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 1033 by chrisfen, Fri Feb 6 19:15:44 2004 UTC

# Line 1 | Line 1
1   %\documentclass[prb,aps,times,twocolumn,tabularx]{revtex4}
2 < \documentclass[11pt]{article}
3 < \usepackage{endfloat}
2 > \documentclass[preprint,aps]{revtex4}
3 > %\documentclass[11pt]{article}
4 > %\usepackage{endfloat}
5   \usepackage{amsmath}
6   \usepackage{epsf}
7   \usepackage{berkeley}
8   \usepackage{setspace}
9   \usepackage{tabularx}
10   \usepackage{graphicx}
11 < \usepackage[ref]{overcite}
11 < %\usepackage{berkeley}
11 > %\usepackage[ref]{overcite}
12   %\usepackage{curves}
13 < \pagestyle{plain}
14 < \pagenumbering{arabic}
15 < \oddsidemargin 0.0cm \evensidemargin 0.0cm
16 < \topmargin -21pt \headsep 10pt
17 < \textheight 9.0in \textwidth 6.5in
18 < \brokenpenalty=10000
19 < \renewcommand{\baselinestretch}{1.2}
20 < \renewcommand\citemid{\ } % no comma in optional reference note
13 > %\pagestyle{plain}
14 > %\pagenumbering{arabic}
15 > %\oddsidemargin 0.0cm \evensidemargin 0.0cm
16 > %\topmargin -21pt \headsep 10pt
17 > %\textheight 9.0in \textwidth 6.5in
18 > %\brokenpenalty=10000
19 > %\renewcommand{\baselinestretch}{1.2}
20 > %\renewcommand\citemid{\ } % no comma in optional reference note
21 > \newcounter{captions}
22 > \newcounter{figs}
23  
24   \begin{document}
25  
26   \title{On the structural and transport properties of the soft sticky
27 < dipole ({\sc ssd}) and related single point water models}
27 > dipole (SSD) and related single point water models}
28  
29 < \author{Christopher J. Fennell and J. Daniel Gezelter\footnote{Corresponding author. \ Electronic mail: gezelter@nd.edu} \\
30 < Department of Chemistry and Biochemistry\\ University of Notre Dame\\
29 > \author{Christopher J. Fennell and J. Daniel Gezelter\footnote{Corresponding author. \ Electronic mail: gezelter@nd.edu}}
30 >
31 > \affiliation{Department of Chemistry and Biochemistry\\ University of Notre Dame\\
32   Notre Dame, Indiana 46556}
33  
34   \date{\today}
35  
33 \maketitle
36  
37   \begin{abstract}
38   The density maximum and temperature dependence of the self-diffusion
39 < constant were investigated for the soft sticky dipole ({\sc ssd}) water
40 < model and two related re-parameterizations of this single-point model.
39 > constant were investigated for the soft sticky dipole (SSD) water
40 > model and two related reparameterizations of this single-point model.
41   A combination of microcanonical and isobaric-isothermal molecular
42   dynamics simulations were used to calculate these properties, both
43   with and without the use of reaction field to handle long-range
44   electrostatics.  The isobaric-isothermal (NPT) simulations of the
45   melting of both ice-$I_h$ and ice-$I_c$ showed a density maximum near
46 < 260 K.  In most cases, the use of the reaction field resulted in
46 > 260~K.  In most cases, the use of the reaction field resulted in
47   calculated densities which were were significantly lower than
48   experimental densities.  Analysis of self-diffusion constants shows
49 < that the original {\sc ssd} model captures the transport properties of
49 > that the original SSD model captures the transport properties of
50   experimental water very well in both the normal and super-cooled
51 < liquid regimes.  We also present our re-parameterized versions of {\sc ssd}
51 > liquid regimes.  We also present our reparameterized versions of SSD
52   for use both with the reaction field or without any long-range
53 < electrostatic corrections.  These are called the {\sc ssd/rf} and {\sc ssd/e}
53 > electrostatic corrections.  These are called the SSD/RF and SSD/E
54   models respectively.  These modified models were shown to maintain or
55   improve upon the experimental agreement with the structural and
56 < transport properties that can be obtained with either the original {\sc ssd}
57 < or the density corrected version of the original model ({\sc ssd1}).
56 > transport properties that can be obtained with either the original SSD
57 > or the density corrected version of the original model (SSD1).
58   Additionally, a novel low-density ice structure is presented
59 < which appears to be the most stable ice structure for the entire {\sc ssd}
59 > which appears to be the most stable ice structure for the entire SSD
60   family.
61   \end{abstract}
62  
63 + \maketitle
64 +
65   \newpage
66  
67   %\narrowtext
# Line 89 | Line 93 | cost is the Soft Sticky Dipole ({\sc ssd}) water
93  
94   One recently developed model that largely succeeds in retaining the
95   accuracy of bulk properties while greatly reducing the computational
96 < cost is the Soft Sticky Dipole ({\sc ssd}) water
97 < model.\cite{Ichiye96,Ichiye96b,Ichiye99,Ichiye03} The {\sc ssd} model was
98 < developed by Ichiye \emph{et al.} as a modified form of the
96 > cost is the Soft Sticky Dipole (SSD) water
97 > model.\cite{Ichiye96,Ichiye96b,Ichiye99,Ichiye03} The SSD model
98 > was developed by Ichiye \emph{et al.} as a modified form of the
99   hard-sphere water model proposed by Bratko, Blum, and
100 < Luzar.\cite{Bratko85,Bratko95} {\sc ssd} is a {\it single point} model which
101 < has an interaction site that is both a point dipole along with a
100 > Luzar.\cite{Bratko85,Bratko95} SSD is a {\it single point} model
101 > which has an interaction site that is both a point dipole and a
102   Lennard-Jones core.  However, since the normal aligned and
103   anti-aligned geometries favored by point dipoles are poor mimics of
104   local structure in liquid water, a short ranged ``sticky'' potential
105   is also added.  The sticky potential directs the molecules to assume
106 < the proper hydrogen bond orientation in the first solvation
103 < shell.  
106 > the proper hydrogen bond orientation in the first solvation shell.
107  
108 < The interaction between two {\sc ssd} water molecules \emph{i} and \emph{j}
108 > The interaction between two SSD water molecules \emph{i} and \emph{j}
109   is given by the potential
110   \begin{equation}
111   u_{ij} = u_{ij}^{LJ} (r_{ij})\ + u_{ij}^{dp}
# Line 159 | Line 162 | can be found in the original {\sc ssd}
162   enhances the tetrahedral geometry for hydrogen bonded structures),
163   while $w^\prime$ is a purely empirical function.  A more detailed
164   description of the functional parts and variables in this potential
165 < can be found in the original {\sc ssd}
165 > can be found in the original SSD
166   articles.\cite{Ichiye96,Ichiye96b,Ichiye99,Ichiye03}
167  
168 < Since {\sc ssd} is a single-point {\it dipolar} model, the force
168 > Since SSD is a single-point {\it dipolar} model, the force
169   calculations are simplified significantly relative to the standard
170   {\it charged} multi-point models. In the original Monte Carlo
171 < simulations using this model, Ichiye {\it et al.} reported that using
172 < {\sc ssd} decreased computer time by a factor of 6-7 compared to other
171 > simulations using this model, Liu and Ichiye reported that using SSD
172 > decreased computer time by a factor of 6-7 compared to other
173   models.\cite{Ichiye96} What is most impressive is that this savings
174   did not come at the expense of accurate depiction of the liquid state
175 < properties.  Indeed, {\sc ssd} maintains reasonable agreement with the Soper
175 > properties.  Indeed, SSD maintains reasonable agreement with the Soper
176   data for the structural features of liquid
177   water.\cite{Soper86,Ichiye96} Additionally, the dynamical properties
178 < exhibited by {\sc ssd} agree with experiment better than those of more
178 > exhibited by SSD agree with experiment better than those of more
179   computationally expensive models (like TIP3P and
180   SPC/E).\cite{Ichiye99} The combination of speed and accurate depiction
181 < of solvent properties makes {\sc ssd} a very attractive model for the
181 > of solvent properties makes SSD a very attractive model for the
182   simulation of large scale biochemical simulations.
183  
184 < One feature of the {\sc ssd} model is that it was parameterized for use with
185 < the Ewald sum to handle long-range interactions.  This would normally
186 < be the best way of handling long-range interactions in systems that
187 < contain other point charges.  However, our group has recently become
188 < interested in systems with point dipoles as mimics for neutral, but
189 < polarized regions on molecules (e.g. the zwitterionic head group
190 < regions of phospholipids).  If the system of interest does not contain
191 < point charges, the Ewald sum and even particle-mesh Ewald become
192 < computational bottlenecks.  Their respective ideal $N^\frac{3}{2}$ and
193 < $N\log N$ calculation scaling orders for $N$ particles can become
194 < prohibitive when $N$ becomes large.\cite{Darden99} In applying this
195 < water model in these types of systems, it would be useful to know its
196 < properties and behavior under the more computationally efficient
197 < reaction field (RF) technique, or even with a simple cutoff. This
198 < study addresses these issues by looking at the structural and
199 < transport behavior of {\sc ssd} over a variety of temperatures with the
200 < purpose of utilizing the RF correction technique.  We then suggest
201 < modifications to the parameters that result in more realistic bulk
202 < phase behavior.  It should be noted that in a recent publication, some
203 < of the original investigators of the {\sc ssd} water model have suggested
204 < adjustments to the {\sc ssd} water model to address abnormal density
205 < behavior (also observed here), calling the corrected model
206 < {\sc ssd1}.\cite{Ichiye03} In what follows, we compare our
207 < reparamaterization of {\sc ssd} with both the original {\sc ssd} and {\sc ssd1} models
208 < with the goal of improving the bulk phase behavior of an {\sc ssd}-derived
209 < model in simulations utilizing the Reaction Field.
184 > One feature of the SSD model is that it was parameterized for
185 > use with the Ewald sum to handle long-range interactions.  This would
186 > normally be the best way of handling long-range interactions in
187 > systems that contain other point charges.  However, our group has
188 > recently become interested in systems with point dipoles as mimics for
189 > neutral, but polarized regions on molecules (e.g. the zwitterionic
190 > head group regions of phospholipids).  If the system of interest does
191 > not contain point charges, the Ewald sum and even particle-mesh Ewald
192 > become computational bottlenecks.  Their respective ideal
193 > $N^\frac{3}{2}$ and $N\log N$ calculation scaling orders for $N$
194 > particles can become prohibitive when $N$ becomes
195 > large.\cite{Darden99} In applying this water model in these types of
196 > systems, it would be useful to know its properties and behavior under
197 > the more computationally efficient reaction field (RF) technique, or
198 > even with a simple cutoff. This study addresses these issues by
199 > looking at the structural and transport behavior of SSD over a
200 > variety of temperatures with the purpose of utilizing the RF
201 > correction technique.  We then suggest modifications to the parameters
202 > that result in more realistic bulk phase behavior.  It should be noted
203 > that in a recent publication, some of the original investigators of
204 > the SSD water model have suggested adjustments to the SSD
205 > water model to address abnormal density behavior (also observed here),
206 > calling the corrected model SSD1.\cite{Ichiye03} In what
207 > follows, we compare our reparamaterization of SSD with both the
208 > original SSD and SSD1 models with the goal of improving
209 > the bulk phase behavior of an SSD-derived model in simulations
210 > utilizing the reaction field.
