolar character. However, this increasing
759 > disorder in the SSD/E model has little effect on the melting
760 > transition. By monitoring $C_p$ throughout these simulations, the
761 > melting transition for SSD/E was shown to occur at 235~K.  The
762 > same transition temperature observed with SSD and SSD1.
763  
764   \begin{figure}
765   \begin{center}
766   \epsfxsize=6in
767   \epsfbox{ssdrfDense.epsi}
768 < \caption{Comparison of densities calculated with {\sc ssd/rf} to {\sc ssd1} with a
769 < reaction field, TIP3P [Ref. \citen{Jorgensen98b}], TIP5P
770 < [Ref. \citen{Jorgensen00}], SPC/E [Ref. \citen{Clancy94}], and
768 > \caption{ Comparison of densities calculated with SSD/RF to
769 > SSD1 with a reaction field, TIP3P [Ref. \citen{Jorgensen98b}],
770 > TIP5P [Ref. \citen{Jorgensen00}], SPC/E [Ref. \citen{Clancy94}], and
771   experiment [Ref. \citen{CRC80}]. The inset shows the necessity of
772   reparameterization when utilizing a reaction field long-ranged
773 < correction - {\sc ssd/rf} provides significantly more accurate densities
774 < than {\sc ssd1} when performing room temperature simulations.}
773 > correction - SSD/RF provides significantly more accurate
774 > densities than SSD1 when performing room temperature
775 > simulations.}
776   \label{ssdrfdense}
777   \end{center}
778   \end{figure}
779  
780   Including the reaction field long-range correction in the simulations
781   results in a more interesting comparison.  A density profile including
782 < {\sc ssd/rf} and {\sc ssd1} with an active reaction field is shown in figure
782 > SSD/RF and SSD1 with an active reaction field is shown in figure
783   \ref{ssdrfdense}.  As observed in the simulations without a reaction
784 < field, the densities of {\sc ssd/rf} and {\sc ssd1} show a dramatic increase over
785 < normal {\sc ssd} (see figure \ref{dense1}). At 298 K, {\sc ssd/rf} has a density
784 > field, the densities of SSD/RF and SSD1 show a dramatic increase over
785 > normal SSD (see figure \ref{dense1}). At 298 K, SSD/RF has a density
786   of 0.997$\pm$0.001 g/cm$^3$, directly in line with experiment and
787 < considerably better than the original {\sc ssd} value of 0.941$\pm$0.001
788 < g/cm$^3$ and the {\sc ssd1} value of 0.972$\pm$0.002 g/cm$^3$. These results
787 > considerably better than the original SSD value of 0.941$\pm$0.001
788 > g/cm$^3$ and the SSD1 value of 0.972$\pm$0.002 g/cm$^3$. These results
789   further emphasize the importance of reparameterization in order to
790   model the density properly under different simulation conditions.
791   Again, these changes have only a minor effect on the melting point,
792 < which observed at 245 K for {\sc ssd/rf}, is identical to {\sc ssd} and only 5 K
793 < lower than {\sc ssd1} with a reaction field. Additionally, the difference in
794 < density maxima is not as extreme, with {\sc ssd/rf} showing a density
795 < maximum at 255 K, fairly close to the density maxima of 260 K and 265
796 < K, shown by {\sc ssd} and {\sc ssd1} respectively.
792 > which observed at 245~K for SSD/RF, is identical to SSD and only 5~K
793 > lower than SSD1 with a reaction field. Additionally, the difference in
794 > density maxima is not as extreme, with SSD/RF showing a density
795 > maximum at 255~K, fairly close to the density maxima of 260~K and
796 > 265~K, shown by SSD and SSD1 respectively.
