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 1027 by chrisfen, Thu Feb 5 18:42:59 2004 UTC vs.
Revision 1029 by gezelter, Thu Feb 5 20:47:50 2004 UTC

# Line 22 | Line 22 | dipole (SSD) and related single point water models}
22   \begin{document}
23  
24   \title{On the structural and transport properties of the soft sticky
25 < dipole (SSD) and related single point water models}
25 > dipole ({\sc ssd}) and related single point water models}
26  
27   \author{Christopher J. Fennell and J. Daniel Gezelter\footnote{Corresponding author. \ Electronic mail: gezelter@nd.edu} \\
28   Department of Chemistry and Biochemistry\\ University of Notre Dame\\
# Line 34 | Line 34 | constant were investigated for the soft sticky dipole
34  
35   \begin{abstract}
36   The density maximum and temperature dependence of the self-diffusion
37 < constant were investigated for the soft sticky dipole (SSD) water
37 > constant were investigated for the soft sticky dipole ({\sc ssd}) water
38   model and two related re-parameterizations of this single-point model.
39   A combination of microcanonical and isobaric-isothermal molecular
40   dynamics simulations were used to calculate these properties, both
# Line 44 | Line 44 | that the original SSD model captures the transport pro
44   260 K.  In most cases, the use of the reaction field resulted in
45   calculated densities which were were significantly lower than
46   experimental densities.  Analysis of self-diffusion constants shows
47 < that the original SSD model captures the transport properties of
47 > that the original {\sc ssd} model captures the transport properties of
48   experimental water very well in both the normal and super-cooled
49 < liquid regimes.  We also present our re-parameterized versions of SSD
49 > liquid regimes.  We also present our re-parameterized versions of {\sc ssd}
50   for use both with the reaction field or without any long-range
51 < electrostatic corrections.  These are called the SSD/RF and SSD/E
51 > electrostatic corrections.  These are called the {\sc ssd/rf} and {\sc ssd/e}
52   models respectively.  These modified models were shown to maintain or
53   improve upon the experimental agreement with the structural and
54 < transport properties that can be obtained with either the original SSD
55 < or the density corrected version of the original model (SSD1).
54 > transport properties that can be obtained with either the original {\sc ssd}
55 > or the density corrected version of the original model ({\sc ssd1}).
56   Additionally, a novel low-density ice structure is presented
57 < which appears to be the most stable ice structure for the entire SSD
57 > which appears to be the most stable ice structure for the entire {\sc ssd}
58   family.
59   \end{abstract}
60  
# Line 89 | Line 89 | cost is the Soft Sticky Dipole (SSD) water
89  
90   One recently developed model that largely succeeds in retaining the
91   accuracy of bulk properties while greatly reducing the computational
92 < cost is the Soft Sticky Dipole (SSD) water
93 < model.\cite{Ichiye96,Ichiye96b,Ichiye99,Ichiye03} The SSD model was
92 > cost is the Soft Sticky Dipole ({\sc ssd}) water
93 > model.\cite{Ichiye96,Ichiye96b,Ichiye99,Ichiye03} The {\sc ssd} model was
94   developed by Ichiye \emph{et al.} as a modified form of the
95   hard-sphere water model proposed by Bratko, Blum, and
96 < Luzar.\cite{Bratko85,Bratko95} SSD is a {\it single point} model which
96 > Luzar.\cite{Bratko85,Bratko95} {\sc ssd} is a {\it single point} model which
97   has an interaction site that is both a point dipole along with a
98   Lennard-Jones core.  However, since the normal aligned and
99   anti-aligned geometries favored by point dipoles are poor mimics of
# Line 102 | Line 102 | The interaction between two SSD water molecules \emph{
102   the proper hydrogen bond orientation in the first solvation
103   shell.  
104  
105 < The interaction between two SSD water molecules \emph{i} and \emph{j}
105 > The interaction between two {\sc ssd} water molecules \emph{i} and \emph{j}
106   is given by the potential
107   \begin{equation}
108   u_{ij} = u_{ij}^{LJ} (r_{ij})\ + u_{ij}^{dp}
# Line 159 | Line 159 | can be found in the original SSD
159   enhances the tetrahedral geometry for hydrogen bonded structures),
160   while $w^\prime$ is a purely empirical function.  A more detailed
161   description of the functional parts and variables in this potential
162 < can be found in the original SSD
162 > can be found in the original {\sc ssd}
163   articles.\cite{Ichiye96,Ichiye96b,Ichiye99,Ichiye03}
164  
165 < Since SSD is a single-point {\it dipolar} model, the force
165 > Since {\sc ssd} is a single-point {\it dipolar} model, the force
166   calculations are simplified significantly relative to the standard
167   {\it charged} multi-point models. In the original Monte Carlo
168   simulations using this model, Ichiye {\it et al.} reported that using
169 < SSD decreased computer time by a factor of 6-7 compared to other
169 > {\sc ssd} decreased computer time by a factor of 6-7 compared to other
170   models.\cite{Ichiye96} What is most impressive is that this savings
171   did not come at the expense of accurate depiction of the liquid state
172 < properties.  Indeed, SSD maintains reasonable agreement with the Soper
172 > properties.  Indeed, {\sc ssd} maintains reasonable agreement with the Soper
173   data for the structural features of liquid
174   water.\cite{Soper86,Ichiye96} Additionally, the dynamical properties
175 < exhibited by SSD agree with experiment better than those of more
175 > exhibited by {\sc ssd} agree with experiment better than those of more
176   computationally expensive models (like TIP3P and
177   SPC/E).\cite{Ichiye99} The combination of speed and accurate depiction
178 < of solvent properties makes SSD a very attractive model for the
178 > of solvent properties makes {\sc ssd} a very attractive model for the
179   simulation of large scale biochemical simulations.
180  
181 < One feature of the SSD model is that it was parameterized for use with
181 > One feature of the {\sc ssd} model is that it was parameterized for use with
182   the Ewald sum to handle long-range interactions.  This would normally
183   be the best way of handling long-range interactions in systems that
184   contain other point charges.  However, our group has recently become
# Line 193 | Line 193 | transport behavior of SSD over a variety of temperatur
193   properties and behavior under the more computationally efficient
194   reaction field (RF) technique, or even with a simple cutoff. This
195   study addresses these issues by looking at the structural and
196 < transport behavior of SSD over a variety of temperatures with the
196 > transport behavior of {\sc ssd} over a variety of temperatures with the
197   purpose of utilizing the RF correction technique.  We then suggest
198   modifications to the parameters that result in more realistic bulk
199   phase behavior.  It should be noted that in a recent publication, some
200 < of the original investigators of the SSD water model have suggested
201 < adjustments to the SSD water model to address abnormal density
200 > of the original investigators of the {\sc ssd} water model have suggested
201 > adjustments to the {\sc ssd} water model to address abnormal density
202   behavior (also observed here), calling the corrected model
203 < SSD1.\cite{Ichiye03} In what follows, we compare our
204 < reparamaterization of SSD with both the original SSD and SSD1 models
205 < with the goal of improving the bulk phase behavior of an SSD-derived
203 > {\sc ssd1}.\cite{Ichiye03} In what follows, we compare our
204 > reparamaterization of {\sc ssd} with both the original {\sc ssd} and {\sc ssd1} models
205 > with the goal of improving the bulk phase behavior of an {\sc ssd}-derived
206   model in simulations utilizing the Reaction Field.
207  
208   \section{Methods}
# Line 215 | Line 215 | field acting on dipole $i$ is
215   field acting on dipole $i$ is
216   \begin{equation}
217   \mathcal{E}_{i} = \frac{2(\varepsilon_{s} - 1)}{2\varepsilon_{s} + 1}
218 < \frac{1}{r_{c}^{3}} \sum_{j\in{\mathcal{R}}} {\bf \mu}_{j} f(r_{ij}),
218 > \frac{1}{r_{c}^{3}} \sum_{j\in{\mathcal{R}}} {\bf \mu}_{j} s(r_{ij}),
219   \label{rfequation}
220   \end{equation}
221   where $\mathcal{R}$ is the cavity defined by the cutoff radius
222   ($r_{c}$), $\varepsilon_{s}$ is the dielectric constant imposed on the
223   system (80 in the case of liquid water), ${\bf \mu}_{j}$ is the dipole
224 < moment vector of particle $j$, and $f(r_{ij})$ is a cubic switching
224 > moment vector of particle $j$, and $s(r_{ij})$ is a cubic switching
225   function.\cite{AllenTildesley} The reaction field contribution to the
226   total energy by particle $i$ is given by $-\frac{1}{2}{\bf
227   \mu}_{i}\cdot\mathcal{E}_{i}$ and the torque on dipole $i$ by ${\bf
# Line 236 | Line 236 | of SSD which could be used either with or without the
236   We have also performed a companion set of simulations {\it without} a
237   surrounding dielectric (i.e. using a simple cubic switching function
238   at the cutoff radius), and as a result we have two reparamaterizations
239 < of SSD which could be used either with or without the reaction field
239 > of {\sc ssd} which could be used either with or without the reaction field
240   turned on.