211  
212   \section{Methods}
213  
214   Long-range dipole-dipole interactions were accounted for in this study
215 < by using either the reaction field method or by resorting to a simple
216 < cubic switching function at a cutoff radius.  The reaction field
217 < method was actually first used in Monte Carlo simulations of liquid
218 < water.\cite{Barker73} Under this method, the magnitude of the reaction
219 < field acting on dipole $i$ is
215 > by using either the reaction field technique or by resorting to a
216 > simple cubic switching function at a cutoff radius.  One of the early
217 > applications of a reaction field was actually in Monte Carlo
218 > simulations of liquid water.\cite{Barker73} Under this method, the
219 > magnitude of the reaction field acting on dipole $i$ is
220   \begin{equation}
221   \mathcal{E}_{i} = \frac{2(\varepsilon_{s} - 1)}{2\varepsilon_{s} + 1}
222   \frac{1}{r_{c}^{3}} \sum_{j\in{\mathcal{R}}} {\bf \mu}_{j} s(r_{ij}),
# Line 226 | Line 230 | field is known to alter the bulk orientational propert
230   total energy by particle $i$ is given by $-\frac{1}{2}{\bf
231   \mu}_{i}\cdot\mathcal{E}_{i}$ and the torque on dipole $i$ by ${\bf
232   \mu}_{i}\times\mathcal{E}_{i}$.\cite{AllenTildesley}  Use of the reaction
233 < field is known to alter the bulk orientational properties, such as the
234 < dielectric relaxation time.  There is particular sensitivity of this
235 < property on changes in the length of the cutoff
236 < radius.\cite{Berendsen98} This variable behavior makes reaction field
237 < a less attractive method than the Ewald sum.  However, for very large
238 < systems, the computational benefit of reaction field is dramatic.
233 > field is known to alter the bulk orientational properties of simulated
234 > water, and there is particular sensitivity of these properties on
235 > changes in the length of the cutoff radius.\cite{Berendsen98} This
236 > variable behavior makes reaction field a less attractive method than
237 > the Ewald sum.  However, for very large systems, the computational
238 > benefit of reaction field is dramatic.
239  
240   We have also performed a companion set of simulations {\it without} a
241   surrounding dielectric (i.e. using a simple cubic switching function
242   at the cutoff radius), and as a result we have two reparamaterizations
243 < of {\sc ssd} which could be used either with or without the reaction field
244 < turned on.
243 > of SSD which could be used either with or without the reaction
244 > field turned on.
245  
246   Simulations to obtain the preferred densities of the models were
247   performed in the isobaric-isothermal (NPT) ensemble, while all
# Line 247 | Line 251 | simulations of {\sc ssd}, because there are no explici
251   implemented using an integral thermostat and barostat as outlined by
252   Hoover.\cite{Hoover85,Hoover86} All molecules were treated as
253   non-linear rigid bodies. Vibrational constraints are not necessary in
254 < simulations of {\sc ssd}, because there are no explicit hydrogen atoms, and
255 < thus no molecular vibrational modes need to be considered.
254 > simulations of SSD, because there are no explicit hydrogen
255 > atoms, and thus no molecular vibrational modes need to be considered.
256  
257   Integration of the equations of motion was carried out using the
258   symplectic splitting method proposed by Dullweber, Leimkuhler, and
259 < McLachlan ({\sc dlm}).\cite{Dullweber1997} Our reason for selecting this
260 < integrator centers on poor energy conservation of rigid body dynamics
261 < using traditional quaternion integration.\cite{Evans77,Evans77b} In
262 < typical microcanonical ensemble simulations, the energy drift when
263 < using quaternions was substantially greater than when using the {\sc dlm}
264 < method (fig. \ref{timestep}).  This steady drift in the total energy
265 < has also been observed by Kol {\it et al.}\cite{Laird97}
259 > McLachlan ({\sc dlm}).\cite{Dullweber1997} Our reason for selecting
260 > this integrator centers on poor energy conservation of rigid body
261 > dynamics using traditional quaternion
262 > integration.\cite{Evans77,Evans77b} In typical microcanonical ensemble
263 > simulations, the energy drift when using quaternions was substantially
264 > greater than when using the {\sc dlm} method (fig. \ref{timestep}).
265 > This steady drift in the total energy has also been observed by Kol
266 > {\it et al.}\cite{Laird97}
267  
268   The key difference in the integration method proposed by Dullweber
269   \emph{et al.} is that the entire rotation matrix is propagated from
# Line 273 | Line 278 | step, a 1000 {\sc ssd} particle simulation shows an av
278   of matrix evaluations to update the rotation
279   matrix.\cite{Dullweber1997} These matrix rotations are more costly
280   than the simpler arithmetic quaternion propagation. With the same time
281 < step, a 1000 {\sc ssd} particle simulation shows an average 7\% increase in
282 < computation time using the {\sc dlm} method in place of quaternions. The
283 < additional expense per step is justified when one considers the
284 < ability to use time steps that are nearly twice as large under {\sc dlm}
285 < than would be usable under quaternion dynamics.  The energy
286 < conservation of the two methods using a number of different time steps
287 < is illustrated in figure
281 > step, a 1000 SSD particle simulation shows an average 7\%
282 > increase in computation time using the {\sc dlm} method in place of
283 > quaternions. The additional expense per step is justified when one
284 > considers the ability to use time steps that are nearly twice as large
285 > under {\sc dlm} than would be usable under quaternion dynamics.  The
286 > energy conservation of the two methods using a number of different
287 > time steps is illustrated in figure
288   \ref{timestep}.
289  
290 < \begin{figure}
291 < \begin{center}
292 < \epsfxsize=6in
293 < \epsfbox{timeStep.epsi}
294 < \caption{Energy conservation using both quaternion-based integration and
295 < the {\sc dlm} method with increasing time step. The larger time step plots
296 < are shifted from the true energy baseline (that of $\Delta t$ = 0.1
297 < fs) for clarity.}
298 < \label{timestep}
299 < \end{center}
300 < \end{figure}
290 > %\begin{figure}
291 > %\begin{center}
292 > %\epsfxsize=6in
293 > %\epsfbox{timeStep.epsi}
294 > %\caption{Energy conservation using both quaternion-based integration and
295 > %the {\sc dlm} method with increasing time step. The larger time step
296 > %plots are shifted from the true energy baseline (that of $\Delta t$ =
297 > %0.1~fs) for clarity.}
298 > %\label{timestep}
299 > %\end{center}
300 > %\end{figure}
301  
302   In figure \ref{timestep}, the resulting energy drift at various time
303 < steps for both the {\sc dlm} and quaternion integration schemes is compared.
304 < All of the 1000 {\sc ssd} particle simulations started with the same
305 < configuration, and the only difference was the method used to handle
306 < orientational motion. At time steps of 0.1 and 0.5 fs, both methods
307 < for propagating the orientational degrees of freedom conserve energy
308 < fairly well, with the quaternion method showing a slight energy drift
309 < over time in the 0.5 fs time step simulation. At time steps of 1 and 2
310 < fs, the energy conservation benefits of the {\sc dlm} method are clearly
311 < demonstrated. Thus, while maintaining the same degree of energy
312 < conservation, one can take considerably longer time steps, leading to
313 < an overall reduction in computation time.
303 > steps for both the {\sc dlm} and quaternion integration schemes is
304 > compared.  All of the 1000 SSD particle simulations started with
305 > the same configuration, and the only difference was the method used to
306 > handle orientational motion. At time steps of 0.1 and 0.5~fs, both
307 > methods for propagating the orientational degrees of freedom conserve
308 > energy fairly well, with the quaternion method showing a slight energy
309 > drift over time in the 0.5~fs time step simulation. At time steps of 1
310 > and 2~fs, the energy conservation benefits of the {\sc dlm} method are
311 > clearly demonstrated. Thus, while maintaining the same degree of
312 > energy conservation, one can take considerably longer time steps,
313 > leading to an overall reduction in computation time.
314  
315 < Energy drift in the simulations using {\sc dlm} integration was unnoticeable
316 < for time steps up to 3 fs. A slight energy drift on the order of 0.012
317 < kcal/mol per nanosecond was observed at a time step of 4 fs, and as
318 < expected, this drift increases dramatically with increasing time
319 < step. To insure accuracy in our microcanonical simulations, time steps
320 < were set at 2 fs and kept at this value for constant pressure
321 < simulations as well.
315 > Energy drift in the simulations using {\sc dlm} integration was
316 > unnoticeable for time steps up to 3~fs. A slight energy drift on the
317 > order of 0.012~kcal/mol per nanosecond was observed at a time step of
318 > 4~fs, and as expected, this drift increases dramatically with
319 > increasing time step. To insure accuracy in our microcanonical
320 > simulations, time steps were set at 2~fs and kept at this value for
321 > constant pressure simulations as well.
322  
323   Proton-disordered ice crystals in both the $I_h$ and $I_c$ lattices
324   were generated as starting points for all simulations. The $I_h$
325 < crystals were formed by first arranging the centers of mass of the {\sc ssd}
325 > crystals were formed by first arranging the centers of mass of the SSD
326   particles into a ``hexagonal'' ice lattice of 1024 particles. Because
327   of the crystal structure of $I_h$ ice, the simulation box assumed an
328   orthorhombic shape with an edge length ratio of approximately
329   1.00$\times$1.06$\times$1.23. The particles were then allowed to
330   orient freely about fixed positions with angular momenta randomized at
331 < 400 K for varying times. The rotational temperature was then scaled
332 < down in stages to slowly cool the crystals to 25 K. The particles were
331 > 400~K for varying times. The rotational temperature was then scaled
332 > down in stages to slowly cool the crystals to 25~K. The particles were
333   then allowed to translate with fixed orientations at a constant
334 < pressure of 1 atm for 50 ps at 25 K. Finally, all constraints were
335 < removed and the ice crystals were allowed to equilibrate for 50 ps at
336 < 25 K and a constant pressure of 1 atm.  This procedure resulted in
334 > pressure of 1 atm for 50~ps at 25~K. Finally, all constraints were
335 > removed and the ice crystals were allowed to equilibrate for 50~ps at
336 > 25~K and a constant pressure of 1~atm.  This procedure resulted in
337   structurally stable $I_h$ ice crystals that obey the Bernal-Fowler
338   rules.\cite{Bernal33,Rahman72} This method was also utilized in the
339   making of diamond lattice $I_c$ ice crystals, with each cubic
# Line 346 | Line 351 | for 100 ps prior to a 200 ps data collection run at ea
351   supercooled regime. An ensemble average from five separate melting
352   simulations was acquired, each starting from different ice crystals
353   generated as described previously. All simulations were equilibrated
354 < for 100 ps prior to a 200 ps data collection run at each temperature
355 < setting. The temperature range of study spanned from 25 to 400 K, with
356 < a maximum degree increment of 25 K. For regions of interest along this
357 < stepwise progression, the temperature increment was decreased from 25
358 < K to 10 and 5 K.  The above equilibration and production times were
354 > for 100~ps prior to a 200~ps data collection run at each temperature
355 > setting. The temperature range of study spanned from 25 to 400~K, with
356 > a maximum degree increment of 25~K. For regions of interest along this
357 > stepwise progression, the temperature increment was decreased from
358 > 25~K to 10 and 5~K.  The above equilibration and production times were
359   sufficient in that fluctuations in the volume autocorrelation function
360 < were damped out in all simulations in under 20 ps.
360 > were damped out in all simulations in under 20~ps.
361  
362   \subsection{Density Behavior}
363  
364 < Our initial simulations focused on the original {\sc ssd} water model, and
365 < an average density versus temperature plot is shown in figure
364 > Our initial simulations focused on the original SSD water model,
365 > and an average density versus temperature plot is shown in figure
366   \ref{dense1}. Note that the density maximum when using a reaction
367 < field appears between 255 and 265 K.  There were smaller fluctuations
368 < in the density at 260 K than at either 255 or 265, so we report this
367 > field appears between 255 and 265~K.  There were smaller fluctuations
368 > in the density at 260~K than at either 255 or 265~K, so we report this
369   value as the location of the density maximum. Figure \ref{dense1} was
370   constructed using ice $I_h$ crystals for the initial configuration;
371   though not pictured, the simulations starting from ice $I_c$ crystal
372   configurations showed similar results, with a liquid-phase density
373 < maximum in this same region (between 255 and 260 K).