797  
798   \begin{figure}
799   \begin{center}
800   \epsfxsize=6in
801   \epsfbox{ssdeDiffuse.epsi}
802 < \caption{The diffusion constants calculated from {\sc ssd/e} and {\sc ssd1} (both
803 < without a reaction field) along with experimental results
802 > \caption{ The diffusion constants calculated from SSD/E and
803 > SSD1 (both without a reaction field) along with experimental results
804   [Refs. \citen{Gillen72} and \citen{Holz00}]. The NVE calculations were
805   performed at the average densities observed in the 1 atm NPT
806 < simulations for the respective models. {\sc ssd/e} is slightly more mobile
806 > simulations for the respective models. SSD/E is slightly more mobile
807   than experiment at all of the temperatures, but it is closer to
808 < experiment at biologically relevant temperatures than {\sc ssd1} without a
808 > experiment at biologically relevant temperatures than SSD1 without a
809   long-range correction.}
810   \label{ssdediffuse}
811   \end{center}
812   \end{figure}
813  
814 < The reparameterization of the {\sc ssd} water model, both for use with and
814 > The reparameterization of the SSD water model, both for use with and
815   without an applied long-range correction, brought the densities up to
816   what is expected for simulating liquid water. In addition to improving
817 < the densities, it is important that the diffusive behavior of {\sc ssd} be
817 > the densities, it is important that the diffusive behavior of SSD be
818   maintained or improved. Figure \ref{ssdediffuse} compares the
819 < temperature dependence of the diffusion constant of {\sc ssd/e} to {\sc ssd1}
819 > temperature dependence of the diffusion constant of SSD/E to SSD1
820   without an active reaction field at the densities calculated from
821   their respective NPT simulations at 1 atm. The diffusion constant for
822 < {\sc ssd/e} is consistently higher than experiment, while {\sc ssd1} remains lower
822 > SSD/E is consistently higher than experiment, while SSD1 remains lower
823   than experiment until relatively high temperatures (around 360
824   K). Both models follow the shape of the experimental curve well below
825 < 300 K but tend to diffuse too rapidly at higher temperatures, as seen
826 < in {\sc ssd1}'s crossing above 360 K.  This increasing diffusion relative to
825 > 300~K but tend to diffuse too rapidly at higher temperatures, as seen
826 > in SSD1's crossing above 360~K.  This increasing diffusion relative to
827   the experimental values is caused by the rapidly decreasing system
828 < density with increasing temperature.  Both {\sc ssd1} and {\sc ssd/e} show this
828 > density with increasing temperature.  Both SSD1 and SSD/E show this
829   deviation in particle mobility, but this trend has different
830 < implications on the diffusive behavior of the models.  While {\sc ssd1}
830 > implications on the diffusive behavior of the models.  While SSD1
831   shows more experimentally accurate diffusive behavior in the high
832 < temperature regimes, {\sc ssd/e} shows more accurate behavior in the
832 > temperature regimes, SSD/E shows more accurate behavior in the
833   supercooled and biologically relevant temperature ranges.  Thus, the
834   changes made to improve the liquid structure may have had an adverse
835   affect on the density maximum, but they improve the transport behavior
836 < of {\sc ssd/e} relative to {\sc ssd1} under the most commonly simulated
836 > of SSD/E relative to SSD1 under the most commonly simulated
837   conditions.
838  
839   \begin{figure}
840   \begin{center}
841   \epsfxsize=6in
842   \epsfbox{ssdrfDiffuse.epsi}
843 < \caption{The diffusion constants calculated from {\sc ssd/rf} and {\sc ssd1} (both
844 < with an active reaction field) along with experimental results
845 < [Refs. \citen{Gillen72} and \citen{Holz00}]. The NVE calculations were
846 < performed at the average densities observed in the 1 atm NPT
847 < simulations for both of the models. {\sc ssd/rf} simulates the diffusion of
848 < water throughout this temperature range very well. The rapidly
849 < increasing diffusion constants at high temperatures for both models
850 < can be attributed to lower calculated densities than those observed in
851 < experiment.}
843 > \caption{ The diffusion constants calculated from SSD/RF and
844 > SSD1 (both with an active reaction field) along with
845 > experimental results [Refs. \citen{Gillen72} and \citen{Holz00}]. The
846 > NVE calculations were performed at the average densities observed in
847 > the 1 atm NPT simulations for both of the models. SSD/RF
848 > simulates the diffusion of water throughout this temperature range
849 > very well. The rapidly increasing diffusion constants at high
850 > temperatures for both models can be attributed to lower calculated
851 > densities than those observed in experiment.}
852   \label{ssdrfdiffuse}
853   \end{center}
854   \end{figure}
855  
856 < In figure \ref{ssdrfdiffuse}, the diffusion constants for {\sc ssd/rf} are
857 < compared to {\sc ssd1} with an active reaction field. Note that {\sc ssd/rf}
856 > In figure \ref{ssdrfdiffuse}, the diffusion constants for SSD/RF are
857 > compared to SSD1 with an active reaction field. Note that SSD/RF
858   tracks the experimental results quantitatively, identical within error
859   throughout most of the temperature range shown and exhibiting only a
860 < slight increasing trend at higher temperatures. {\sc ssd1} tends to diffuse
860 > slight increasing trend at higher temperatures. SSD1 tends to diffuse
861   more slowly at low temperatures and deviates to diffuse too rapidly at
862 < temperatures greater than 330 K.  As stated above, this deviation away
862 > temperatures greater than 330~K.  As stated above, this deviation away
863   from the ideal trend is due to a rapid decrease in density at higher
864 < temperatures. {\sc ssd/rf} does not suffer from this problem as much as {\sc ssd1}
864 > temperatures. SSD/RF does not suffer from this problem as much as SSD1
865   because the calculated densities are closer to the experimental
866   values. These results again emphasize the importance of careful
867   reparameterization when using an altered long-range correction.