241  
242 < Simulations to obtain the preferred density were performed in the
243 < isobaric-isothermal (NPT) ensemble, while all dynamical properties
244 < were obtained from microcanonical (NVE) simulations done at densities
245 < matching the NPT density for a particular target temperature.  The
246 < constant pressure simulations were implemented using an integral
247 < thermostat and barostat as outlined by Hoover.\cite{Hoover85,Hoover86}
248 < All molecules were treated as non-linear rigid bodies. Vibrational
249 < constraints are not necessary in simulations of SSD, because there are
250 < no explicit hydrogen atoms, and thus no molecular vibrational modes
251 < need to be considered.
242 > Simulations to obtain the preferred densities of the models were
243 > performed in the isobaric-isothermal (NPT) ensemble, while all
244 > dynamical properties were obtained from microcanonical (NVE)
245 > simulations done at densities matching the NPT density for a
246 > particular target temperature.  The constant pressure simulations were
247 > implemented using an integral thermostat and barostat as outlined by
248 > Hoover.\cite{Hoover85,Hoover86} All molecules were treated as
249 > non-linear rigid bodies. Vibrational constraints are not necessary in
250 > simulations of {\sc ssd}, because there are no explicit hydrogen atoms, and
251 > thus no molecular vibrational modes need to be considered.
252  
253   Integration of the equations of motion was carried out using the
254   symplectic splitting method proposed by Dullweber, Leimkuhler, and
255 < McLachlan (DLM).\cite{Dullweber1997} Our reason for selecting this
255 > McLachlan ({\sc dlm}).\cite{Dullweber1997} Our reason for selecting this
256   integrator centers on poor energy conservation of rigid body dynamics
257   using traditional quaternion integration.\cite{Evans77,Evans77b} In
258   typical microcanonical ensemble simulations, the energy drift when
259 < using quaternions was substantially greater than when using the DLM
259 > using quaternions was substantially greater than when using the {\sc dlm}
260   method (fig. \ref{timestep}).  This steady drift in the total energy
261   has also been observed by Kol {\it et al.}\cite{Laird97}
262  
# Line 267 | Line 267 | The DML method allows for Verlet style integration of
267   rotation matrix into quaternions for storage purposes makes trajectory
268   data quite compact.
269  
270 < The DML method allows for Verlet style integration of both
270 > The {\sc dlm} method allows for Verlet style integration of both
271   translational and orientational motion of rigid bodies. In this
272   integration method, the orientational propagation involves a sequence
273   of matrix evaluations to update the rotation
274   matrix.\cite{Dullweber1997} These matrix rotations are more costly
275   than the simpler arithmetic quaternion propagation. With the same time
276 < step, a 1000 SSD particle simulation shows an average 7\% increase in
277 < computation time using the DML method in place of quaternions. The
276 > step, a 1000 {\sc ssd} particle simulation shows an average 7\% increase in
277 > computation time using the {\sc dlm} method in place of quaternions. The
278   additional expense per step is justified when one considers the
279 < ability to use time steps that are nearly twice as large under DML
279 > ability to use time steps that are nearly twice as large under {\sc dlm}
280   than would be usable under quaternion dynamics.  The energy
281   conservation of the two methods using a number of different time steps
282   is illustrated in figure
# Line 286 | Line 286 | is illustrated in figure
286   \begin{center}
287   \epsfxsize=6in
288   \epsfbox{timeStep.epsi}
289 < \caption{Energy conservation using both quaternion based integration and
290 < the symplectic splitting method proposed by Dullweber \emph{et al.}
291 < with increasing time step. The larger time step plots are shifted from
292 < the true energy baseline (that of $\Delta t$ = 0.1 fs) for clarity.}
289 > \caption{Energy conservation using both quaternion-based integration and
290 > the {\sc dlm} method with increasing time step. The larger time step plots
291 > are shifted from the true energy baseline (that of $\Delta t$ = 0.1
292 > fs) for clarity.}
293   \label{timestep}
294   \end{center}
295   \end{figure}
296  
297   In figure \ref{timestep}, the resulting energy drift at various time
298 < steps for both the DML and quaternion integration schemes is compared.
299 < All of the 1000 SSD particle simulations started with the same
298 > steps for both the {\sc dlm} and quaternion integration schemes is compared.
299 > All of the 1000 {\sc ssd} particle simulations started with the same
300   configuration, and the only difference was the method used to handle
301   orientational motion. At time steps of 0.1 and 0.5 fs, both methods
302   for propagating the orientational degrees of freedom conserve energy
303   fairly well, with the quaternion method showing a slight energy drift
304   over time in the 0.5 fs time step simulation. At time steps of 1 and 2
305 < fs, the energy conservation benefits of the DML method are clearly
305 > fs, the energy conservation benefits of the {\sc dlm} method are clearly
306   demonstrated. Thus, while maintaining the same degree of energy
307   conservation, one can take considerably longer time steps, leading to
308   an overall reduction in computation time.
309  
310 < Energy drift in the simulations using DML integration was unnoticeable
310 > Energy drift in the simulations using {\sc dlm} integration was unnoticeable
311   for time steps up to 3 fs. A slight energy drift on the order of 0.012
312   kcal/mol per nanosecond was observed at a time step of 4 fs, and as
313   expected, this drift increases dramatically with increasing time
# Line 317 | Line 317 | crystals were formed by first arranging the centers of
317  
318   Proton-disordered ice crystals in both the $I_h$ and $I_c$ lattices
319   were generated as starting points for all simulations. The $I_h$
320 < crystals were formed by first arranging the centers of mass of the SSD
320 > crystals were formed by first arranging the centers of mass of the {\sc ssd}
321   particles into a ``hexagonal'' ice lattice of 1024 particles. Because
322   of the crystal structure of $I_h$ ice, the simulation box assumed an
323   orthorhombic shape with an edge length ratio of approximately
# Line 356 | Line 356 | Our initial simulations focused on the original SSD wa
356  
357   \subsection{Density Behavior}
358  
359 < Our initial simulations focused on the original SSD water model, and
359 > Our initial simulations focused on the original {\sc ssd} water model, and
360   an average density versus temperature plot is shown in figure
361   \ref{dense1}. Note that the density maximum when using a reaction
362   field appears between 255 and 265 K.  There were smaller fluctuations
# Line 372 | Line 372 | maximum in this same region (between 255 and 260 K).
372   \epsfxsize=6in
373   \epsfbox{denseSSD.eps}
374   \caption{Density versus temperature for TIP4P [Ref. \citen{Jorgensen98b}],
375 < TIP3P [Ref. \citen{Jorgensen98b}], SPC/E [Ref. \citen{Clancy94}], SSD
376 < without Reaction Field, SSD, and experiment [Ref. \citen{CRC80}]. The
375 > TIP3P [Ref. \citen{Jorgensen98b}], SPC/E [Ref. \citen{Clancy94}], {\sc ssd}
376 > without Reaction Field, {\sc ssd}, and experiment [Ref. \citen{CRC80}]. The
377   arrows indicate the change in densities observed when turning off the
378 < reaction field. The the lower than expected densities for the SSD
379 < model were what prompted the original reparameterization of SSD1
378 > reaction field. The the lower than expected densities for the {\sc ssd}
379 > model were what prompted the original reparameterization of {\sc ssd1}
380   [Ref. \citen{Ichiye03}].}
381   \label{dense1}
382   \end{center}
383   \end{figure}
384  
385 < The density maximum for SSD compares quite favorably to other simple
385 > The density maximum for {\sc ssd} compares quite favorably to other simple
386   water models. Figure \ref{dense1} also shows calculated densities of
387   several other models and experiment obtained from other
388   sources.\cite{Jorgensen98b,Clancy94,CRC80} Of the listed simple water
389 < models, SSD has a temperature closest to the experimentally observed
389 > models, {\sc ssd} has a temperature closest to the experimentally observed
390   density maximum. Of the {\it charge-based} models in
391   Fig. \ref{dense1}, TIP4P has a density maximum behavior most like that
392 < seen in SSD. Though not included in this plot, it is useful
392 > seen in {\sc ssd}. Though not included in this plot, it is useful
393   to note that TIP5P has a density maximum nearly identical to the
394   experimentally measured temperature.