373 > maximum in this same region (between 255 and 260~K).
374  
375 < \begin{figure}
376 < \begin{center}
377 < \epsfxsize=6in
378 < \epsfbox{denseSSD.eps}
379 < \caption{Density versus temperature for TIP4P [Ref. \citen{Jorgensen98b}],
380 < TIP3P [Ref. \citen{Jorgensen98b}], SPC/E [Ref. \citen{Clancy94}], {\sc ssd}
381 < without Reaction Field, {\sc ssd}, and experiment [Ref. \citen{CRC80}]. The
382 < arrows indicate the change in densities observed when turning off the
383 < reaction field. The the lower than expected densities for the {\sc ssd}
384 < model were what prompted the original reparameterization of {\sc ssd1}
385 < [Ref. \citen{Ichiye03}].}
386 < \label{dense1}
387 < \end{center}
388 < \end{figure}
375 > %\begin{figure}
376 > %\begin{center}
377 > %\epsfxsize=6in
378 > %\epsfbox{denseSSDnew.eps}
379 > %\caption{Density versus temperature for TIP4P [Ref. \onlinecite{Jorgensen98b}],
380 > % TIP3P [Ref. \onlinecite{Jorgensen98b}], SPC/E [Ref. \onlinecite{Clancy94}], SSD
381 > % without Reaction Field, SSD, and experiment [Ref. \onlinecite{CRC80}]. The
382 > % arrows indicate the change in densities observed when turning off the
383 > % reaction field. The the lower than expected densities for the SSD
384 > % model were what prompted the original reparameterization of SSD1
385 > % [Ref. \onlinecite{Ichiye03}].}
386 > %\label{dense1}
387 > %\end{center}
388 > %\end{figure}
389  
390 < The density maximum for {\sc ssd} compares quite favorably to other simple
391 < water models. Figure \ref{dense1} also shows calculated densities of
392 < several other models and experiment obtained from other
390 > The density maximum for SSD compares quite favorably to other
391 > simple water models. Figure \ref{dense1} also shows calculated
392 > densities of several other models and experiment obtained from other
393   sources.\cite{Jorgensen98b,Clancy94,CRC80} Of the listed simple water
394 < models, {\sc ssd} has a temperature closest to the experimentally observed
395 < density maximum. Of the {\it charge-based} models in
394 > models, SSD has a temperature closest to the experimentally
395 > observed density maximum. Of the {\it charge-based} models in
396   Fig. \ref{dense1}, TIP4P has a density maximum behavior most like that
397 < seen in {\sc ssd}. Though not included in this plot, it is useful
398 < to note that TIP5P has a density maximum nearly identical to the
397 > seen in SSD. Though not included in this plot, it is useful to
398 > note that TIP5P has a density maximum nearly identical to the
399   experimentally measured temperature.
400  
401   It has been observed that liquid state densities in water are
402   dependent on the cutoff radius used both with and without the use of
403   reaction field.\cite{Berendsen98} In order to address the possible
404   effect of cutoff radius, simulations were performed with a dipolar
405 < cutoff radius of 12.0 \AA\ to complement the previous {\sc ssd} simulations,
406 < all performed with a cutoff of 9.0 \AA. All of the resulting densities
407 < overlapped within error and showed no significant trend toward lower
408 < or higher densities as a function of cutoff radius, for simulations
409 < both with and without reaction field. These results indicate that
410 < there is no major benefit in choosing a longer cutoff radius in
411 < simulations using {\sc ssd}. This is advantageous in that the use of a
412 < longer cutoff radius results in a significant increase in the time
413 < required to obtain a single trajectory.
405 > cutoff radius of 12.0~\AA\ to complement the previous SSD
406 > simulations, all performed with a cutoff of 9.0~\AA. All of the
407 > resulting densities overlapped within error and showed no significant
408 > trend toward lower or higher densities as a function of cutoff radius,
409 > for simulations both with and without reaction field. These results
410 > indicate that there is no major benefit in choosing a longer cutoff
411 > radius in simulations using SSD. This is advantageous in that
412 > the use of a longer cutoff radius results in a significant increase in
413 > the time required to obtain a single trajectory.
414  
415   The key feature to recognize in figure \ref{dense1} is the density
416 < scaling of {\sc ssd} relative to other common models at any given
417 < temperature. {\sc ssd} assumes a lower density than any of the other listed
418 < models at the same pressure, behavior which is especially apparent at
419 < temperatures greater than 300 K. Lower than expected densities have
420 < been observed for other systems using a reaction field for long-range
421 < electrostatic interactions, so the most likely reason for the
422 < significantly lower densities seen in these simulations is the
416 > scaling of SSD relative to other common models at any given
417 > temperature. SSD assumes a lower density than any of the other
418 > listed models at the same pressure, behavior which is especially
419 > apparent at temperatures greater than 300~K. Lower than expected
420 > densities have been observed for other systems using a reaction field
421 > for long-range electrostatic interactions, so the most likely reason
422 > for the significantly lower densities seen in these simulations is the
423   presence of the reaction field.\cite{Berendsen98,Nezbeda02} In order
424   to test the effect of the reaction field on the density of the
425   systems, the simulations were repeated without a reaction field
# Line 422 | Line 427 | the curve produced from {\sc ssd} simulations using re
427   \ref{dense1}. Without the reaction field, the densities increase
428   to more experimentally reasonable values, especially around the
429   freezing point of liquid water. The shape of the curve is similar to
430 < the curve produced from {\sc ssd} simulations using reaction field,
430 > the curve produced from SSD simulations using reaction field,
431   specifically the rapidly decreasing densities at higher temperatures;
432 < however, a shift in the density maximum location, down to 245 K, is
432 > however, a shift in the density maximum location, down to 245~K, is
433   observed. This is a more accurate comparison to the other listed water
434   models, in that no long range corrections were applied in those
435   simulations.\cite{Clancy94,Jorgensen98b} However, even without the
436 < reaction field, the density around 300 K is still significantly lower
436 > reaction field, the density around 300~K is still significantly lower
437   than experiment and comparable water models. This anomalous behavior
438   was what lead Tan {\it et al.} to recently reparameterize
439 < {\sc ssd}.\cite{Ichiye03} Throughout the remainder of the paper our
440 < reparamaterizations of {\sc ssd} will be compared with their newer {\sc ssd1}
439 > SSD.\cite{Ichiye03} Throughout the remainder of the paper our
440 > reparamaterizations of SSD will be compared with their newer SSD1
441   model.
442  
443   \subsection{Transport Behavior}
# Line 443 | Line 448 | underwent 50 ps of temperature scaling and 50 ps of co
448   constant energy (NVE) simulations were performed at the average
449   density obtained by the NPT simulations at an identical target
450   temperature. Simulations started with randomized velocities and
451 < underwent 50 ps of temperature scaling and 50 ps of constant energy
452 < equilibration before a 200 ps data collection run. Diffusion constants
451 > underwent 50~ps of temperature scaling and 50~ps of constant energy
452 > equilibration before a 200~ps data collection run. Diffusion constants
453   were calculated via linear fits to the long-time behavior of the
454   mean-square displacement as a function of time. The averaged results
455   from five sets of NVE simulations are displayed in figure
456   \ref{diffuse}, alongside experimental, SPC/E, and TIP5P
457   results.\cite{Gillen72,Holz00,Clancy94,Jorgensen01}
458  
459 < \begin{figure}
460 < \begin{center}
461 < \epsfxsize=6in
462 < \epsfbox{betterDiffuse.epsi}
463 < \caption{Average self-diffusion constant as a function of temperature for
464 < {\sc ssd}, SPC/E [Ref. \citen{Clancy94}], and TIP5P
465 < [Ref. \citen{Jorgensen01}] compared with experimental data
466 < [Refs. \citen{Gillen72} and \citen{Holz00}]. Of the three water models
467 < shown, {\sc ssd} has the least deviation from the experimental values. The
468 < rapidly increasing diffusion constants for TIP5P and {\sc ssd} correspond to
469 < significant decreases in density at the higher temperatures.}
470 < \label{diffuse}
471 < \end{center}
472 < \end{figure}
459 > %\begin{figure}
460 > %\begin{center}
461 > %\epsfxsize=6in
462 > %\epsfbox{betterDiffuse.epsi}
463 > %\caption{Average self-diffusion constant as a function of temperature for
464 > %SSD, SPC/E [Ref. \onlinecite{Clancy94}], and TIP5P
465 > %[Ref. \onlinecite{Jorgensen01}] compared with experimental data
466 > %[Refs. \onlinecite{Gillen72} and \onlinecite{Holz00}]. Of the three water models
467 > %shown, SSD has the least deviation from the experimental values. The
468 > %rapidly increasing diffusion constants for TIP5P and SSD correspond to
469 > %significant decreases in density at the higher temperatures.}
470 > %\label{diffuse}
471 > %\end{center}
472 > %\end{figure}
473  
474   The observed values for the diffusion constant point out one of the
475 < strengths of the {\sc ssd} model. Of the three models shown, the {\sc ssd} model
475 > strengths of the SSD model. Of the three models shown, the SSD model
476   has the most accurate depiction of self-diffusion in both the
477   supercooled and liquid regimes.  SPC/E does a respectable job by
478 < reproducing values similar to experiment around 290 K; however, it
478 > reproducing values similar to experiment around 290~K; however, it
479   deviates at both higher and lower temperatures, failing to predict the
480 < correct thermal trend. TIP5P and {\sc ssd} both start off low at colder
480 > correct thermal trend. TIP5P and SSD both start off low at colder
481   temperatures and tend to diffuse too rapidly at higher temperatures.
482   This behavior at higher temperatures is not particularly surprising
483 < since the densities of both TIP5P and {\sc ssd} are lower than experimental
483 > since the densities of both TIP5P and SSD are lower than experimental
484   water densities at higher temperatures.  When calculating the
485 < diffusion coefficients for {\sc ssd} at experimental densities (instead of
486 < the densities from the NPT simulations), the resulting values fall
487 < more in line with experiment at these temperatures.
485 > diffusion coefficients for SSD at experimental densities
486 > (instead of the densities from the NPT simulations), the resulting
487 > values fall more in line with experiment at these temperatures.
488  
489   \subsection{Structural Changes and Characterization}
490  
# Line 489 | Line 494 | at 245 K, indicating a first order phase transition fo
494   capacity (C$_\text{p}$) was monitored to locate the melting transition
495   in each of the simulations. In the melting simulations of the 1024
496   particle ice $I_h$ simulations, a large spike in C$_\text{p}$ occurs
497 < at 245 K, indicating a first order phase transition for the melting of
497 > at 245~K, indicating a first order phase transition for the melting of
498   these ice crystals. When the reaction field is turned off, the melting
499 < transition occurs at 235 K.  These melting transitions are
499 > transition occurs at 235~K.  These melting transitions are
500   considerably lower than the experimental value.