# Line 866 | Line 870 | reparameterization when using an altered long-range co
870   \begin{minipage}{\linewidth}
871   \renewcommand{\thefootnote}{\thempfootnote}
872   \begin{center}
873 < \caption{Properties of the single-point water models compared with
874 < experimental data at ambient conditions}
873 > \caption{ Properties of the single-point water models compared with
874 > experimental data at ambient conditions. Deviations of the of the
875 > averages are given in parentheses.}
876   \begin{tabular}{ l  c  c  c  c  c }
877   \hline \\[-3mm]
878 < \ \ \ \ \ \  & \ \ \ {\sc ssd1} \ \ \ & \ {\sc ssd/e} \ \ \ & \ {\sc ssd1} (RF) \ \
879 < \ & \ {\sc ssd/rf} \ \ \ & \ Expt. \\
878 > \ \ \ \ \ \  & \ \ \ SSD1 \ \ \ & \ \ SSD/E \ \ \ & \ \ SSD1 (RF) \ \
879 > \ & \ \ SSD/RF \ \ \ & \ \ Expt. \\
880   \hline \\[-3mm]
881 < \ \ \ $\rho$ (g/cm$^3$) & 0.999 $\pm$0.001 & 0.996 $\pm$0.001 & 0.972 $\pm$0.002 & 0.997 $\pm$0.001 & 0.997 \\
882 < \ \ \ $C_p$ (cal/mol K) & 28.80 $\pm$0.11 & 25.45 $\pm$0.09 & 28.28 $\pm$0.06 & 23.83 $\pm$0.16 & 17.98 \\
883 < \ \ \ $D$ ($10^{-5}$ cm$^2$/s) & 1.78 $\pm$0.07 & 2.51 $\pm$0.18 &
884 < 2.00 $\pm$0.17 & 2.32 $\pm$0.06 & 2.299\cite{Mills73} \\
880 < \ \ \ Coordination Number ($n_C$) & 3.9 & 4.3 & 3.8 & 4.4 &
881 > \ \ $\rho$ (g/cm$^3$) & 0.999(0.001) & 0.996(0.001) & 0.972(0.002) & 0.997(0.001) & 0.997 \\
882 > \ \ $C_p$ (cal/mol K) & 28.80(0.11) & 25.45(0.09) & 28.28(0.06) & 23.83(0.16) & 17.98 \\
883 > \ \ $D$ ($10^{-5}$ cm$^2$/s) & 1.78(0.7) & 2.51(0.18) & 2.00(0.17) & 2.32(0.06) & 2.299\cite{Mills73} \\
884 > \ \ Coordination Number ($n_C$) & 3.9 & 4.3 & 3.8 & 4.4 &
885   4.7\footnote{Calculated by integrating $g_{\text{OO}}(r)$ in
886   Ref. \citen{Head-Gordon00_1}} \\
887 < \ \ \ H-bonds per particle ($n_H$) & 3.7 & 3.6 & 3.7 & 3.7 &
887 > \ \ H-bonds per particle ($n_H$) & 3.7 & 3.6 & 3.7 & 3.7 &
888   3.5\footnote{Calculated by integrating $g_{\text{OH}}(r)$ in
889   Ref. \citen{Soper86}}  \\
890 < \ \ \ $\tau_1$ (ps) & 10.9 $\pm$0.6 & 7.3 $\pm$0.4 & 7.5 $\pm$0.7 &
891 < 7.2 $\pm$0.4 & 5.7\footnote{Calculated for 298 K from data in Ref. \citen{Eisenberg69}} \\
888 < \ \ \ $\tau_2$ (ps) & 4.7 $\pm$0.4 & 3.1 $\pm$0.2 & 3.5 $\pm$0.3 & 3.2
889 < $\pm$0.2 & 2.3\footnote{Calculated for 298 K from data in
890 > \ \ $\tau_1$ (ps) & 10.9(0.6) & 7.3(0.4) & 7.5(0.7) & 7.2(0.4) & 5.7\footnote{Calculated for 298 K from data in Ref. \citen{Eisenberg69}} \\
891 > \ \ $\tau_2$ (ps) & 4.7(0.4) & 3.1(0.2) & 3.5(0.3) & 3.2(0.2) & 2.3\footnote{Calculated for 298 K from data in
892   Ref. \citen{Krynicki66}}
893   \end{tabular}
894   \label{liquidproperties}
# Line 909 | Line 911 | but the coordination numbers of {\sc ssd/e} and {\sc s
911   $a$ and $b$ are the radial locations of the minima following the first
912   peak in g$_\text{OO}(r)$ or g$_\text{OH}(r)$ respectively. The number