395  
# Line 397 | Line 397 | cutoff radius of 12.0 \AA\ to complement the previous
397   dependent on the cutoff radius used both with and without the use of
398   reaction field.\cite{Berendsen98} In order to address the possible
399   effect of cutoff radius, simulations were performed with a dipolar
400 < cutoff radius of 12.0 \AA\ to complement the previous SSD simulations,
400 > cutoff radius of 12.0 \AA\ to complement the previous {\sc ssd} simulations,
401   all performed with a cutoff of 9.0 \AA. All of the resulting densities
402   overlapped within error and showed no significant trend toward lower
403   or higher densities as a function of cutoff radius, for simulations
404   both with and without reaction field. These results indicate that
405   there is no major benefit in choosing a longer cutoff radius in
406 < simulations using SSD. This is advantageous in that the use of a
406 > simulations using {\sc ssd}. This is advantageous in that the use of a
407   longer cutoff radius results in a significant increase in the time
408   required to obtain a single trajectory.
409  
410   The key feature to recognize in figure \ref{dense1} is the density
411 < scaling of SSD relative to other common models at any given
412 < temperature. SSD assumes a lower density than any of the other listed
411 > scaling of {\sc ssd} relative to other common models at any given
412 > temperature. {\sc ssd} assumes a lower density than any of the other listed
413   models at the same pressure, behavior which is especially apparent at
414   temperatures greater than 300 K. Lower than expected densities have
415   been observed for other systems using a reaction field for long-range
# Line 422 | Line 422 | the curve produced from SSD simulations using reaction
422   \ref{dense1}. Without the reaction field, the densities increase
423   to more experimentally reasonable values, especially around the
424   freezing point of liquid water. The shape of the curve is similar to
425 < the curve produced from SSD simulations using reaction field,
425 > the curve produced from {\sc ssd} simulations using reaction field,
426   specifically the rapidly decreasing densities at higher temperatures;
427   however, a shift in the density maximum location, down to 245 K, is
428   observed. This is a more accurate comparison to the other listed water
# Line 431 | Line 431 | SSD.\cite{Ichiye03} Throughout the remainder of the pa
431   reaction field, the density around 300 K is still significantly lower
432   than experiment and comparable water models. This anomalous behavior
433   was what lead Tan {\it et al.} to recently reparameterize
434 < SSD.\cite{Ichiye03} Throughout the remainder of the paper our
435 < reparamaterizations of SSD will be compared with the newer SSD1 model.
434 > {\sc ssd}.\cite{Ichiye03} Throughout the remainder of the paper our
435 > reparamaterizations of {\sc ssd} will be compared with their newer {\sc ssd1}
436 > model.
437  
438   \subsection{Transport Behavior}
439  
# Line 455 | Line 456 | SSD, SPC/E [Ref. \citen{Clancy94}], TIP5P [Ref. \citen
456   \epsfxsize=6in
457   \epsfbox{betterDiffuse.epsi}
458   \caption{Average self-diffusion constant as a function of temperature for
459 < SSD, SPC/E [Ref. \citen{Clancy94}], TIP5P [Ref. \citen{Jorgensen01}],
460 < and Experimental data [Refs. \citen{Gillen72} and \citen{Holz00}]. Of
461 < the three water models shown, SSD has the least deviation from the
462 < experimental values. The rapidly increasing diffusion constants for
463 < TIP5P and SSD correspond to significant decrease in density at the
464 < higher temperatures.}
459 > {\sc ssd}, SPC/E [Ref. \citen{Clancy94}], and TIP5P
460 > [Ref. \citen{Jorgensen01}] compared with experimental data
461 > [Refs. \citen{Gillen72} and \citen{Holz00}]. Of the three water models
462 > shown, {\sc ssd} has the least deviation from the experimental values. The
463 > rapidly increasing diffusion constants for TIP5P and {\sc ssd} correspond to
464 > significant decreases in density at the higher temperatures.}
465   \label{diffuse}
466   \end{center}
467   \end{figure}
468  
469   The observed values for the diffusion constant point out one of the
470 < strengths of the SSD model. Of the three models shown, the SSD model
470 > strengths of the {\sc ssd} model. Of the three models shown, the {\sc ssd} model
471   has the most accurate depiction of self-diffusion in both the
472   supercooled and liquid regimes.  SPC/E does a respectable job by
473   reproducing values similar to experiment around 290 K; however, it
474   deviates at both higher and lower temperatures, failing to predict the
475 < correct thermal trend. TIP5P and SSD both start off low at colder
475 > correct thermal trend. TIP5P and {\sc ssd} both start off low at colder
476   temperatures and tend to diffuse too rapidly at higher temperatures.
477   This behavior at higher temperatures is not particularly surprising
478 < since the densities of both TIP5P and SSD are lower than experimental
478 > since the densities of both TIP5P and {\sc ssd} are lower than experimental
479   water densities at higher temperatures.  When calculating the
480 < diffusion coefficients for SSD at experimental densities (instead of
480 > diffusion coefficients for {\sc ssd} at experimental densities (instead of
481   the densities from the NPT simulations), the resulting values fall
482   more in line with experiment at these temperatures.
483  
# Line 497 | Line 498 | considerably lower than the experimental value.
498   \begin{center}
499   \epsfxsize=6in
500   \epsfbox{corrDiag.eps}
501 < \caption{Two dimensional illustration of angles involved in the
501 < correlations observed in Fig. \ref{contour}.}
501 > \caption{An illustration of angles involved in the correlations observed in Fig. \ref{contour}.}
502   \label{corrAngle}
503   \end{center}
504   \end{figure}
# Line 507 | Line 507 | correlations observed in Fig. \ref{contour}.}
507   \begin{center}
508   \epsfxsize=6in
509   \epsfbox{fullContours.eps}
510 < \caption{Contour plots of 2D angular g($r$)'s for 512 SSD systems at
511 < 100 K (A \& B) and 300 K (C \& D). Contour colors are inverted for
512 < clarity: dark areas signify peaks while light areas signify
513 < depressions. White areas have $g(r)$ values below 0.5 and black
514 < areas have values above 1.5.}
510 > \caption{Contour plots of 2D angular pair correlation functions for
511 > 512 {\sc ssd} molecules at 100 K (A \& B) and 300 K (C \& D). Dark areas
512 > signify regions of enhanced density while light areas signify
513 > depletion relative to the bulk density. White areas have pair
514 > correlation values below 0.5 and black areas have values above 1.5.}
515   \label{contour}
516   \end{center}
517   \end{figure}
# Line 550 | Line 550 | oxygen-oxygen $g_\mathrm{OO}(r)$.\cite{Ichiye96} At lo
550  
551   This complex interplay between dipole and sticky interactions was
552   remarked upon as a possible reason for the split second peak in the
553 < oxygen-oxygen $g_\mathrm{OO}(r)$.\cite{Ichiye96} At low temperatures,
554 < the second solvation shell peak appears to have two distinct
555 < components that blend together to form one observable peak. At higher
556 < temperatures, this split character alters to show the leading 4 \AA\
557 < peak dominated by equatorial anti-parallel dipole orientations. There
558 < is also a tightly bunched group of axially arranged dipoles that most
559 < likely consist of the smaller fraction of aligned dipole pairs. The
560 < trailing component of the split peak at 5 \AA\ is dominated by aligned
561 < dipoles that assume hydrogen bond arrangements similar to those seen
562 < in the first solvation shell. This evidence indicates that the dipole
563 < pair interaction begins to dominate outside of the range of the
564 < dipolar repulsion term.  The energetically favorable dipole
565 < arrangements populate the region immediately outside this repulsion
566 < region (around 4 \AA), while arrangements that seek to satisfy both
567 < the sticky and dipole forces locate themselves just beyond this
568 < initial buildup (around 5 \AA).