501  
502 < \begin{figure}
503 < \begin{center}
504 < \epsfxsize=6in
505 < \epsfbox{corrDiag.eps}
506 < \caption{An illustration of angles involved in the correlations observed in Fig. \ref{contour}.}
507 < \label{corrAngle}
508 < \end{center}
509 < \end{figure}
502 > %\begin{figure}
503 > %\begin{center}
504 > %\epsfxsize=6in
505 > %\epsfbox{corrDiag.eps}
506 > %\caption{An illustration of angles involved in the correlations observed in Fig. \ref{contour}.}
507 > %\label{corrAngle}
508 > %\end{center}
509 > %\end{figure}
510  
511 < \begin{figure}
512 < \begin{center}
513 < \epsfxsize=6in
514 < \epsfbox{fullContours.eps}
515 < \caption{Contour plots of 2D angular pair correlation functions for
516 < 512 {\sc ssd} molecules at 100 K (A \& B) and 300 K (C \& D). Dark areas
517 < signify regions of enhanced density while light areas signify
518 < depletion relative to the bulk density. White areas have pair
519 < correlation values below 0.5 and black areas have values above 1.5.}
520 < \label{contour}
521 < \end{center}
522 < \end{figure}
511 > %\begin{figure}
512 > %\begin{center}
513 > %\epsfxsize=6in
514 > %\epsfbox{fullContours.eps}
515 > %\caption{Contour plots of 2D angular pair correlation functions for
516 > %512 SSD molecules at 100~K (A \& B) and 300~K (C \& D). Dark areas
517 > %signify regions of enhanced density while light areas signify
518 > %depletion relative to the bulk density. White areas have pair
519 > %correlation values below 0.5 and black areas have values above 1.5.}
520 > %\label{contour}
521 > %\end{center}
522 > %\end{figure}
523  
524   Additional analysis of the melting process was performed using
525   two-dimensional structure and dipole angle correlations. Expressions
# Line 554 | Line 559 | this split character alters to show the leading 4 \AA\
559   $g_\mathrm{OO}(r)$.\cite{Ichiye96} At low temperatures, the second
560   solvation shell peak appears to have two distinct components that
561   blend together to form one observable peak. At higher temperatures,
562 < this split character alters to show the leading 4 \AA\ peak dominated
562 > this split character alters to show the leading 4~\AA\ peak dominated
563   by equatorial anti-parallel dipole orientations. There is also a
564   tightly bunched group of axially arranged dipoles that most likely
565   consist of the smaller fraction of aligned dipole pairs. The trailing
566 < component of the split peak at 5 \AA\ is dominated by aligned dipoles
566 > component of the split peak at 5~\AA\ is dominated by aligned dipoles
567   that assume hydrogen bond arrangements similar to those seen in the
568   first solvation shell. This evidence indicates that the dipole pair
569   interaction begins to dominate outside of the range of the dipolar
570   repulsion term.  The energetically favorable dipole arrangements
571   populate the region immediately outside this repulsion region (around
572 < 4 \AA), while arrangements that seek to satisfy both the sticky and
572 > 4~\AA), while arrangements that seek to satisfy both the sticky and
573   dipole forces locate themselves just beyond this initial buildup
574 < (around 5 \AA).
574 > (around 5~\AA).
575  
576   From these findings, the split second peak is primarily the product of
577   the dipolar repulsion term of the sticky potential. In fact, the inner
578   peak can be pushed out and merged with the outer split peak just by
579   extending the switching function ($s^\prime(r_{ij})$) from its normal
580 < 4.0 \AA\ cutoff to values of 4.5 or even 5 \AA. This type of
580 > 4.0~\AA\ cutoff to values of 4.5 or even 5~\AA. This type of
581   correction is not recommended for improving the liquid structure,
582   since the second solvation shell would still be shifted too far
583   out. In addition, this would have an even more detrimental effect on
584   the system densities, leading to a liquid with a more open structure
585 < and a density considerably lower than the already low {\sc ssd} density.  A
586 < better correction would be to include the quadrupole-quadrupole
587 < interactions for the water particles outside of the first solvation
588 < shell, but this would remove the simplicity and speed advantage of
589 < {\sc ssd}.
585 > and a density considerably lower than the already low SSD
586 > density.  A better correction would be to include the
587 > quadrupole-quadrupole interactions for the water particles outside of
588 > the first solvation shell, but this would remove the simplicity and
589 > speed advantage of SSD.
590  
591 < \subsection{Adjusted Potentials: {\sc ssd/rf} and {\sc ssd/e}}
591 > \subsection{Adjusted Potentials: SSD/RF and SSD/E}
592  
593 < The propensity of {\sc ssd} to adopt lower than expected densities under
593 > The propensity of SSD to adopt lower than expected densities under
594   varying conditions is troubling, especially at higher temperatures. In
595   order to correct this model for use with a reaction field, it is
596   necessary to adjust the force field parameters for the primary
# Line 601 | Line 606 | Sticky Dipole / Reaction Field ({\sc ssd/rf} - for use
606   function cutoffs ($r_l$, $r_u$ and $r_l^\prime$, $r_u^\prime$
607   respectively). The results of the reparameterizations are shown in
608   table \ref{params}. We are calling these reparameterizations the Soft
609 < Sticky Dipole / Reaction Field ({\sc ssd/rf} - for use with a reaction
610 < field) and Soft Sticky Dipole Extended ({\sc ssd/e} - an attempt to improve
609 > Sticky Dipole / Reaction Field (SSD/RF - for use with a reaction
610 > field) and Soft Sticky Dipole Extended (SSD/E - an attempt to improve
611   the liquid structure in simulations without a long-range correction).
612  
613   \begin{table}
# Line 610 | Line 615 | the liquid structure in simulations without a long-ran
615   \caption{Parameters for the original and adjusted models}
616   \begin{tabular}{ l  c  c  c  c }
617   \hline \\[-3mm]
618 < \ \ \ Parameters\ \ \  & \ \ \ {\sc ssd} [Ref. \citen{Ichiye96}] \ \ \
619 < & \ {\sc ssd1} [Ref. \citen{Ichiye03}]\ \  & \ {\sc ssd/e}\ \  & \ {\sc ssd/rf} \\
618 > \ \ \ Parameters\ \ \  & \ \ \ SSD [Ref. \onlinecite{Ichiye96}] \ \ \
619 > & \ SSD1 [Ref. \onlinecite{Ichiye03}]\ \  & \ SSD/E\ \  & \ \ SSD/RF \\
620   \hline \\[-3mm]
621   \ \ \ $\sigma$ (\AA)  & 3.051 & 3.016 & 3.035 & 3.019\\
622   \ \ \ $\epsilon$ (kcal/mol) & 0.152 & 0.152 & 0.152 & 0.152\\
# Line 627 | Line 632 | the liquid structure in simulations without a long-ran
632   \end{center}
633   \end{table}
634  
635 < \begin{figure}
636 < \begin{center}
637 < \epsfxsize=5in
638 < \epsfbox{GofRCompare.epsi}
639 < \caption{Plots comparing experiment [Ref. \citen{Head-Gordon00_1}] with {\sc ssd/e}
640 < and {\sc ssd1} without reaction field (top), as well as {\sc ssd/rf} and {\sc ssd1} with
641 < reaction field turned on (bottom). The insets show the respective
642 < first peaks in detail. Note how the changes in parameters have lowered
643 < and broadened the first peak of {\sc ssd/e} and {\sc ssd/rf}.}
644 < \label{grcompare}
645 < \end{center}
646 < \end{figure}
635 > %\begin{figure}
636 > %\begin{center}
637 > %\epsfxsize=5in
638 > %\epsfbox{GofRCompare.epsi}
639 > %\caption{Plots comparing experiment [Ref. \onlinecite{Head-Gordon00_1}] with
640 > %SSD/E and SSD1 without reaction field (top), as well as
641 > %SSD/RF and SSD1 with reaction field turned on
642 > %(bottom). The insets show the respective first peaks in detail. Note
643 > %how the changes in parameters have lowered and broadened the first
644 > %peak of SSD/E and SSD/RF.}
645 > %\label{grcompare}
646 > %\end{center}
647 > %\end{figure}
648  
649 < \begin{figure}
650 < \begin{center}
651 < \epsfxsize=6in
652 < \epsfbox{dualsticky_bw.eps}
653 < \caption{Positive and negative isosurfaces of the sticky potential for
654 < {\sc ssd1} (left) and {\sc ssd/e} \& {\sc ssd/rf} (right). Light areas correspond to the
655 < tetrahedral attractive component, and darker areas correspond to the
656 < dipolar repulsive component.}
657 < \label{isosurface}
658 < \end{center}
659 < \end{figure}
649 > %\begin{figure}
650 > %\begin{center}
651 > %\epsfxsize=6in
652 > %\epsfbox{dualsticky_bw.eps}
653 > %\caption{Positive and negative isosurfaces of the sticky potential for
654 > %SSD1 (left) and SSD/E \& SSD/RF (right). Light areas
655 > %correspond to the tetrahedral attractive component, and darker areas
656 > %correspond to the dipolar repulsive component.}
657 > %\label{isosurface}
658 > %\end{center}
659 > %\end{figure}
660  
661 < In the original paper detailing the development of {\sc ssd}, Liu and Ichiye
661 > In the original paper detailing the development of SSD, Liu and Ichiye
662   placed particular emphasis on an accurate description of the first
663   solvation shell. This resulted in a somewhat tall and narrow first
664   peak in $g(r)$ that integrated to give similar coordination numbers to
665   the experimental data obtained by Soper and
666   Phillips.\cite{Ichiye96,Soper86} New experimental x-ray scattering
667   data from the Head-Gordon lab indicates a slightly lower and shifted
668 < first peak in the g$_\mathrm{OO}(r)$, so our adjustments to {\sc ssd} were
668 > first peak in the g$_\mathrm{OO}(r)$, so our adjustments to SSD were
669   made after taking into consideration the new experimental
670   findings.\cite{Head-Gordon00_1} Figure \ref{grcompare} shows the
671   relocation of the first peak of the oxygen-oxygen $g(r)$ by comparing
672 < the revised {\sc ssd} model ({\sc ssd1}), {\sc ssd/e}, and {\sc ssd/rf} to the new
672 > the revised SSD model (SSD1), SSD/E, and SSD/RF to the new
673   experimental results. Both modified water models have shorter peaks
674   that match more closely to the experimental peak (as seen in the
675   insets of figure \ref{grcompare}).  This structural alteration was
# Line 676 | Line 682 | density for the overall system.  This change in intera
682   see how altering the cutoffs changes the repulsive and attractive
683   character of the particles. With a reduced repulsive surface (darker
684   region), the particles can move closer to one another, increasing the
685 < density for the overall system.  This change in interaction cutoff also
686 < results in a more gradual orientational motion by allowing the
685 > density for the overall system.  This change in interaction cutoff
686 > also results in a more gradual orientational motion by allowing the
687   particles to maintain preferred dipolar arrangements before they begin
688   to feel the pull of the tetrahedral restructuring. As the particles
689   move closer together, the dipolar repulsion term becomes active and
690   excludes unphysical nearest-neighbor arrangements. This compares with
691 < how {\sc ssd} and {\sc ssd1} exclude preferred dipole alignments before the
691 > how SSD and SSD1 exclude preferred dipole alignments before the
692   particles feel the pull of the ``hydrogen bonds''. Aside from
693   improving the shape of the first peak in the g(\emph{r}), this
694   modification improves the densities considerably by allowing the
695 < persistence of full dipolar character below the previous 4.0 \AA\
695 > persistence of full dipolar character below the previous 4.0~\AA\
696   cutoff.
697  
698   While adjusting the location and shape of the first peak of $g(r)$
699   improves the densities, these changes alone are insufficient to bring
700   the system densities up to the values observed experimentally.  To
701   further increase the densities, the dipole moments were increased in
702 < both of our adjusted models. Since {\sc ssd} is a dipole based model, the
702 > both of our adjusted models. Since SSD is a dipole based model, the
703   structure and transport are very sensitive to changes in the dipole
704 < moment. The original {\sc ssd} simply used the dipole moment calculated from
705 < the TIP3P water model, which at 2.35 D is significantly greater than
706 < the experimental gas phase value of 1.84 D. The larger dipole moment
704 > moment. The original SSD simply used the dipole moment calculated from
705 > the TIP3P water model, which at 2.35~D is significantly greater than
706 > the experimental gas phase value of 1.84~D. The larger dipole moment
707   is a more realistic value and improves the dielectric properties of
708   the fluid. Both theoretical and experimental measurements indicate a
709 < liquid phase dipole moment ranging from 2.4 D to values as high as
710 < 3.11 D, providing a substantial range of reasonable values for a
709 > liquid phase dipole moment ranging from 2.4~D to values as high as
710 > 3.11~D, providing a substantial range of reasonable values for a
711   dipole moment.\cite{Sprik91,Kusalik02,Badyal00,Barriol64} Moderately
712 < increasing the dipole moments to 2.42 and 2.48 D for {\sc ssd/e} and {\sc ssd/rf},
712 > increasing the dipole moments to 2.42 and 2.48~D for SSD/E and SSD/RF,
713   respectively, leads to significant changes in the density and
714   transport of the water models.