913   of hydrogen bonds stays relatively constant across all of the models,
914 < but the coordination numbers of {\sc ssd/e} and {\sc ssd/rf} show an improvement
915 < over {\sc ssd1}.  This improvement is primarily due to extension of the
916 < first solvation shell in the new parameter sets.  Because $n_H$ and
917 < $n_C$ are nearly identical in {\sc ssd1}, it appears that all molecules in
918 < the first solvation shell are involved in hydrogen bonds.  Since $n_H$
919 < and $n_C$ differ in the newly parameterized models, the orientations
920 < in the first solvation shell are a bit more ``fluid''.  Therefore {\sc ssd1}
921 < overstructures the first solvation shell and our reparameterizations
922 < have returned this shell to more realistic liquid-like behavior.
914 > but the coordination numbers of SSD/E and SSD/RF show an
915 > improvement over SSD1.  This improvement is primarily due to
916 > extension of the first solvation shell in the new parameter sets.
917 > Because $n_H$ and $n_C$ are nearly identical in SSD1, it appears
918 > that all molecules in the first solvation shell are involved in
919 > hydrogen bonds.  Since $n_H$ and $n_C$ differ in the newly
920 > parameterized models, the orientations in the first solvation shell
921 > are a bit more ``fluid''.  Therefore SSD1 overstructures the
922 > first solvation shell and our reparameterizations have returned this
923 > shell to more realistic liquid-like behavior.
924  
925   The time constants for the orientational autocorrelation functions
926   are also displayed in Table \ref{liquidproperties}. The dipolar
# Line 937 | Line 940 | dependence of $\tau_2$, it is necessary to recalculate
940   the commonly cited value of 1.9 ps for $\tau_2$ was determined from
941   the NMR data in Ref. \citen{Krynicki66} at a temperature near
942   34$^\circ$C.\cite{Rahman71} Because of the strong temperature
943 < dependence of $\tau_2$, it is necessary to recalculate it at 298 K to
943 > dependence of $\tau_2$, it is necessary to recalculate it at 298~K to
944   make proper comparisons. The value shown in Table
945   \ref{liquidproperties} was calculated from the same NMR data in the
946   fashion described in Ref. \citen{Krynicki66}. Similarly, $\tau_1$ was
947 < recomputed for 298 K from the data in Ref. \citen{Eisenberg69}.
948 < Again, {\sc ssd/e} and {\sc ssd/rf} show improved behavior over {\sc ssd1}, both with
947 > recomputed for 298~K from the data in Ref. \citen{Eisenberg69}.
948 > Again, SSD/E and SSD/RF show improved behavior over SSD1, both with
949   and without an active reaction field. Turning on the reaction field
950 < leads to much improved time constants for {\sc ssd1}; however, these results
950 > leads to much improved time constants for SSD1; however, these results
951   also include a corresponding decrease in system density.