553 > oxygen-oxygen pair correlation function,
554 > $g_\mathrm{OO}(r)$.\cite{Ichiye96} At low temperatures, the second
555 > solvation shell peak appears to have two distinct components that
556 > blend together to form one observable peak. At higher temperatures,
557 > this split character alters to show the leading 4 \AA\ peak dominated
558 > by equatorial anti-parallel dipole orientations. There is also a
559 > tightly bunched group of axially arranged dipoles that most likely
560 > consist of the smaller fraction of aligned dipole pairs. The trailing
561 > component of the split peak at 5 \AA\ is dominated by aligned dipoles
562 > that assume hydrogen bond arrangements similar to those seen in the
563 > first solvation shell. This evidence indicates that the dipole pair
564 > interaction begins to dominate outside of the range of the dipolar
565 > repulsion term.  The energetically favorable dipole arrangements
566 > populate the region immediately outside this repulsion region (around
567 > 4 \AA), while arrangements that seek to satisfy both the sticky and
568 > dipole forces locate themselves just beyond this initial buildup
569 > (around 5 \AA).
570  
571   From these findings, the split second peak is primarily the product of
572   the dipolar repulsion term of the sticky potential. In fact, the inner
# Line 576 | Line 577 | and a density considerably lower than the already low
577   since the second solvation shell would still be shifted too far
578   out. In addition, this would have an even more detrimental effect on
579   the system densities, leading to a liquid with a more open structure
580 < and a density considerably lower than the already low SSD density.  A
580 > and a density considerably lower than the already low {\sc ssd} density.  A
581   better correction would be to include the quadrupole-quadrupole
582   interactions for the water particles outside of the first solvation
583   shell, but this would remove the simplicity and speed advantage of
584 < SSD.
584 > {\sc ssd}.
585  
586 < \subsection{Adjusted Potentials: SSD/RF and SSD/E}
586 > \subsection{Adjusted Potentials: {\sc ssd/rf} and {\sc ssd/e}}
587  
588 < The propensity of SSD to adopt lower than expected densities under
588 > The propensity of {\sc ssd} to adopt lower than expected densities under
589   varying conditions is troubling, especially at higher temperatures. In
590   order to correct this model for use with a reaction field, it is
591   necessary to adjust the force field parameters for the primary
# Line 595 | Line 596 | strength of the sticky potential ($\nu_0$), and the st
596  
597   The parameters available for tuning include the $\sigma$ and
598   $\epsilon$ Lennard-Jones parameters, the dipole strength ($\mu$), the
599 < strength of the sticky potential ($\nu_0$), and the sticky attractive
600 < and dipole repulsive cubic switching function cutoffs ($r_l$, $r_u$
601 < and $r_l^\prime$, $r_u^\prime$ respectively). The results of the
602 < reparameterizations are shown in table \ref{params}. We are calling
603 < these reparameterizations the Soft Sticky Dipole / Reaction Field
604 < (SSD/RF - for use with a reaction field) and Soft Sticky Dipole
605 < Extended (SSD/E - an attempt to improve the liquid structure in
606 < simulations without a long-range correction).
599 > strength of the sticky potential ($\nu_0$), and the cutoff distances
600 > for the sticky attractive and dipole repulsive cubic switching
601 > function cutoffs ($r_l$, $r_u$ and $r_l^\prime$, $r_u^\prime$
602 > respectively). The results of the reparameterizations are shown in
603 > table \ref{params}. We are calling these reparameterizations the Soft
604 > Sticky Dipole / Reaction Field ({\sc ssd/rf} - for use with a reaction
605 > field) and Soft Sticky Dipole Extended ({\sc ssd/e} - an attempt to improve
606 > the liquid structure in simulations without a long-range correction).
607  
608   \begin{table}
609   \begin{center}
610   \caption{Parameters for the original and adjusted models}
611   \begin{tabular}{ l  c  c  c  c }
612   \hline \\[-3mm]
613 < \ \ \ Parameters\ \ \  & \ \ \ SSD [Ref. \citen{Ichiye96}] \ \ \
614 < & \ SSD1 [Ref. \citen{Ichiye03}]\ \  & \ SSD/E\ \  & \ SSD/RF \\
613 > \ \ \ Parameters\ \ \  & \ \ \ {\sc ssd} [Ref. \citen{Ichiye96}] \ \ \
614 > & \ {\sc ssd1} [Ref. \citen{Ichiye03}]\ \  & \ {\sc ssd/e}\ \  & \ {\sc ssd/rf} \\
615   \hline \\[-3mm]
616   \ \ \ $\sigma$ (\AA)  & 3.051 & 3.016 & 3.035 & 3.019\\
617   \ \ \ $\epsilon$ (kcal/mol) & 0.152 & 0.152 & 0.152 & 0.152\\
# Line 630 | Line 631 | simulations without a long-range correction).
631   \begin{center}
632   \epsfxsize=5in
633   \epsfbox{GofRCompare.epsi}
634 < \caption{Plots comparing experiment [Ref. \citen{Head-Gordon00_1}] with SSD/E
635 < and SSD1 without reaction field (top), as well as SSD/RF and SSD1 with
634 > \caption{Plots comparing experiment [Ref. \citen{Head-Gordon00_1}] with {\sc ssd/e}
635 > and {\sc ssd1} without reaction field (top), as well as {\sc ssd/rf} and {\sc ssd1} with
636   reaction field turned on (bottom). The insets show the respective
637   first peaks in detail. Note how the changes in parameters have lowered
638 < and broadened the first peak of SSD/E and SSD/RF.}
638 > and broadened the first peak of {\sc ssd/e} and {\sc ssd/rf}.}
639   \label{grcompare}
640   \end{center}
641   \end{figure}
# Line 643 | Line 644 | and broadened the first peak of SSD/E and SSD/RF.}
644   \begin{center}
645   \epsfxsize=6in
646   \epsfbox{dualsticky_bw.eps}
647 < \caption{Isosurfaces of the sticky potential for SSD1 (left) and SSD/E \&
648 < SSD/RF (right). Light areas correspond to the tetrahedral attractive
649 < component, and darker areas correspond to the dipolar repulsive
650 < component.}
647 > \caption{Positive and negative isosurfaces of the sticky potential for
648 > {\sc ssd1} (left) and {\sc ssd/e} \& {\sc ssd/rf} (right). Light areas correspond to the
649 > tetrahedral attractive component, and darker areas correspond to the
650 > dipolar repulsive component.}
651   \label{isosurface}
652   \end{center}
653   \end{figure}
654  
655 < In the original paper detailing the development of SSD, Liu and Ichiye
655 > In the original paper detailing the development of {\sc ssd}, Liu and Ichiye
656   placed particular emphasis on an accurate description of the first
657   solvation shell. This resulted in a somewhat tall and narrow first
658   peak in $g(r)$ that integrated to give similar coordination numbers to
659   the experimental data obtained by Soper and
660   Phillips.\cite{Ichiye96,Soper86} New experimental x-ray scattering
661   data from the Head-Gordon lab indicates a slightly lower and shifted
662 < first peak in the g$_\mathrm{OO}(r)$, so our adjustments to SSD were
663 < made while taking into consideration the new experimental
662 > first peak in the g$_\mathrm{OO}(r)$, so our adjustments to {\sc ssd} were
663 > made after taking into consideration the new experimental
664   findings.\cite{Head-Gordon00_1} Figure \ref{grcompare} shows the
665   relocation of the first peak of the oxygen-oxygen $g(r)$ by comparing
666 < the revised SSD model (SSD1), SSD/E, and SSD/RF to the new
666 > the revised {\sc ssd} model ({\sc ssd1}), {\sc ssd/e}, and {\sc ssd/rf} to the new
667   experimental results. Both modified water models have shorter peaks
668   that match more closely to the experimental peak (as seen in the
669   insets of figure \ref{grcompare}).  This structural alteration was
# Line 681 | Line 682 | how SSD and SSD1 exclude preferred dipole alignments b
682   to feel the pull of the tetrahedral restructuring. As the particles
683   move closer together, the dipolar repulsion term becomes active and
684   excludes unphysical nearest-neighbor arrangements. This compares with
685 < how SSD and SSD1 exclude preferred dipole alignments before the
685 > how {\sc ssd} and {\sc ssd1} exclude preferred dipole alignments before the
686   particles feel the pull of the ``hydrogen bonds''. Aside from
687   improving the shape of the first peak in the g(\emph{r}), this
688   modification improves the densities considerably by allowing the
# Line 692 | Line 693 | both of our adjusted models. Since SSD is a dipole bas
693   improves the densities, these changes alone are insufficient to bring
694   the system densities up to the values observed experimentally.  To
695   further increase the densities, the dipole moments were increased in
696 < both of our adjusted models. Since SSD is a dipole based model, the
696 > both of our adjusted models. Since {\sc ssd} is a dipole based model, the
697   structure and transport are very sensitive to changes in the dipole
698 < moment. The original SSD simply used the dipole moment calculated from
698 > moment. The original {\sc ssd} simply used the dipole moment calculated from
699   the TIP3P water model, which at 2.35 D is significantly greater than
700   the experimental gas phase value of 1.84 D. The larger dipole moment
701   is a more realistic value and improves the dielectric properties of
# Line 702 | Line 703 | increasing the dipole moments to 2.42 and 2.48 D for S
703   liquid phase dipole moment ranging from 2.4 D to values as high as
704   3.11 D, providing a substantial range of reasonable values for a
705   dipole moment.\cite{Sprik91,Kusalik02,Badyal00,Barriol64} Moderately
706 < increasing the dipole moments to 2.42 and 2.48 D for SSD/E and SSD/RF,
706 > increasing the dipole moments to 2.42 and 2.48 D for {\sc ssd/e} and {\sc ssd/rf},
707   respectively, leads to significant changes in the density and
708   transport of the water models.