715  
716   In order to demonstrate the benefits of these reparameterizations, a
717   series of NPT and NVE simulations were performed to probe the density
718   and transport properties of the adapted models and compare the results
719 < to the original {\sc ssd} model. This comparison involved full NPT melting
720 < sequences for both {\sc ssd/e} and {\sc ssd/rf}, as well as NVE transport
719 > to the original SSD model. This comparison involved full NPT melting
720 > sequences for both SSD/E and SSD/RF, as well as NVE transport
721   calculations at the calculated self-consistent densities. Again, the
722   results are obtained from five separate simulations of 1024 particle
723   systems, and the melting sequences were started from different ice
724   $I_h$ crystals constructed as described previously. Each NPT
725 < simulation was equilibrated for 100 ps before a 200 ps data collection
725 > simulation was equilibrated for 100~ps before a 200~ps data collection
726   run at each temperature step, and the final configuration from the
727   previous temperature simulation was used as a starting point. All NVE
728   simulations had the same thermalization, equilibration, and data
729   collection times as stated previously.
730  
731 < \begin{figure}
732 < \begin{center}
733 < \epsfxsize=6in
734 < \epsfbox{ssdeDense.epsi}
735 < \caption{Comparison of densities calculated with {\sc ssd/e} to {\sc ssd1} without a
736 < reaction field, TIP3P [Ref. \citen{Jorgensen98b}], TIP5P
737 < [Ref. \citen{Jorgensen00}], SPC/E [Ref. \citen{Clancy94}] and
738 < experiment [Ref. \citen{CRC80}]. The window shows a expansion around
739 < 300 K with error bars included to clarify this region of
740 < interest. Note that both {\sc ssd1} and {\sc ssd/e} show good agreement with
741 < experiment when the long-range correction is neglected.}
742 < \label{ssdedense}
743 < \end{center}
744 < \end{figure}
731 > %\begin{figure}
732 > %\begin{center}
733 > %\epsfxsize=6in
734 > %\epsfbox{ssdeDense.epsi}
735 > %\caption{Comparison of densities calculated with SSD/E to
736 > %SSD1 without a reaction field, TIP3P [Ref. \onlinecite{Jorgensen98b}],
737 > %TIP5P [Ref. \onlinecite{Jorgensen00}], SPC/E [Ref. \onlinecite{Clancy94}] and
738 > %experiment [Ref. \onlinecite{CRC80}]. The window shows a expansion around
739 > %300 K with error bars included to clarify this region of
740 > %interest. Note that both SSD1 and SSD/E show good agreement with
741 > %experiment when the long-range correction is neglected.}
742 > %\label{ssdedense}
743 > %\end{center}
744 > %\end{figure}
745  
746 < Fig. \ref{ssdedense} shows the density profile for the {\sc ssd/e} model
747 < in comparison to {\sc ssd1} without a reaction field, other common water
748 < models, and experimental results. The calculated densities for both
749 < {\sc ssd/e} and {\sc ssd1} have increased significantly over the original {\sc ssd}
750 < model (see fig. \ref{dense1}) and are in better agreement with the
751 < experimental values. At 298 K, the densities of {\sc ssd/e} and {\sc ssd1} without
746 > Fig. \ref{ssdedense} shows the density profile for the SSD/E
747 > model in comparison to SSD1 without a reaction field, other
748 > common water models, and experimental results. The calculated
749 > densities for both SSD/E and SSD1 have increased
750 > significantly over the original SSD model (see
751 > fig. \ref{dense1}) and are in better agreement with the experimental
752 > values. At 298 K, the densities of SSD/E and SSD1 without
753   a long-range correction are 0.996$\pm$0.001 g/cm$^3$ and
754   0.999$\pm$0.001 g/cm$^3$ respectively.  These both compare well with
755   the experimental value of 0.997 g/cm$^3$, and they are considerably
756 < better than the {\sc ssd} value of 0.967$\pm$0.003 g/cm$^3$. The changes to
757 < the dipole moment and sticky switching functions have improved the
758 < structuring of the liquid (as seen in figure \ref{grcompare}, but they
759 < have shifted the density maximum to much lower temperatures. This
760 < comes about via an increase in the liquid disorder through the
761 < weakening of the sticky potential and strengthening of the dipolar
762 < character. However, this increasing disorder in the {\sc ssd/e} model has
763 < little effect on the melting transition. By monitoring $C_p$
764 < throughout these simulations, the melting transition for {\sc ssd/e} was
765 < shown to occur at 235 K.  The same transition temperature observed
766 < with {\sc ssd} and {\sc ssd1}.
756 > better than the SSD value of 0.967$\pm$0.003 g/cm$^3$. The
757 > changes to the dipole moment and sticky switching functions have
758 > improved the structuring of the liquid (as seen in figure
759 > \ref{grcompare}, but they have shifted the density maximum to much
760 > lower temperatures. This comes about via an increase in the liquid
761 > disorder through the weakening of the sticky potential and
762 > strengthening of the dipolar character. However, this increasing
763 > disorder in the SSD/E model has little effect on the melting
764 > transition. By monitoring $C_p$ throughout these simulations, the
765 > melting transition for SSD/E was shown to occur at 235~K.  The
766 > same transition temperature observed with SSD and SSD1.
767  
768 < \begin{figure}
769 < \begin{center}
770 < \epsfxsize=6in
771 < \epsfbox{ssdrfDense.epsi}
772 < \caption{Comparison of densities calculated with {\sc ssd/rf} to {\sc ssd1} with a
773 < reaction field, TIP3P [Ref. \citen{Jorgensen98b}], TIP5P
774 < [Ref. \citen{Jorgensen00}], SPC/E [Ref. \citen{Clancy94}], and
775 < experiment [Ref. \citen{CRC80}]. The inset shows the necessity of
776 < reparameterization when utilizing a reaction field long-ranged
777 < correction - {\sc ssd/rf} provides significantly more accurate densities
778 < than {\sc ssd1} when performing room temperature simulations.}
779 < \label{ssdrfdense}
780 < \end{center}
781 < \end{figure}
768 > %\begin{figure}
769 > %\begin{center}
770 > %\epsfxsize=6in
771 > %\epsfbox{ssdrfDense.epsi}
772 > %\caption{Comparison of densities calculated with SSD/RF to
773 > %SSD1 with a reaction field, TIP3P [Ref. \onlinecite{Jorgensen98b}],
774 > %TIP5P [Ref. \onlinecite{Jorgensen00}], SPC/E [Ref. \onlinecite{Clancy94}], and
775 > %experiment [Ref. \onlinecite{CRC80}]. The inset shows the necessity of
776 > %reparameterization when utilizing a reaction field long-ranged
777 > %correction - SSD/RF provides significantly more accurate
778 > %densities than SSD1 when performing room temperature
779 > %simulations.}
780 > %\label{ssdrfdense}
781 > %\end{center}
782 > %\end{figure}
783  
784   Including the reaction field long-range correction in the simulations
785   results in a more interesting comparison.  A density profile including
786 < {\sc ssd/rf} and {\sc ssd1} with an active reaction field is shown in figure
786 > SSD/RF and SSD1 with an active reaction field is shown in figure
787   \ref{ssdrfdense}.  As observed in the simulations without a reaction
788 < field, the densities of {\sc ssd/rf} and {\sc ssd1} show a dramatic increase over
789 < normal {\sc ssd} (see figure \ref{dense1}). At 298 K, {\sc ssd/rf} has a density
788 > field, the densities of SSD/RF and SSD1 show a dramatic increase over
789 > normal SSD (see figure \ref{dense1}). At 298 K, SSD/RF has a density
790   of 0.997$\pm$0.001 g/cm$^3$, directly in line with experiment and
791 < considerably better than the original {\sc ssd} value of 0.941$\pm$0.001
792 < g/cm$^3$ and the {\sc ssd1} value of 0.972$\pm$0.002 g/cm$^3$. These results
791 > considerably better than the original SSD value of 0.941$\pm$0.001
792 > g/cm$^3$ and the SSD1 value of 0.972$\pm$0.002 g/cm$^3$. These results
793   further emphasize the importance of reparameterization in order to
794   model the density properly under different simulation conditions.
795   Again, these changes have only a minor effect on the melting point,
796 < which observed at 245 K for {\sc ssd/rf}, is identical to {\sc ssd} and only 5 K
797 < lower than {\sc ssd1} with a reaction field. Additionally, the difference in
798 < density maxima is not as extreme, with {\sc ssd/rf} showing a density
799 < maximum at 255 K, fairly close to the density maxima of 260 K and 265
800 < K, shown by {\sc ssd} and {\sc ssd1} respectively.
796 > which observed at 245~K for SSD/RF, is identical to SSD and only 5~K
797 > lower than SSD1 with a reaction field. Additionally, the difference in
798 > density maxima is not as extreme, with SSD/RF showing a density
799 > maximum at 255~K, fairly close to the density maxima of 260~K and
800 > 265~K, shown by SSD and SSD1 respectively.
801  
802 < \begin{figure}
803 < \begin{center}
804 < \epsfxsize=6in
805 < \epsfbox{ssdeDiffuse.epsi}
806 < \caption{The diffusion constants calculated from {\sc ssd/e} and {\sc ssd1} (both
807 < without a reaction field) along with experimental results
808 < [Refs. \citen{Gillen72} and \citen{Holz00}]. The NVE calculations were
809 < performed at the average densities observed in the 1 atm NPT
810 < simulations for the respective models. {\sc ssd/e} is slightly more mobile
811 < than experiment at all of the temperatures, but it is closer to
812 < experiment at biologically relevant temperatures than {\sc ssd1} without a
813 < long-range correction.}
814 < \label{ssdediffuse}
815 < \end{center}
816 < \end{figure}
802 > %\begin{figure}
803 > %\begin{center}
804 > %\epsfxsize=6in
805 > %\epsfbox{ssdeDiffuse.epsi}
806 > %\caption{The diffusion constants calculated from SSD/E and
807 > %SSD1 (both without a reaction field) along with experimental results
808 > %[Refs. \onlinecite{Gillen72} and \onlinecite{Holz00}]. The NVE calculations were
809 > %performed at the average densities observed in the 1 atm NPT
810 > %simulations for the respective models. SSD/E is slightly more mobile
811 > %than experiment at all of the temperatures, but it is closer to
812 > %experiment at biologically relevant temperatures than SSD1 without a
813 > %long-range correction.}
814 > %\label{ssdediffuse}
815 > %\end{center}
816 > %\end{figure}
817  
818 < The reparameterization of the {\sc ssd} water model, both for use with and
818 > The reparameterization of the SSD water model, both for use with and
819   without an applied long-range correction, brought the densities up to
820   what is expected for simulating liquid water. In addition to improving
821 < the densities, it is important that the diffusive behavior of {\sc ssd} be
821 > the densities, it is important that the diffusive behavior of SSD be
822   maintained or improved. Figure \ref{ssdediffuse} compares the
823 < temperature dependence of the diffusion constant of {\sc ssd/e} to {\sc ssd1}
823 > temperature dependence of the diffusion constant of SSD/E to SSD1
824   without an active reaction field at the densities calculated from
825   their respective NPT simulations at 1 atm. The diffusion constant for
826 < {\sc ssd/e} is consistently higher than experiment, while {\sc ssd1} remains lower
826 > SSD/E is consistently higher than experiment, while SSD1 remains lower
827   than experiment until relatively high temperatures (around 360
828   K). Both models follow the shape of the experimental curve well below
829 < 300 K but tend to diffuse too rapidly at higher temperatures, as seen
830 < in {\sc ssd1}'s crossing above 360 K.  This increasing diffusion relative to
829 > 300~K but tend to diffuse too rapidly at higher temperatures, as seen
830 > in SSD1's crossing above 360~K.  This increasing diffusion relative to
831   the experimental values is caused by the rapidly decreasing system
832 < density with increasing temperature.  Both {\sc ssd1} and {\sc ssd/e} show this
832 > density with increasing temperature.  Both SSD1 and SSD/E show this
833   deviation in particle mobility, but this trend has different
834 < implications on the diffusive behavior of the models.  While {\sc ssd1}
834 > implications on the diffusive behavior of the models.  While SSD1
835   shows more experimentally accurate diffusive behavior in the high
836 < temperature regimes, {\sc ssd/e} shows more accurate behavior in the
836 > temperature regimes, SSD/E shows more accurate behavior in the
837   supercooled and biologically relevant temperature ranges.  Thus, the
838   changes made to improve the liquid structure may have had an adverse
839   affect on the density maximum, but they improve the transport behavior
840 < of {\sc ssd/e} relative to {\sc ssd1} under the most commonly simulated
840 > of SSD/E relative to SSD1 under the most commonly simulated
841   conditions.