952 < Orientational relaxation times published in the original {\sc ssd} dynamics
952 > Orientational relaxation times published in the original SSD dynamics
953   papers are smaller than the values observed here, and this difference
954   can be attributed to the use of the Ewald sum.\cite{Ichiye99}
955  
# Line 956 | Line 959 | can be attributed to the use of the Ewald sum.\cite{Ic
959   \begin{center}
960   \epsfxsize=6in
961   \epsfbox{icei_bw.eps}
962 < \caption{The most stable crystal structure assumed by the {\sc ssd} family
962 > \caption{ The most stable crystal structure assumed by the SSD family
963   of water models.  We refer to this structure as Ice-{\it i} to
964   indicate its origins in computer simulation.  This image was taken of
965   the (001) face of the crystal.}
# Line 965 | Line 968 | of {\sc ssd/e} not discussed in this paper, we observe
968   \end{figure}
969  
970   While performing a series of melting simulations on an early iteration
971 < of {\sc ssd/e} not discussed in this paper, we observed recrystallization
972 < into a novel structure not previously known for water.  After melting
973 < at 235 K, two of five systems underwent crystallization events near
974 < 245 K.  The two systems remained crystalline up to 320 and 330 K,
975 < respectively.  The crystal exhibits an expanded zeolite-like structure
976 < that does not correspond to any known form of ice.  This appears to be
977 < an artifact of the point dipolar models, so to distinguish it from the
978 < experimentally observed forms of ice, we have denoted the structure
971 > of SSD/E not discussed in this paper, we observed
972 > recrystallization into a novel structure not previously known for
973 > water.  After melting at 235~K, two of five systems underwent
974 > crystallization events near 245~K.  The two systems remained
975 > crystalline up to 320 and 330~K, respectively.  The crystal exhibits
976 > an expanded zeolite-like structure that does not correspond to any
977 > known form of ice.  This appears to be an artifact of the point
978 > dipolar models, so to distinguish it from the experimentally observed
979 > forms of ice, we have denoted the structure
980   Ice-$\sqrt{\smash[b]{-\text{I}}}$ (Ice-{\it i}).  A large enough
981   portion of the sample crystallized that we have been able to obtain a
982   near ideal crystal structure of Ice-{\it i}. Figure \ref{weirdice}
# Line 987 | Line 991 | structure for at least the {\sc ssd/e} model. To verif
991   favorable enough to stabilize an entire crystal generated around them.
992  
993   Initial simulations indicated that Ice-{\it i} is the preferred ice
994 < structure for at least the {\sc ssd/e} model. To verify this, a
995 < comparison was made between near ideal crystals of ice~$I_h$,
996 < ice~$I_c$, and Ice-{\it i} at constant pressure with {\sc ssd/e}, {\sc
997 < ssd/rf}, and {\sc ssd1}. Near-ideal versions of the three types of
998 < crystals were cooled to 1 K, and enthalpies of formation of each were
999 < compared using all three water models.  Enthalpies were estimated from
1000 < the isobaric-isothermal simulations using $H=U+P_{\text ext}V$ where
1001 < $P_{\text ext}$ is the applied pressure.  A constant value of
1002 < -60.158 kcal / mol has been added to place our zero for the
1003 < enthalpies of formation for these systems at the traditional state
1004 < (elemental forms at standard temperature and pressure).  With every
1005 < model in the {\sc ssd} family, Ice-{\it i} had the lowest calculated
1002 < enthalpy of formation.
994 > structure for at least the SSD/E model. To verify this, a comparison
995 > was made between near ideal crystals of ice~$I_h$, ice~$I_c$, and
996 > Ice-{\it i} at constant pressure with SSD/E, SSD/RF, and
997 > SSD1. Near-ideal versions of the three types of crystals were cooled
998 > to 1 K, and enthalpies of formation of each were compared using all
999 > three water models.  Enthalpies were estimated from the
1000 > isobaric-isothermal simulations using $H=U+P_{\text ext}V$ where
1001 > $P_{\text ext}$ is the applied pressure.  A constant value of -60.158
1002 > kcal / mol has been added to place our zero for the enthalpies of
1003 > formation for these systems at the traditional state (elemental forms
1004 > at standard temperature and pressure).  With every model in the SSD
1005 > family, Ice-{\it i} had the lowest calculated enthalpy of formation.