709  
710   In order to demonstrate the benefits of these reparameterizations, a
711   series of NPT and NVE simulations were performed to probe the density
712   and transport properties of the adapted models and compare the results
713 < to the original SSD model. This comparison involved full NPT melting
714 < sequences for both SSD/E and SSD/RF, as well as NVE transport
713 > to the original {\sc ssd} model. This comparison involved full NPT melting
714 > sequences for both {\sc ssd/e} and {\sc ssd/rf}, as well as NVE transport
715   calculations at the calculated self-consistent densities. Again, the
716   results are obtained from five separate simulations of 1024 particle
717   systems, and the melting sequences were started from different ice
# Line 725 | Line 726 | collection times as stated previously.
726   \begin{center}
727   \epsfxsize=6in
728   \epsfbox{ssdeDense.epsi}
729 < \caption{Comparison of densities calculated with SSD/E to SSD1 without a
729 > \caption{Comparison of densities calculated with {\sc ssd/e} to {\sc ssd1} without a
730   reaction field, TIP3P [Ref. \citen{Jorgensen98b}], TIP5P
731   [Ref. \citen{Jorgensen00}], SPC/E [Ref. \citen{Clancy94}] and
732   experiment [Ref. \citen{CRC80}]. The window shows a expansion around
733   300 K with error bars included to clarify this region of
734 < interest. Note that both SSD1 and SSD/E show good agreement with
734 > interest. Note that both {\sc ssd1} and {\sc ssd/e} show good agreement with
735   experiment when the long-range correction is neglected.}
736   \label{ssdedense}
737   \end{center}
738   \end{figure}
739  
740 < Fig. \ref{ssdedense} shows the density profile for the SSD/E model
741 < in comparison to SSD1 without a reaction field, other common water
740 > Fig. \ref{ssdedense} shows the density profile for the {\sc ssd/e} model
741 > in comparison to {\sc ssd1} without a reaction field, other common water
742   models, and experimental results. The calculated densities for both
743 < SSD/E and SSD1 have increased significantly over the original SSD
743 > {\sc ssd/e} and {\sc ssd1} have increased significantly over the original {\sc ssd}
744   model (see fig. \ref{dense1}) and are in better agreement with the
745 < experimental values. At 298 K, the densities of SSD/E and SSD1 without
745 > experimental values. At 298 K, the densities of {\sc ssd/e} and {\sc ssd1} without
746   a long-range correction are 0.996$\pm$0.001 g/cm$^3$ and
747   0.999$\pm$0.001 g/cm$^3$ respectively.  These both compare well with
748   the experimental value of 0.997 g/cm$^3$, and they are considerably
749 < better than the SSD value of 0.967$\pm$0.003 g/cm$^3$. The changes to
749 > better than the {\sc ssd} value of 0.967$\pm$0.003 g/cm$^3$. The changes to
750   the dipole moment and sticky switching functions have improved the
751   structuring of the liquid (as seen in figure \ref{grcompare}, but they
752   have shifted the density maximum to much lower temperatures. This
753   comes about via an increase in the liquid disorder through the
754   weakening of the sticky potential and strengthening of the dipolar
755 < character. However, this increasing disorder in the SSD/E model has
755 > character. However, this increasing disorder in the {\sc ssd/e} model has
756   little effect on the melting transition. By monitoring $C_p$
757 < throughout these simulations, the melting transition for SSD/E was
757 > throughout these simulations, the melting transition for {\sc ssd/e} was
758   shown to occur at 235 K.  The same transition temperature observed
759 < with SSD and SSD1.
759 > with {\sc ssd} and {\sc ssd1}.
760  
761   \begin{figure}
762   \begin{center}
763   \epsfxsize=6in
764   \epsfbox{ssdrfDense.epsi}
765 < \caption{Comparison of densities calculated with SSD/RF to SSD1 with a
765 > \caption{Comparison of densities calculated with {\sc ssd/rf} to {\sc ssd1} with a
766   reaction field, TIP3P [Ref. \citen{Jorgensen98b}], TIP5P
767   [Ref. \citen{Jorgensen00}], SPC/E [Ref. \citen{Clancy94}], and
768   experiment [Ref. \citen{CRC80}]. The inset shows the necessity of
769   reparameterization when utilizing a reaction field long-ranged
770 < correction - SSD/RF provides significantly more accurate densities
771 < than SSD1 when performing room temperature simulations.}
770 > correction - {\sc ssd/rf} provides significantly more accurate densities
771 > than {\sc ssd1} when performing room temperature simulations.}
772   \label{ssdrfdense}
773   \end{center}
774   \end{figure}
775  
776   Including the reaction field long-range correction in the simulations
777   results in a more interesting comparison.  A density profile including
778 < SSD/RF and SSD1 with an active reaction field is shown in figure
778 > {\sc ssd/rf} and {\sc ssd1} with an active reaction field is shown in figure
779   \ref{ssdrfdense}.  As observed in the simulations without a reaction
780 < field, the densities of SSD/RF and SSD1 show a dramatic increase over
781 < normal SSD (see figure \ref{dense1}). At 298 K, SSD/RF has a density
780 > field, the densities of {\sc ssd/rf} and {\sc ssd1} show a dramatic increase over
781 > normal {\sc ssd} (see figure \ref{dense1}). At 298 K, {\sc ssd/rf} has a density
782   of 0.997$\pm$0.001 g/cm$^3$, directly in line with experiment and
783 < considerably better than the original SSD value of 0.941$\pm$0.001
784 < g/cm$^3$ and the SSD1 value of 0.972$\pm$0.002 g/cm$^3$. These results
783 > considerably better than the original {\sc ssd} value of 0.941$\pm$0.001
784 > g/cm$^3$ and the {\sc ssd1} value of 0.972$\pm$0.002 g/cm$^3$. These results
785   further emphasize the importance of reparameterization in order to
786   model the density properly under different simulation conditions.
787   Again, these changes have only a minor effect on the melting point,
788 < which observed at 245 K for SSD/RF, is identical to SSD and only 5 K
789 < lower than SSD1 with a reaction field. Additionally, the difference in
790 < density maxima is not as extreme, with SSD/RF showing a density
788 > which observed at 245 K for {\sc ssd/rf}, is identical to {\sc ssd} and only 5 K
789 > lower than {\sc ssd1} with a reaction field. Additionally, the difference in
790 > density maxima is not as extreme, with {\sc ssd/rf} showing a density
791   maximum at 255 K, fairly close to the density maxima of 260 K and 265
792 < K, shown by SSD and SSD1 respectively.
792 > K, shown by {\sc ssd} and {\sc ssd1} respectively.