842  
843 < \begin{figure}
844 < \begin{center}
845 < \epsfxsize=6in
846 < \epsfbox{ssdrfDiffuse.epsi}
847 < \caption{The diffusion constants calculated from {\sc ssd/rf} and {\sc ssd1} (both
848 < with an active reaction field) along with experimental results
849 < [Refs. \citen{Gillen72} and \citen{Holz00}]. The NVE calculations were
850 < performed at the average densities observed in the 1 atm NPT
851 < simulations for both of the models. {\sc ssd/rf} simulates the diffusion of
852 < water throughout this temperature range very well. The rapidly
853 < increasing diffusion constants at high temperatures for both models
854 < can be attributed to lower calculated densities than those observed in
855 < experiment.}
856 < \label{ssdrfdiffuse}
857 < \end{center}
858 < \end{figure}
843 > %\begin{figure}
844 > %\begin{center}
845 > %\epsfxsize=6in
846 > %\epsfbox{ssdrfDiffuse.epsi}
847 > %\caption{The diffusion constants calculated from SSD/RF and
848 > %SSD1 (both with an active reaction field) along with
849 > %experimental results [Refs. \onlinecite{Gillen72} and \onlinecite{Holz00}]. The
850 > %NVE calculations were performed at the average densities observed in
851 > %the 1 atm NPT simulations for both of the models. SSD/RF
852 > %simulates the diffusion of water throughout this temperature range
853 > %very well. The rapidly increasing diffusion constants at high
854 > %temperatures for both models can be attributed to lower calculated
855 > %densities than those observed in experiment.}
856 > %\label{ssdrfdiffuse}
857 > %\end{center}
858 > %\end{figure}
859  
860 < In figure \ref{ssdrfdiffuse}, the diffusion constants for {\sc ssd/rf} are
861 < compared to {\sc ssd1} with an active reaction field. Note that {\sc ssd/rf}
860 > In figure \ref{ssdrfdiffuse}, the diffusion constants for SSD/RF are
861 > compared to SSD1 with an active reaction field. Note that SSD/RF
862   tracks the experimental results quantitatively, identical within error
863   throughout most of the temperature range shown and exhibiting only a
864 < slight increasing trend at higher temperatures. {\sc ssd1} tends to diffuse
864 > slight increasing trend at higher temperatures. SSD1 tends to diffuse
865   more slowly at low temperatures and deviates to diffuse too rapidly at
866 < temperatures greater than 330 K.  As stated above, this deviation away
866 > temperatures greater than 330~K.  As stated above, this deviation away
867   from the ideal trend is due to a rapid decrease in density at higher
868 < temperatures. {\sc ssd/rf} does not suffer from this problem as much as {\sc ssd1}
868 > temperatures. SSD/RF does not suffer from this problem as much as SSD1
869   because the calculated densities are closer to the experimental
870   values. These results again emphasize the importance of careful
871   reparameterization when using an altered long-range correction.
# Line 867 | Line 875 | experimental data at ambient conditions}
875   \renewcommand{\thefootnote}{\thempfootnote}
876   \begin{center}
877   \caption{Properties of the single-point water models compared with
878 < experimental data at ambient conditions}
878 > experimental data at ambient conditions. Deviations of the of the
879 > averages are given in parentheses.}
880   \begin{tabular}{ l  c  c  c  c  c }
881   \hline \\[-3mm]
882 < \ \ \ \ \ \  & \ \ \ {\sc ssd1} \ \ \ & \ {\sc ssd/e} \ \ \ & \ {\sc ssd1} (RF) \ \
883 < \ & \ {\sc ssd/rf} \ \ \ & \ Expt. \\
882 > \ \ \ \ \ \  & \ \ \ SSD1 \ \ \ & \ \ SSD/E \ \ \ & \ \ SSD1 (RF) \ \
883 > \ & \ \ SSD/RF \ \ \ & \ \ Expt. \\
884   \hline \\[-3mm]
885 < \ \ \ $\rho$ (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 \\
886 < \ \ \ $C_p$ (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 \\
887 < \ \ \ $D$ ($10^{-5}$ cm$^2$/s) & 1.78 $\pm$0.07 & 2.51 $\pm$0.18 &
888 < 2.00 $\pm$0.17 & 2.32 $\pm$0.06 & 2.299\cite{Mills73} \\
889 < \ \ \ Coordination Number ($n_C$) & 3.9 & 4.3 & 3.8 & 4.4 &
890 < 4.7\footnote{Calculated by integrating $g_{\text{OO}}(r)$ in
891 < Ref. \citen{Head-Gordon00_1}} \\
883 < \ \ \ H-bonds per particle ($n_H$) & 3.7 & 3.6 & 3.7 & 3.7 &
885 > \ \ $\rho$ (g/cm$^3$) & 0.999(0.001) & 0.996(0.001) & 0.972(0.002) & 0.997(0.001) & 0.997 \\
886 > \ \ $C_p$ (cal/mol K) & 28.80(0.11) & 25.45(0.09) & 28.28(0.06) & 23.83(0.16) & 17.98 \\
887 > \ \ $D$ ($10^{-5}$ cm$^2$/s) & 1.78(0.7) & 2.51(0.18) & 2.00(0.17) & 2.32(0.06) & 2.299\cite{Mills73} \\
888 > \ \ Coordination Number ($n_C$) & 3.9 & 4.3 & 3.8 & 4.4 &
889 > 4.7\footnote{Calculated by integrating $g_{\text{OO}}(r)$ in
890 > Ref. \onlinecite{Head-Gordon00_1}} \\
891 > \ \ H-bonds per particle ($n_H$) & 3.7 & 3.6 & 3.7 & 3.7 &
892   3.5\footnote{Calculated by integrating $g_{\text{OH}}(r)$ in
893 < Ref. \citen{Soper86}}  \\
894 < \ \ \ $\tau_1$ (ps) & 10.9 $\pm$0.6 & 7.3 $\pm$0.4 & 7.5 $\pm$0.7 &
895 < 7.2 $\pm$0.4 & 5.7\footnote{Calculated for 298 K from data in Ref. \citen{Eisenberg69}} \\
896 < \ \ \ $\tau_2$ (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 < Ref. \citen{Krynicki66}}
893 > Ref. \onlinecite{Soper86}}  \\
894 > \ \ $\tau_1$ (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. \onlinecite{Eisenberg69}} \\
895 > \ \ $\tau_2$ (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
896 > Ref. \onlinecite{Krynicki66}}
897   \end{tabular}
898   \label{liquidproperties}
899   \end{center}
# Line 909 | Line 915 | but the coordination numbers of {\sc ssd/e} and {\sc s
915   $a$ and $b$ are the radial locations of the minima following the first
916   peak in g$_\text{OO}(r)$ or g$_\text{OH}(r)$ respectively. The number
917   of hydrogen bonds stays relatively constant across all of the models,
918 < but the coordination numbers of {\sc ssd/e} and {\sc ssd/rf} show an improvement
919 < over {\sc ssd1}.  This improvement is primarily due to extension of the
920 < first solvation shell in the new parameter sets.  Because $n_H$ and
921 < $n_C$ are nearly identical in {\sc ssd1}, it appears that all molecules in
922 < the first solvation shell are involved in hydrogen bonds.  Since $n_H$
923 < and $n_C$ differ in the newly parameterized models, the orientations
924 < in the first solvation shell are a bit more ``fluid''.  Therefore {\sc ssd1}
925 < overstructures the first solvation shell and our reparameterizations
926 < have returned this shell to more realistic liquid-like behavior.
918 > but the coordination numbers of SSD/E and SSD/RF show an
919 > improvement over SSD1.  This improvement is primarily due to
920 > extension of the first solvation shell in the new parameter sets.
921 > Because $n_H$ and $n_C$ are nearly identical in SSD1, it appears
922 > that all molecules in the first solvation shell are involved in
923 > hydrogen bonds.  Since $n_H$ and $n_C$ differ in the newly
924 > parameterized models, the orientations in the first solvation shell
925 > are a bit more ``fluid''.  Therefore SSD1 overstructures the
926 > first solvation shell and our reparameterizations have returned this
927 > shell to more realistic liquid-like behavior.
928  
929   The time constants for the orientational autocorrelation functions
930   are also displayed in Table \ref{liquidproperties}. The dipolar
# Line 935 | Line 942 | the NMR data in Ref. \citen{Krynicki66} at a temperatu
942   averaged over five detailed NVE simulations performed at the ambient
943   conditions for each of the respective models. It should be noted that
944   the commonly cited value of 1.9 ps for $\tau_2$ was determined from
945 < the NMR data in Ref. \citen{Krynicki66} at a temperature near
945 > the NMR data in Ref. \onlinecite{Krynicki66} at a temperature near
946   34$^\circ$C.\cite{Rahman71} Because of the strong temperature
947 < dependence of $\tau_2$, it is necessary to recalculate it at 298 K to
947 > dependence of $\tau_2$, it is necessary to recalculate it at 298~K to
948   make proper comparisons. The value shown in Table
949   \ref{liquidproperties} was calculated from the same NMR data in the
950 < fashion described in Ref. \citen{Krynicki66}. Similarly, $\tau_1$ was
951 < recomputed for 298 K from the data in Ref. \citen{Eisenberg69}.
952 < Again, {\sc ssd/e} and {\sc ssd/rf} show improved behavior over {\sc ssd1}, both with
950 > fashion described in Ref. \onlinecite{Krynicki66}. Similarly, $\tau_1$ was
951 > recomputed for 298~K from the data in Ref. \onlinecite{Eisenberg69}.
952 > Again, SSD/E and SSD/RF show improved behavior over SSD1, both with
953   and without an active reaction field. Turning on the reaction field
954 < leads to much improved time constants for {\sc ssd1}; however, these results
954 > leads to much improved time constants for SSD1; however, these results
955   also include a corresponding decrease in system density.