1006  
1007   \begin{table}
1008   \begin{center}
1009 < \caption{Enthalpies of Formation (in kcal / mol) of the three crystal
1010 < structures (at 1 K) exhibited by the {\sc ssd} family of water models}
1009 > \caption{ Enthalpies of Formation (in kcal / mol) of the three crystal
1010 > structures (at 1 K) exhibited by the SSD family of water models}
1011   \begin{tabular}{ l  c  c  c  }
1012   \hline \\[-3mm]
1013 < \ \ \ Water Model \ \ \  & \ \ \ Ice-$I_h$ \ \ \ & \ Ice-$I_c$\ \  & \
1014 < Ice-{\it i} \\
1013 > \ \ \ Water Model \ \ \  & \ \ \ Ice-$I_h$ \ \ \ & \ \ \ Ice-$I_c$ \ \ \  &
1014 > \ \ \ \ Ice-{\it i} \\
1015   \hline \\[-3mm]
1016 < \ \ \ {\sc ssd/e} & -12.286 & -12.292 & -13.590 \\
1017 < \ \ \ {\sc ssd/rf} & -12.935 & -12.917 & -14.022 \\
1018 < \ \ \ {\sc ssd1} & -12.496 & -12.411 & -13.417 \\
1019 < \ \ \ {\sc ssd1} (RF) & -12.504 & -12.411 & -13.134 \\
1016 > \ \ \ SSD/E & -72.444 & -72.450 & -73.748 \\
1017 > \ \ \ SSD/RF & -73.093 & -73.075 & -74.180 \\
1018 > \ \ \ SSD1 & -72.654 & -72.569 & -73.575 \\
1019 > \ \ \ SSD1 (RF) & -72.662 & -72.569 & -73.292 \\
1020   \end{tabular}
1021   \label{iceenthalpy}
1022   \end{center}
1023   \end{table}
1024  
1025   In addition to these energetic comparisons, melting simulations were
1026 < performed with ice-{\it i} as the initial configuration using {\sc ssd/e},
1027 < {\sc ssd/rf}, and {\sc ssd1} both with and without a reaction field. The melting
1028 < transitions for both {\sc ssd/e} and {\sc ssd1} without reaction field occurred at
1029 < temperature in excess of 375~K.  {\sc ssd/rf} and {\sc ssd1} with a reaction field
1026 > performed with Ice-{\it i} as the initial configuration using SSD/E,
1027 > SSD/RF, and SSD1 both with and without a reaction field. The melting
1028 > transitions for both SSD/E and SSD1 without reaction field occurred at
1029 > temperature in excess of 375~K.  SSD/RF and SSD1 with a reaction field
1030   showed more reasonable melting transitions near 325~K.  These melting
1031 < point observations clearly show that all of the {\sc ssd}-derived models
1031 > point observations clearly show that all of the SSD-derived models
1032   prefer the ice-{\it i} structure.
1033  
1034   \section{Conclusions}
1035  
1036   The density maximum and temperature dependence of the self-diffusion
1037 < constant were studied for the {\sc ssd} water model, both with and without
1038 < the use of reaction field, via a series of NPT and NVE
1037 > constant were studied for the SSD water model, both with and
1038 > without the use of reaction field, via a series of NPT and NVE
1039   simulations. The constant pressure simulations showed a density
1040   maximum near 260 K. In most cases, the calculated densities were
1041   significantly lower than the densities obtained from other water
1042 < models (and experiment). Analysis of self-diffusion showed {\sc ssd} to
1043 < capture the transport properties of water well in both the liquid and
1044 < supercooled liquid regimes.
1042 > models (and experiment). Analysis of self-diffusion showed SSD
1043 > to capture the transport properties of water well in both the liquid
1044 > and supercooled liquid regimes.
1045  
1046 < In order to correct the density behavior, the original {\sc ssd} model was
1047 < reparameterized for use both with and without a reaction field ({\sc ssd/rf}
1048 < and {\sc ssd/e}), and comparisons were made with {\sc ssd1}, Ichiye's density
1049 < corrected version of {\sc ssd}. Both models improve the liquid structure,
1046 > In order to correct the density behavior, the original SSD model was
1047 > reparameterized for use both with and without a reaction field (SSD/RF
1048 > and SSD/E), and comparisons were made with SSD1, Ichiye's density
1049 > corrected version of SSD. Both models improve the liquid structure,
1050   densities, and diffusive properties under their respective simulation
1051   conditions, indicating the necessity of reparameterization when
1052   changing the method of calculating long-range electrostatic
# Line 1052 | Line 1055 | by the {\sc ssd} family of water models is somewhat tr
1055   simulations of biochemical systems.
1056  
1057   The existence of a novel low-density ice structure that is preferred
1058 < by the {\sc ssd} family of water models is somewhat troubling, since liquid
1059 < simulations on this family of water models at room temperature are
1060 < effectively simulations of supercooled or metastable liquids.  One
1058 > by the SSD family of water models is somewhat troubling, since
1059 > liquid simulations on this family of water models at room temperature
1060 > are effectively simulations of supercooled or metastable liquids.  One
1061   way to destabilize this unphysical ice structure would be to make the
1062   range of angles preferred by the attractive part of the sticky
1063   potential much narrower.  This would require extensive
# Line 1078 | Line 1081 | DMR-0079647.
1081   \newpage
1082  
1083   \bibliographystyle{jcp}
1084 < \bibliography{nptSSD}
1084 > \bibliography{nptSSD}
1085  
1083 %\pagebreak
1086  
1087   \end{document}

Diff Legend

Removed lines
+ Added lines
< Changed lines
> Changed lines