793  
794   \begin{figure}
795   \begin{center}
796   \epsfxsize=6in
797   \epsfbox{ssdeDiffuse.epsi}
798 < \caption{The diffusion constants calculated from SSD/E and SSD1,
799 < both without a reaction field, along with experimental results
800 < [Refs. \citen{Gillen72} and \citen{Holz00}]. The NVE calculations
801 < were performed at the average densities observed in the 1 atm NPT
802 < simulations for the respective models. SSD/E is slightly more mobile
803 < than experiment at all of the temperatures, but it is closer to
804 < experiment at biologically relevant temperatures than SSD1 without a
805 < long-range correction.}
798 > \caption{The diffusion constants calculated from {\sc ssd/e} and {\sc ssd1} (both
799 > without a reaction field) along with experimental results
800 > [Refs. \citen{Gillen72} and \citen{Holz00}]. The NVE calculations were
801 > performed at the average densities observed in the 1 atm NPT
802 > simulations for the respective models. {\sc ssd/e} is slightly more mobile
803 > than experiment at all of the temperatures, but it is closer to
804 > experiment at biologically relevant temperatures than {\sc ssd1} without a
805 > long-range correction.}
806   \label{ssdediffuse}
807   \end{center}
808   \end{figure}
809  
810 < The reparameterization of the SSD water model, both for use with and
810 > The reparameterization of the {\sc ssd} water model, both for use with and
811   without an applied long-range correction, brought the densities up to
812   what is expected for simulating liquid water. In addition to improving
813 < the densities, it is important that the excellent diffusive behavior
814 < of SSD be maintained or improved. Figure \ref{ssdediffuse} compares
815 < the temperature dependence of the diffusion constant of SSD/E to SSD1
813 > the densities, it is important that the diffusive behavior of {\sc ssd} be
814 > maintained or improved. Figure \ref{ssdediffuse} compares the
815 > temperature dependence of the diffusion constant of {\sc ssd/e} to {\sc ssd1}
816   without an active reaction field at the densities calculated from
817   their respective NPT simulations at 1 atm. The diffusion constant for
818 < SSD/E is consistently higher than experiment, while SSD1 remains lower
818 > {\sc ssd/e} is consistently higher than experiment, while {\sc ssd1} remains lower
819   than experiment until relatively high temperatures (around 360
820   K). Both models follow the shape of the experimental curve well below
821   300 K but tend to diffuse too rapidly at higher temperatures, as seen
822 < in SSD1's crossing above 360 K.  This increasing diffusion relative to
822 > in {\sc ssd1}'s crossing above 360 K.  This increasing diffusion relative to
823   the experimental values is caused by the rapidly decreasing system
824 < density with increasing temperature.  Both SSD1 and SSD/E show this
824 > density with increasing temperature.  Both {\sc ssd1} and {\sc ssd/e} show this
825   deviation in particle mobility, but this trend has different
826 < implications on the diffusive behavior of the models.  While SSD1
826 > implications on the diffusive behavior of the models.  While {\sc ssd1}
827   shows more experimentally accurate diffusive behavior in the high
828 < temperature regimes, SSD/E shows more accurate behavior in the
828 > temperature regimes, {\sc ssd/e} shows more accurate behavior in the
829   supercooled and biologically relevant temperature ranges.  Thus, the
830   changes made to improve the liquid structure may have had an adverse
831   affect on the density maximum, but they improve the transport behavior
832 < of SSD/E relative to SSD1 under the most commonly simulated
832 > of {\sc ssd/e} relative to {\sc ssd1} under the most commonly simulated
833   conditions.
834  
835   \begin{figure}
836   \begin{center}
837   \epsfxsize=6in
838   \epsfbox{ssdrfDiffuse.epsi}
839 < \caption{The diffusion constants calculated from SSD/RF and SSD1,
840 < both with an active reaction field, along with experimental results
841 < [Refs. \citen{Gillen72} and \citen{Holz00}]. The NVE calculations
842 < were performed at the average densities observed in the 1 atm NPT
843 < simulations for both of the models. Note how accurately SSD/RF
844 < simulates the diffusion of water throughout this temperature
845 < range. The more rapidly increasing diffusion constants at high
846 < temperatures for both models is attributed to lower calculated
847 < densities than those observed in experiment.}
839 > \caption{The diffusion constants calculated from {\sc ssd/rf} and {\sc ssd1} (both
840 > with an active reaction field) along with experimental results
841 > [Refs. \citen{Gillen72} and \citen{Holz00}]. The NVE calculations were
842 > performed at the average densities observed in the 1 atm NPT
843 > simulations for both of the models. {\sc ssd/rf} simulates the diffusion of
844 > water throughout this temperature range very well. The rapidly
845 > increasing diffusion constants at high temperatures for both models
846 > can be attributed to lower calculated densities than those observed in
847 > experiment.}
848   \label{ssdrfdiffuse}
849   \end{center}
850   \end{figure}
851  
852 < In figure \ref{ssdrfdiffuse}, the diffusion constants for SSD/RF are
853 < compared to SSD1 with an active reaction field. Note that SSD/RF
852 > In figure \ref{ssdrfdiffuse}, the diffusion constants for {\sc ssd/rf} are
853 > compared to {\sc ssd1} with an active reaction field. Note that {\sc ssd/rf}
854   tracks the experimental results quantitatively, identical within error
855   throughout most of the temperature range shown and exhibiting only a
856 < slight increasing trend at higher temperatures. SSD1 tends to diffuse
856 > slight increasing trend at higher temperatures. {\sc ssd1} tends to diffuse
857   more slowly at low temperatures and deviates to diffuse too rapidly at
858   temperatures greater than 330 K.  As stated above, this deviation away
859   from the ideal trend is due to a rapid decrease in density at higher
860 < temperatures. SSD/RF does not suffer from this problem as much as SSD1
860 > temperatures. {\sc ssd/rf} does not suffer from this problem as much as {\sc ssd1}
861   because the calculated densities are closer to the experimental
862   values. These results again emphasize the importance of careful
863   reparameterization when using an altered long-range correction.
864  
865   \begin{table}
866 + \begin{minipage}{\linewidth}
867 + \renewcommand{\thefootnote}{\thempfootnote}
868   \begin{center}
869 < \caption{Calculated and experimental properties of the single point waters and liquid water at 298 K and 1 atm. (a) Ref. [\citen{Mills73}]. (b) Calculated by integrating the data in ref. \citen{Head-Gordon00_1}. (c) Calculated by integrating the data in ref. \citen{Soper86}. (d) Calculated for 298 K from data in ref. [\citen{Eisenberg69}]. (e) Calculated for 298 K from data in ref. \citen{Krynicki66}.}
869 > \caption{Properties of the single-point water models compared with
870 > experimental data at ambient conditions}
871   \begin{tabular}{ l  c  c  c  c  c }
872   \hline \\[-3mm]
873 < \ \ \ \ \ \  & \ \ \ SSD1 \ \ \ & \ SSD/E \ \ \ & \ SSD1 (RF) \ \
874 < \ & \ SSD/RF \ \ \ & \ Expt. \\
873 > \ \ \ \ \ \  & \ \ \ {\sc ssd1} \ \ \ & \ {\sc ssd/e} \ \ \ & \ {\sc ssd1} (RF) \ \
874 > \ & \ {\sc ssd/rf} \ \ \ & \ Expt. \\
875   \hline \\[-3mm]
876   \ \ \ $\rho$ (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 \\
877   \ \ \ $C_p$ (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 \\
878 < \ \ \ $D$ ($10^{-5}$ cm$^2$/s) & 1.78 $\pm$0.07 & 2.51 $\pm$0.18 & 2.00 $\pm$0.17 & 2.32 $\pm$0.06 & 2.299$^\text{a}$ \\