956 < Orientational relaxation times published in the original {\sc ssd} dynamics
956 > Orientational relaxation times published in the original SSD dynamics
957   papers are smaller than the values observed here, and this difference
958   can be attributed to the use of the Ewald sum.\cite{Ichiye99}
959  
960   \subsection{Additional Observations}
961  
962 < \begin{figure}
963 < \begin{center}
964 < \epsfxsize=6in
965 < \epsfbox{icei_bw.eps}
966 < \caption{The most stable crystal structure assumed by the {\sc ssd} family
967 < of water models.  We refer to this structure as Ice-{\it i} to
968 < indicate its origins in computer simulation.  This image was taken of
969 < the (001) face of the crystal.}
970 < \label{weirdice}
971 < \end{center}
972 < \end{figure}
962 > %\begin{figure}
963 > %\begin{center}
964 > %\epsfxsize=6in
965 > %\epsfbox{icei_bw.eps}
966 > %\caption{The most stable crystal structure assumed by the SSD family
967 > %of water models.  We refer to this structure as Ice-{\it i} to
968 > %indicate its origins in computer simulation.  This image was taken of
969 > %the (001) face of the crystal.}
970 > %\label{weirdice}
971 > %\end{center}
972 > %\end{figure}
973  
974   While performing a series of melting simulations on an early iteration
975 < of {\sc ssd/e} not discussed in this paper, we observed recrystallization
976 < into a novel structure not previously known for water.  After melting
977 < at 235 K, two of five systems underwent crystallization events near
978 < 245 K.  The two systems remained crystalline up to 320 and 330 K,
979 < respectively.  The crystal exhibits an expanded zeolite-like structure
980 < that does not correspond to any known form of ice.  This appears to be
981 < an artifact of the point dipolar models, so to distinguish it from the
982 < experimentally observed forms of ice, we have denoted the structure
975 > of SSD/E not discussed in this paper, we observed
976 > recrystallization into a novel structure not previously known for
977 > water.  After melting at 235~K, two of five systems underwent
978 > crystallization events near 245~K.  The two systems remained
979 > crystalline up to 320 and 330~K, respectively.  The crystal exhibits
980 > an expanded zeolite-like structure that does not correspond to any
981 > known form of ice.  This appears to be an artifact of the point
982 > dipolar models, so to distinguish it from the experimentally observed
983 > forms of ice, we have denoted the structure
984   Ice-$\sqrt{\smash[b]{-\text{I}}}$ (Ice-{\it i}).  A large enough
985   portion of the sample crystallized that we have been able to obtain a
986   near ideal crystal structure of Ice-{\it i}. Figure \ref{weirdice}
# Line 987 | Line 995 | structure for at least the {\sc ssd/e} model. To verif
995   favorable enough to stabilize an entire crystal generated around them.
996  
997   Initial simulations indicated that Ice-{\it i} is the preferred ice
998 < structure for at least the {\sc ssd/e} model. To verify this, a
999 < comparison was made between near ideal crystals of ice~$I_h$,
1000 < ice~$I_c$, and Ice-{\it i} at constant pressure with {\sc ssd/e}, {\sc
1001 < ssd/rf}, and {\sc ssd1}. Near-ideal versions of the three types of
1002 < crystals were cooled to 1 K, and enthalpies of formation of each were
1003 < compared using all three water models.  Enthalpies were estimated from
1004 < the isobaric-isothermal simulations using $H=U+P_{\text ext}V$ where
1005 < $P_{\text ext}$ is the applied pressure.  A constant value of
1006 < -60.158 kcal / mol has been added to place our zero for the
1007 < enthalpies of formation for these systems at the traditional state
1008 < (elemental forms at standard temperature and pressure).  With every
1009 < model in the {\sc ssd} family, Ice-{\it i} had the lowest calculated
1002 < enthalpy of formation.
998 > structure for at least the SSD/E model. To verify this, a comparison
999 > was made between near ideal crystals of ice~$I_h$, ice~$I_c$, and
1000 > Ice-{\it i} at constant pressure with SSD/E, SSD/RF, and
1001 > SSD1. Near-ideal versions of the three types of crystals were cooled
1002 > to 1 K, and enthalpies of formation of each were compared using all
1003 > three water models.  Enthalpies were estimated from the
1004 > isobaric-isothermal simulations using $H=U+P_{\text ext}V$ where
1005 > $P_{\text ext}$ is the applied pressure.  A constant value of -60.158
1006 > kcal / mol has been added to place our zero for the enthalpies of
1007 > formation for these systems at the traditional state (elemental forms
1008 > at standard temperature and pressure).  With every model in the SSD
1009 > family, Ice-{\it i} had the lowest calculated enthalpy of formation.
1010  
1011   \begin{table}
1012   \begin{center}
1013   \caption{Enthalpies of Formation (in kcal / mol) of the three crystal
1014 < structures (at 1 K) exhibited by the {\sc ssd} family of water models}
1014 > structures (at 1 K) exhibited by the SSD family of water models}
1015   \begin{tabular}{ l  c  c  c  }
1016   \hline \\[-3mm]
1017 < \ \ \ Water Model \ \ \  & \ \ \ Ice-$I_h$ \ \ \ & \ Ice-$I_c$\ \  & \
1018 < Ice-{\it i} \\
1017 > \ \ \ Water Model \ \ \  & \ \ \ Ice-$I_h$ \ \ \ & \ \ \ Ice-$I_c$ \ \ \  &
1018 > \ \ \ \ Ice-{\it i} \\
1019   \hline \\[-3mm]
1020 < \ \ \ {\sc ssd/e} & -12.286 & -12.292 & -13.590 \\
1021 < \ \ \ {\sc ssd/rf} & -12.935 & -12.917 & -14.022 \\
1022 < \ \ \ {\sc ssd1} & -12.496 & -12.411 & -13.417 \\
1023 < \ \ \ {\sc ssd1} (RF) & -12.504 & -12.411 & -13.134 \\
1020 > \ \ \ SSD/E & -72.444 & -72.450 & -73.748 \\
1021 > \ \ \ SSD/RF & -73.093 & -73.075 & -74.180 \\
1022 > \ \ \ SSD1 & -72.654 & -72.569 & -73.575 \\
1023 > \ \ \ SSD1 (RF) & -72.662 & -72.569 & -73.292 \\
1024   \end{tabular}
1025   \label{iceenthalpy}
1026   \end{center}
1027   \end{table}
1028  
1029   In addition to these energetic comparisons, melting simulations were
1030 < performed with ice-{\it i} as the initial configuration using {\sc ssd/e},
1031 < {\sc ssd/rf}, and {\sc ssd1} both with and without a reaction field. The melting
1032 < transitions for both {\sc ssd/e} and {\sc ssd1} without reaction field occurred at
1033 < temperature in excess of 375~K.  {\sc ssd/rf} and {\sc ssd1} with a reaction field
1030 > performed with Ice-{\it i} as the initial configuration using SSD/E,
1031 > SSD/RF, and SSD1 both with and without a reaction field. The melting
1032 > transitions for both SSD/E and SSD1 without reaction field occurred at
1033 > temperature in excess of 375~K.  SSD/RF and SSD1 with a reaction field
1034   showed more reasonable melting transitions near 325~K.  These melting
1035 < point observations clearly show that all of the {\sc ssd}-derived models
1035 > point observations clearly show that all of the SSD-derived models
1036   prefer the ice-{\it i} structure.
1037  
1038   \section{Conclusions}
1039  
1040   The density maximum and temperature dependence of the self-diffusion
1041 < constant were studied for the {\sc ssd} water model, both with and without
1042 < the use of reaction field, via a series of NPT and NVE
1041 > constant were studied for the SSD water model, both with and
1042 > without the use of reaction field, via a series of NPT and NVE
1043   simulations. The constant pressure simulations showed a density
1044   maximum near 260 K. In most cases, the calculated densities were
1045   significantly lower than the densities obtained from other water
1046 < models (and experiment). Analysis of self-diffusion showed {\sc ssd} to
1047 < capture the transport properties of water well in both the liquid and
1048 < supercooled liquid regimes.
1046 > models (and experiment). Analysis of self-diffusion showed SSD
1047 > to capture the transport properties of water well in both the liquid
1048 > and supercooled liquid regimes.
1049  
1050 < In order to correct the density behavior, the original {\sc ssd} model was
1051 < reparameterized for use both with and without a reaction field ({\sc ssd/rf}
1052 < and {\sc ssd/e}), and comparisons were made with {\sc ssd1}, Ichiye's density
1053 < corrected version of {\sc ssd}. Both models improve the liquid structure,
1050 > In order to correct the density behavior, the original SSD model was
1051 > reparameterized for use both with and without a reaction field (SSD/RF
1052 > and SSD/E), and comparisons were made with SSD1, Ichiye's density
1053 > corrected version of SSD. Both models improve the liquid structure,
1054   densities, and diffusive properties under their respective simulation
1055   conditions, indicating the necessity of reparameterization when
1056   changing the method of calculating long-range electrostatic
# Line 1052 | Line 1059 | by the {\sc ssd} family of water models is somewhat tr
1059   simulations of biochemical systems.
1060  
1061   The existence of a novel low-density ice structure that is preferred
1062 < by the {\sc ssd} family of water models is somewhat troubling, since liquid
1063 < simulations on this family of water models at room temperature are
1064 < effectively simulations of supercooled or metastable liquids.  One
1062 > by the SSD family of water models is somewhat troubling, since
1063 > liquid simulations on this family of water models at room temperature
1064 > are effectively simulations of supercooled or metastable liquids.  One
1065   way to destabilize this unphysical ice structure would be to make the
1066   range of angles preferred by the attractive part of the sticky
1067   potential much narrower.  This would require extensive
# Line 1078 | Line 1085 | DMR-0079647.
1085   \newpage
1086  
1087   \bibliographystyle{jcp}
1088 < \bibliography{nptSSD}
1088 > \bibliography{nptSSD}
1089  
1090 < %\pagebreak
1090 > \newpage
1091  
1092 + \begin{list}
1093 +  {Figure \arabic{captions}: }{\usecounter{captions}
1094 +        \setlength{\rightmargin}{\leftmargin}}
1095 +        
1096 + \item Energy conservation using both quaternion-based integration and
1097 + the {\sc dlm} method with increasing time step. The larger time step
1098 + plots are shifted from the true energy baseline (that of $\Delta t$ =
1099 + 0.1~fs) for clarity.
1100 +
1101 + \item Density versus temperature for TIP4P [Ref. \onlinecite{Jorgensen98b}],
1102 + TIP3P [Ref. \onlinecite{Jorgensen98b}], SPC/E
1103 + [Ref. \onlinecite{Clancy94}], SSD without Reaction Field, SSD, and
1104 + experiment [Ref. \onlinecite{CRC80}]. The arrows indicate the change
1105 + in densities observed when turning off the reaction field. The the
1106 + lower than expected densities for the SSD model were what prompted the
1107 + original reparameterization of SSD1 [Ref. \onlinecite{Ichiye03}].
1108 +
1109 + \item Average self-diffusion constant as a function of temperature for
1110 + SSD, SPC/E [Ref. \onlinecite{Clancy94}], and TIP5P
1111 + [Ref. \onlinecite{Jorgensen01}] compared with experimental data
1112 + [Refs. \onlinecite{Gillen72} and \onlinecite{Holz00}]. Of the three
1113 + water models shown, SSD has the least deviation from the experimental
1114 + values. The rapidly increasing diffusion constants for TIP5P and SSD
1115 + correspond to significant decreases in density at the higher
1116 + temperatures.
1117 +
1118 + \item An illustration of angles involved in the correlations observed in
1119 + Fig. \ref{contour}.
1120 +
1121 + \item Contour plots of 2D angular pair correlation functions for
1122 + 512 SSD molecules at 100~K (A \& B) and 300~K (C \& D). Dark areas
1123 + signify regions of enhanced density while light areas signify
1124 + depletion relative to the bulk density. White areas have pair
1125 + correlation values below 0.5 and black areas have values above 1.5.
1126 +
1127 + \item Plots comparing experiment [Ref. \onlinecite{Head-Gordon00_1}] with
1128 + SSD/E and SSD1 without reaction field (top), as well as SSD/RF and
1129 + SSD1 with reaction field turned on (bottom). The insets show the
1130 + respective first peaks in detail. Note how the changes in parameters
1131 + have lowered and broadened the first peak of SSD/E and SSD/RF.
1132 +
1133 + \item Positive and negative isosurfaces of the sticky potential for
1134 + SSD1 (left) and SSD/E \& SSD/RF (right). Light areas
1135 + correspond to the tetrahedral attractive component, and darker areas
1136 + correspond to the dipolar repulsive component.
1137 +
1138 + \item Comparison of densities calculated with SSD/E to
1139 + SSD1 without a reaction field, TIP3P [Ref. \onlinecite{Jorgensen98b}],
1140 + TIP5P [Ref. \onlinecite{Jorgensen00}], SPC/E [Ref. \onlinecite{Clancy94}] and
1141 + experiment [Ref. \onlinecite{CRC80}]. The window shows a expansion around
1142 + 300 K with error bars included to clarify this region of
1143 + interest. Note that both SSD1 and SSD/E show good agreement with
1144 + experiment when the long-range correction is neglected.
1145 +
1146 + \item Comparison of densities calculated with SSD/RF to
1147 + SSD1 with a reaction field, TIP3P [Ref. \onlinecite{Jorgensen98b}],
1148 + TIP5P [Ref. \onlinecite{Jorgensen00}], SPC/E [Ref. \onlinecite{Clancy94}], and
1149 + experiment [Ref. \onlinecite{CRC80}]. The inset shows the necessity of
1150 + reparameterization when utilizing a reaction field long-ranged
1151 + correction - SSD/RF provides significantly more accurate
1152 + densities than SSD1 when performing room temperature
1153 + simulations.
1154 +
1155 + \item The diffusion constants calculated from SSD/E and
1156 + SSD1 (both without a reaction field) along with experimental results
1157 + [Refs. \onlinecite{Gillen72} and \onlinecite{Holz00}]. The NVE calculations were
1158 + performed at the average densities observed in the 1 atm NPT
1159 + simulations for the respective models. SSD/E is slightly more mobile
1160 + than experiment at all of the temperatures, but it is closer to
1161 + experiment at biologically relevant temperatures than SSD1 without a
1162 + long-range correction.
1163 +
1164 + \item The diffusion constants calculated from SSD/RF and
1165 + SSD1 (both with an active reaction field) along with
1166 + experimental results [Refs. \onlinecite{Gillen72} and \onlinecite{Holz00}]. The
1167 + NVE calculations were performed at the average densities observed in
1168 + the 1 atm NPT simulations for both of the models. SSD/RF
1169 + simulates the diffusion of water throughout this temperature range
1170 + very well. The rapidly increasing diffusion constants at high
1171 + temperatures for both models can be attributed to lower calculated
1172 + densities than those observed in experiment.
1173 +
1174 + \item The most stable crystal structure assumed by the SSD family
1175 + of water models.  We refer to this structure as Ice-{\it i} to
1176 + indicate its origins in computer simulation.  This image was taken of
1177 + the (001) face of the crystal.
1178 + \end{list}
1179 +
1180 + \newpage
1181 +
1182 + \begin{figure}
1183 + \begin{center}
1184 + \epsfxsize=6in
1185 + \epsfbox{timeStep.epsi}
1186 + %\caption{Energy conservation using both quaternion-based integration and
1187 + %the {\sc dlm} method with increasing time step. The larger time step
1188 + %plots are shifted from the true energy baseline (that of $\Delta t$ =
1189 + %0.1~fs) for clarity.}
1190 + \label{timestep}
1191 + \end{center}
1192 + \end{figure}
1193 +
1194 + \newpage
1195 +
1196 + \begin{figure}
1197 + \begin{center}
1198 + \epsfxsize=6in
1199 + \epsfbox{denseSSDnew.eps}
1200 + %\caption{Density versus temperature for TIP4P [Ref. \onlinecite{Jorgensen98b}],
1201 + % TIP3P [Ref. \onlinecite{Jorgensen98b}], SPC/E [Ref. \onlinecite{Clancy94}], SSD
1202 + % without Reaction Field, SSD, and experiment [Ref. \onlinecite{CRC80}]. The
1203 + % arrows indicate the change in densities observed when turning off the
1204 + % reaction field. The the lower than expected densities for the SSD
1205 + % model were what prompted the original reparameterization of SSD1
1206 + % [Ref. \onlinecite{Ichiye03}].}
1207 + \label{dense1}
1208 + \end{center}
1209 + \end{figure}
1210 +
1211 + \newpage
1212 +
1213 + \begin{figure}
1214 + \begin{center}
1215 + \epsfxsize=6in
1216 + \epsfbox{betterDiffuse.epsi}
1217 + %\caption{Average self-diffusion constant as a function of temperature for
1218 + %SSD, SPC/E [Ref. \onlinecite{Clancy94}], and TIP5P
1219 + %[Ref. \onlinecite{Jorgensen01}] compared with experimental data
1220 + %[Refs. \onlinecite{Gillen72} and \onlinecite{Holz00}]. Of the three water models
1221 + %shown, SSD has the least deviation from the experimental values. The
1222 + %rapidly increasing diffusion constants for TIP5P and SSD correspond to
1223 + %significant decreases in density at the higher temperatures.}
1224 + \label{diffuse}
1225 + \end{center}
1226 + \end{figure}
1227 +
1228 + \newpage
1229 +
1230 + \begin{figure}
1231 + \begin{center}
1232 + \epsfxsize=6in
1233 + \epsfbox{corrDiag.eps}
1234 + %\caption{An illustration of angles involved in the correlations observed in Fig. \ref{contour}.}
1235 + \label{corrAngle}
1236 + \end{center}
1237 + \end{figure}
1238 +
1239 + \newpage
1240 +
1241 + \begin{figure}
1242 + \begin{center}
1243 + \epsfxsize=6in
1244 + \epsfbox{fullContours.eps}
1245 + %\caption{Contour plots of 2D angular pair correlation functions for
1246 + %512 SSD molecules at 100~K (A \& B) and 300~K (C \& D). Dark areas
1247 + %signify regions of enhanced density while light areas signify
1248 + %depletion relative to the bulk density. White areas have pair
1249 + %correlation values below 0.5 and black areas have values above 1.5.}
1250 + \label{contour}
1251 + \end{center}
1252 + \end{figure}
1253 +
1254 + \newpage
1255 +
1256 + \begin{figure}
1257 + \begin{center}
1258 + \epsfxsize=6in
1259 + \epsfbox{GofRCompare.epsi}
1260 + %\caption{Plots comparing experiment [Ref. \onlinecite{Head-Gordon00_1}] with
1261 + %SSD/E and SSD1 without reaction field (top), as well as
1262 + %SSD/RF and SSD1 with reaction field turned on
1263 + %(bottom). The insets show the respective first peaks in detail. Note
1264 + %how the changes in parameters have lowered and broadened the first
1265 + %peak of SSD/E and SSD/RF.}
1266 + \label{grcompare}
1267 + \end{center}
1268 + \end{figure}
1269 +
1270 + \newpage
1271 +
1272 + \begin{figure}
1273 + \begin{center}
1274 + \epsfxsize=7in
1275 + \epsfbox{dualsticky_bw.eps}
1276 + %\caption{Positive and negative isosurfaces of the sticky potential for
1277 + %SSD1 (left) and SSD/E \& SSD/RF (right). Light areas
1278 + %correspond to the tetrahedral attractive component, and darker areas
1279 + %correspond to the dipolar repulsive component.}
1280 + \label{isosurface}
1281 + \end{center}
1282 + \end{figure}
1283 +
1284 + \newpage
1285 +
1286 + \begin{figure}
1287 + \begin{center}
1288 + \epsfxsize=6in
1289 + \epsfbox{ssdeDense.epsi}
1290 + %\caption{Comparison of densities calculated with SSD/E to
1291 + %SSD1 without a reaction field, TIP3P [Ref. \onlinecite{Jorgensen98b}],
1292 + %TIP5P [Ref. \onlinecite{Jorgensen00}], SPC/E [Ref. \onlinecite{Clancy94}] and
1293 + %experiment [Ref. \onlinecite{CRC80}]. The window shows a expansion around
1294 + %300 K with error bars included to clarify this region of
1295 + %interest. Note that both SSD1 and SSD/E show good agreement with
1296 + %experiment when the long-range correction is neglected.}
1297 + \label{ssdedense}
1298 + \end{center}
1299 + \end{figure}
1300 +
1301 + \newpage
1302 +
1303 + \begin{figure}
1304 + \begin{center}
1305 + \epsfxsize=6in
1306 + \epsfbox{ssdrfDense.epsi}
1307 + %\caption{Comparison of densities calculated with SSD/RF to
1308 + %SSD1 with a reaction field, TIP3P [Ref. \onlinecite{Jorgensen98b}],
1309 + %TIP5P [Ref. \onlinecite{Jorgensen00}], SPC/E [Ref. \onlinecite{Clancy94}], and
1310 + %experiment [Ref. \onlinecite{CRC80}]. The inset shows the necessity of
1311 + %reparameterization when utilizing a reaction field long-ranged
1312 + %correction - SSD/RF provides significantly more accurate
1313 + %densities than SSD1 when performing room temperature
1314 + %simulations.}
1315 + \label{ssdrfdense}
1316 + \end{center}
1317 + \end{figure}
1318 +
1319 + \newpage
1320 +
1321 + \begin{figure}
1322 + \begin{center}
1323 + \epsfxsize=6in
1324 + \epsfbox{ssdeDiffuse.epsi}
1325 + %\caption{The diffusion constants calculated from SSD/E and
1326 + %SSD1 (both without a reaction field) along with experimental results
1327 + %[Refs. \onlinecite{Gillen72} and \onlinecite{Holz00}]. The NVE calculations were
1328 + %performed at the average densities observed in the 1 atm NPT
1329 + %simulations for the respective models. SSD/E is slightly more mobile
1330 + %than experiment at all of the temperatures, but it is closer to
1331 + %experiment at biologically relevant temperatures than SSD1 without a
1332 + %long-range correction.}
1333 + \label{ssdediffuse}
1334 + \end{center}
1335 + \end{figure}
1336 +
1337 + \newpage
1338 +
1339 + \begin{figure}
1340 + \begin{center}
1341 + \epsfxsize=6in
1342 + \epsfbox{ssdrfDiffuse.epsi}
1343 + %\caption{The diffusion constants calculated from SSD/RF and
1344 + %SSD1 (both with an active reaction field) along with
1345 + %experimental results [Refs. \onlinecite{Gillen72} and \onlinecite{Holz00}]. The
1346 + %NVE calculations were performed at the average densities observed in
1347 + %the 1 atm NPT simulations for both of the models. SSD/RF
1348 + %simulates the diffusion of water throughout this temperature range
1349 + %very well. The rapidly increasing diffusion constants at high
1350 + %temperatures for both models can be attributed to lower calculated
1351 + %densities than those observed in experiment.}
1352 + \label{ssdrfdiffuse}
1353 + \end{center}
1354 + \end{figure}
1355 +
1356 + \newpage
1357 +
1358 + \begin{figure}
1359 + \begin{center}
1360 + \epsfxsize=6in
1361 + \epsfbox{icei_bw.eps}
1362 + %\caption{The most stable crystal structure assumed by the SSD family
1363 + %of water models.  We refer to this structure as Ice-{\it i} to
1364 + %indicate its origins in computer simulation.  This image was taken of
1365 + %the (001) face of the crystal.}
1366 + \label{weirdice}
1367 + \end{center}
1368 + \end{figure}
1369 +
1370   \end{document}

Diff Legend

Removed lines
+ Added lines
< Changed lines
> Changed lines