879 < \ \ \ Coordination Number & 3.9 & 4.3 & 3.8 & 4.4 & 4.7$^\text{b}$ \\
880 < \ \ \ H-bonds per particle & 3.7 & 3.6 & 3.7 & 3.7 & 3.5$^\text{c}$ \\
881 < \ \ \ $\tau_1$ (ps) & 10.9 $\pm$0.6 & 7.3 $\pm$0.4 & 7.5 $\pm$0.7 & 7.2 $\pm$0.4 & 5.7$^\text{d}$ \\
882 < \ \ \ $\tau_2$ (ps) & 4.7 $\pm$0.4 & 3.1 $\pm$0.2 & 3.5 $\pm$0.3 & 3.2 $\pm$0.2 & 2.3$^\text{e}$ \\
878 > \ \ \ $D$ ($10^{-5}$ cm$^2$/s) & 1.78 $\pm$0.07 & 2.51 $\pm$0.18 &
879 > 2.00 $\pm$0.17 & 2.32 $\pm$0.06 & 2.299\cite{Mills73} \\
880 > \ \ \ Coordination Number ($n_C$) & 3.9 & 4.3 & 3.8 & 4.4 &
881 > 4.7\footnote{Calculated by integrating $g_{\text{OO}}(r)$ in
882 > Ref. \citen{Head-Gordon00_1}} \\
883 > \ \ \ H-bonds per particle ($n_H$) & 3.7 & 3.6 & 3.7 & 3.7 &
884 > 3.5\footnote{Calculated by integrating $g_{\text{OH}}(r)$ in
885 > Ref. \citen{Soper86}}  \\
886 > \ \ \ $\tau_1$ (ps) & 10.9 $\pm$0.6 & 7.3 $\pm$0.4 & 7.5 $\pm$0.7 &
887 > 7.2 $\pm$0.4 & 5.7\footnote{Calculated for 298 K from data in Ref. \citen{Eisenberg69}} \\
888 > \ \ \ $\tau_2$ (ps) & 4.7 $\pm$0.4 & 3.1 $\pm$0.2 & 3.5 $\pm$0.3 & 3.2
889 > $\pm$0.2 & 2.3\footnote{Calculated for 298 K from data in
890 > Ref. \citen{Krynicki66}}
891   \end{tabular}
892   \label{liquidproperties}
893   \end{center}
894 + \end{minipage}
895   \end{table}
896  
897   Table \ref{liquidproperties} gives a synopsis of the liquid state
898   properties of the water models compared in this study along with the
899   experimental values for liquid water at ambient conditions. The
900 < coordination number ($N_C$) and hydrogen bonds per particle ($N_H$)
901 < were calculated by integrating the following relations:
900 > coordination number ($n_C$) and number of hydrogen bonds per particle
901 > ($n_H$) were calculated by integrating the following relations:
902   \begin{equation}
903 < N_C = 4\pi\rho_{\text{OO}}\int_{0}^{a}r^2\text{g}_{\text{OO}}(r)dr,
903 > n_C = 4\pi\rho_{\text{OO}}\int_{0}^{a}r^2\text{g}_{\text{OO}}(r)dr,
904   \end{equation}
905   \begin{equation}
906 < N_H = 4\pi\rho_{\text{OH}}\int_{0}^{b}r^2\text{g}_{\text{OH}}(r)dr,
906 > n_H = 4\pi\rho_{\text{OH}}\int_{0}^{b}r^2\text{g}_{\text{OH}}(r)dr,
907   \end{equation}
908   where $\rho$ is the number density of the specified pair interactions,
909   $a$ and $b$ are the radial locations of the minima following the first
910 < solvation shell peak in g$_\text{OO}(r)$ or g$_\text{OH}(r)$
911 < respectively. The number of hydrogen bonds stays relatively constant
912 < across all of the models, but the coordination numbers of SSD/E and
913 < SSD/RF show an improvement over SSD1. This improvement is primarily
914 < due to the widening of the first solvation shell peak, allowing the
915 < first minima to push outward. Comparing the coordination number with
916 < the number of hydrogen bonds can lead to more insight into the
917 < structural character of the liquid.  Because of the near identical
918 < values for SSD1, it appears to be a little too exclusive, in that all
919 < molecules in the first solvation shell are involved in forming ideal
920 < hydrogen bonds.  The differing numbers for the newly parameterized
921 < models indicate the allowance of more fluid configurations in addition
922 < to the formation of an acceptable number of ideal hydrogen bonds.
910 <
911 < The time constants for the self orientational autocorrelation function
910 > peak in g$_\text{OO}(r)$ or g$_\text{OH}(r)$ respectively. The number
911 > of hydrogen bonds stays relatively constant across all of the models,
912 > but the coordination numbers of {\sc ssd/e} and {\sc ssd/rf} show an improvement
913 > over {\sc ssd1}.  This improvement is primarily due to extension of the
914 > first solvation shell in the new parameter sets.  Because $n_H$ and
915 > $n_C$ are nearly identical in {\sc ssd1}, it appears that all molecules in
916 > the first solvation shell are involved in hydrogen bonds.  Since $n_H$
917 > and $n_C$ differ in the newly parameterized models, the orientations
918 > in the first solvation shell are a bit more ``fluid''.  Therefore {\sc ssd1}
919 > overstructures the first solvation shell and our reparameterizations
920 > have returned this shell to more realistic liquid-like behavior.
921 >
922 > The time constants for the orientational autocorrelation functions
923   are also displayed in Table \ref{liquidproperties}. The dipolar
924 < orientational time correlation function ($\Gamma_{l}$) is described
924 > orientational time correlation functions ($C_{l}$) are described
925   by:
926   \begin{equation}
927 < \Gamma_{l}(t) = \langle P_l[\mathbf{u}_j(0)\cdot\mathbf{u}_j(t)]\rangle,
927 > C_{l}(t) = \langle P_l[\hat{\mathbf{u}}_j(0)\cdot\hat{\mathbf{u}}_j(t)]\rangle,
928   \end{equation}
929 < where $P_l$ is a Legendre polynomial of order $l$ and $\mathbf{u}_j$
930 < is the unit vector of the particle dipole.\cite{Rahman71} From these
931 < correlation functions, the orientational relaxation time of the dipole
932 < vector can be calculated from an exponential fit in the long-time
933 < regime ($t > \tau_l$).\cite{Rothschild84} Calculation of these
934 < time constants were averaged from five detailed NVE simulations
935 < performed at the STP density for each of the respective models. It
936 < should be noted that the commonly cited value for $\tau_2$ of 1.9 ps
937 < was determined from the NMR data in reference \citen{Krynicki66} at a
938 < temperature near 34$^\circ$C.\cite{Rahman71} Because of the strong
939 < temperature dependence of $\tau_2$, it is necessary to recalculate it
940 < at 298 K to make proper comparisons. The value shown in Table
929 > where $P_l$ are Legendre polynomials of order $l$ and
930 > $\hat{\mathbf{u}}_j$ is the unit vector pointing along the molecular
931 > dipole.\cite{Rahman71} From these correlation functions, the
932 > orientational relaxation time of the dipole vector can be calculated
933 > from an exponential fit in the long-time regime ($t >
934 > \tau_l$).\cite{Rothschild84} Calculation of these time constants were
935 > averaged over five detailed NVE simulations performed at the ambient
936 > conditions for each of the respective models. It should be noted that
937 > the commonly cited value of 1.9 ps for $\tau_2$ was determined from
938 > the NMR data in Ref. \citen{Krynicki66} at a temperature near
939 > 34$^\circ$C.\cite{Rahman71} Because of the strong temperature
940 > dependence of $\tau_2$, it is necessary to recalculate it at 298 K to
941 > make proper comparisons. The value shown in Table
942   \ref{liquidproperties} was calculated from the same NMR data in the
943 < fashion described in reference \citen{Krynicki66}. Similarly, $\tau_1$
944 < was recomputed for 298 K from the data in ref \citen{Eisenberg69}.
945 < Again, SSD/E and SSD/RF show improved behavior over SSD1, both with
943 > fashion described in Ref. \citen{Krynicki66}. Similarly, $\tau_1$ was
944 > recomputed for 298 K from the data in Ref. \citen{Eisenberg69}.
945 > Again, {\sc ssd/e} and {\sc ssd/rf} show improved behavior over {\sc ssd1}, both with
946   and without an active reaction field. Turning on the reaction field
947 < leads to much improved time constants for SSD1; however, these results
948 < also include a corresponding decrease in system density. Numbers
949 < published from the original SSD dynamics studies are shorter than the
950 < values observed here, and this difference can be attributed to the use
951 < of the Ewald sum technique versus a reaction field.\cite{Ichiye99}
947 > leads to much improved time constants for {\sc ssd1}; however, these results
948 > also include a corresponding decrease in system density.
949 > Orientational relaxation times published in the original {\sc ssd} dynamics
950 > papers are smaller than the values observed here, and this difference
951 > can be attributed to the use of the Ewald sum.\cite{Ichiye99}
952  
953   \subsection{Additional Observations}
954  
# Line 944 | Line 956 | of the Ewald sum technique versus a reaction field.\ci
956   \begin{center}
957   \epsfxsize=6in
958   \epsfbox{icei_bw.eps}
959 < \caption{A water lattice built from the crystal structure assumed by
960 < SSD/E when undergoing an extremely restricted temperature NPT
961 < simulation. This form of ice is referred to as ice-{\it i} to
962 < emphasize its simulation origins. This image was taken of the (001)
951 < face of the crystal.}
959 > \caption{The most stable crystal structure assumed by the {\sc ssd} family
960 > of water models.  We refer to this structure as Ice-{\it i} to
961 > indicate its origins in computer simulation.  This image was taken of
962 > the (001) face of the crystal.}
963   \label{weirdice}
964   \end{center}
965   \end{figure}
966  
967   While performing a series of melting simulations on an early iteration
968 < of SSD/E not discussed in this paper, we observed recrystallization
968 > of {\sc ssd/e} not discussed in this paper, we observed recrystallization
969   into a novel structure not previously known for water.  After melting
970   at 235 K, two of five systems underwent crystallization events near
971   245 K.  The two systems remained crystalline up to 320 and 330 K,
# Line 962 | Line 973 | Ice-$\sqrt{\smash[b]{-\text{I}}}$ (ice-{\it i}).  A la
973   that does not correspond to any known form of ice.  This appears to be
974   an artifact of the point dipolar models, so to distinguish it from the
975   experimentally observed forms of ice, we have denoted the structure
976 < Ice-$\sqrt{\smash[b]{-\text{I}}}$ (ice-{\it i}).  A large enough
976 > Ice-$\sqrt{\smash[b]{-\text{I}}}$ (Ice-{\it i}).  A large enough
977   portion of the sample crystallized that we have been able to obtain a
978 < near ideal crystal structure of ice-{\it i}. Figure \ref{weirdice}
978 > near ideal crystal structure of Ice-{\it i}. Figure \ref{weirdice}
979   shows the repeating crystal structure of a typical crystal at 5
980   K. Each water molecule is hydrogen bonded to four others; however, the
981   hydrogen bonds are bent rather than perfectly straight. This results
# Line 975 | Line 986 | Initial simulations indicated that ice-{\it i} is the
986   configuration. Though not ideal, these flexed hydrogen bonds are
987   favorable enough to stabilize an entire crystal generated around them.
988  
989 < Initial simulations indicated that ice-{\it i} is the preferred ice
990 < structure for at least the SSD/E model. To verify this, a comparison
991 < was made between near ideal crystals of ice~$I_h$, ice~$I_c$, and
992 < ice-{\it i} at constant pressure with SSD/E, SSD/RF, and
993 < SSD1. Near-ideal versions of the three types of crystals were cooled
994 < to 1 K, and the enthalpies of each were compared using all three water
995 < models. With every model in the SSD family, ice-{\it i} had the lowest
996 < calculated enthalpy: 5\% lower than $I_h$ with SSD1, 6.5\% lower with
997 < SSD/E, and 7.5\% lower with SSD/RF.  The enthalpy data is summarized
998 < in Table \ref{iceenthalpy}.
989 > Initial simulations indicated that Ice-{\it i} is the preferred ice
990 > structure for at least the {\sc ssd/e} model. To verify this, a
991 > comparison was made between near ideal crystals of ice~$I_h$,
992 > ice~$I_c$, and Ice-{\it i} at constant pressure with {\sc ssd/e}, {\sc
993 > ssd/rf}, and {\sc ssd1}. Near-ideal versions of the three types of
994 > crystals were cooled to 1 K, and enthalpies of formation of each were
995 > compared using all three water models.  Enthalpies were estimated from
996 > the isobaric-isothermal simulations using $H=U+P_{\text ext}V$ where
997 > $P_{\text ext}$ is the applied pressure.  A constant value of
998 > -60.158 kcal / mol has been added to place our zero for the
999 > enthalpies of formation for these systems at the traditional state
1000 > (elemental forms at standard temperature and pressure).  With every
1001 > model in the {\sc ssd} family, Ice-{\it i} had the lowest calculated
1002 > enthalpy of formation.
1003  
1004   \begin{table}
1005   \begin{center}
1006 < \caption{Enthalpies (in kcal / mol) of the three crystal structures (at 1
1007 < K) exhibited by the SSD family of water models}
1006 > \caption{Enthalpies of Formation (in kcal / mol) of the three crystal
1007 > structures (at 1 K) exhibited by the {\sc ssd} family of water models}
1008   \begin{tabular}{ l  c  c  c  }
1009   \hline \\[-3mm]
1010   \ \ \ Water Model \ \ \  & \ \ \ Ice-$I_h$ \ \ \ & \ Ice-$I_c$\ \  & \
1011   Ice-{\it i} \\
1012   \hline \\[-3mm]
1013 < \ \ \ SSD/E & -12.286 & -12.292 & -13.590 \\
1014 < \ \ \ SSD/RF & -12.935 & -12.917 & -14.022 \\
1015 < \ \ \ SSD1 & -12.496 & -12.411 & -13.417 \\
1016 < \ \ \ SSD1 (RF) & -12.504 & -12.411 & -13.134 \\
1013 > \ \ \ {\sc ssd/e} & -12.286 & -12.292 & -13.590 \\
1014 > \ \ \ {\sc ssd/rf} & -12.935 & -12.917 & -14.022 \\
1015 > \ \ \ {\sc ssd1} & -12.496 & -12.411 & -13.417 \\
1016 > \ \ \ {\sc ssd1} (RF) & -12.504 & -12.411 & -13.134 \\
1017   \end{tabular}
1018   \label{iceenthalpy}
1019   \end{center}
1020   \end{table}
1021  
1022   In addition to these energetic comparisons, melting simulations were
1023 < performed with ice-{\it i} as the initial configuration using SSD/E,
1024 < SSD/RF, and SSD1 both with and without a reaction field. The melting
1025 < transitions for both SSD/E and SSD1 without reaction field occurred at
1026 < temperature in excess of 375~K.  SSD/RF and SSD1 with a reaction field
1023 > performed with ice-{\it i} as the initial configuration using {\sc ssd/e},
1024 > {\sc ssd/rf}, and {\sc ssd1} both with and without a reaction field. The melting
1025 > transitions for both {\sc ssd/e} and {\sc ssd1} without reaction field occurred at
1026 > temperature in excess of 375~K.  {\sc ssd/rf} and {\sc ssd1} with a reaction field
1027   showed more reasonable melting transitions near 325~K.  These melting
1028 < point observations clearly show that all of the SSD-derived models
1028 > point observations clearly show that all of the {\sc ssd}-derived models
1029   prefer the ice-{\it i} structure.
1030  
1031   \section{Conclusions}
1032  
1033   The density maximum and temperature dependence of the self-diffusion
1034 < constant were studied for the SSD water model, both with and without
1034 > constant were studied for the {\sc ssd} water model, both with and without
1035   the use of reaction field, via a series of NPT and NVE
1036   simulations. The constant pressure simulations showed a density
1037   maximum near 260 K. In most cases, the calculated densities were
1038   significantly lower than the densities obtained from other water
1039 < models (and experiment). Analysis of self-diffusion showed SSD to
1039 > models (and experiment). Analysis of self-diffusion showed {\sc ssd} to
1040   capture the transport properties of water well in both the liquid and
1041   supercooled liquid regimes.
1042  
1043 < In order to correct the density behavior, the original SSD model was
1044 < reparameterized for use both with and without a reaction field (SSD/RF
1045 < and SSD/E), and comparisons were made with SSD1, Ichiye's density
1046 < corrected version of SSD. Both models improve the liquid structure,
1043 > In order to correct the density behavior, the original {\sc ssd} model was
1044 > reparameterized for use both with and without a reaction field ({\sc ssd/rf}
1045 > and {\sc ssd/e}), and comparisons were made with {\sc ssd1}, Ichiye's density
1046 > corrected version of {\sc ssd}. Both models improve the liquid structure,
1047   densities, and diffusive properties under their respective simulation
1048   conditions, indicating the necessity of reparameterization when
1049   changing the method of calculating long-range electrostatic
# Line 1037 | Line 1052 | by the SSD family of water models is somewhat troublin
1052   simulations of biochemical systems.
1053  
1054   The existence of a novel low-density ice structure that is preferred
1055 < by the SSD family of water models is somewhat troubling, since liquid
1055 > by the {\sc ssd} family of water models is somewhat troubling, since liquid
1056   simulations on this family of water models at room temperature are
1057   effectively simulations of supercooled or metastable liquids.  One
1058   way to destabilize this unphysical ice structure would be to make the
# Line 1046 | Line 1061 | Additionally, our initial calculations show that the i
1061   reparameterization to maintain the same level of agreement with the
1062   experiments.
1063  
1064 < Additionally, our initial calculations show that the ice-{\it i}
1064 > Additionally, our initial calculations show that the Ice-{\it i}
1065   structure may also be a preferred crystal structure for at least one
1066   other popular multi-point water model (TIP3P), and that much of the
1067   simulation work being done using this popular model could also be at

Diff Legend

Removed lines
+ Added lines
< Changed lines
> Changed lines