ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/ssdePaper/nptSSD.tex
Revision: 856
Committed: Thu Nov 6 23:00:00 2003 UTC (20 years, 7 months ago) by chrisfen
Content type: application/x-tex
File size: 49811 byte(s)
Log Message:
Fixed plots by including SSD1 data.  Made several images more condensed and
easier to read.  Begun altering text to match the new focus.

File Contents

# Content
1 \documentclass[prb,aps,times,twocolumn,tabularx]{revtex4}
2 %\documentclass[prb,aps,times,tabularx,preprint]{revtex4}
3 \usepackage{amsmath}
4 \usepackage{graphicx}
5
6 %\usepackage{endfloat}
7 %\usepackage{berkeley}
8 %\usepackage{epsf}
9 %\usepackage[ref]{overcite}
10 %\usepackage{setspace}
11 %\usepackage{tabularx}
12 %\usepackage{graphicx}
13 %\usepackage{curves}
14 %\usepackage{amsmath}
15 %\pagestyle{plain}
16 %\pagenumbering{arabic}
17 %\oddsidemargin 0.0cm \evensidemargin 0.0cm
18 %\topmargin -21pt \headsep 10pt
19 %\textheight 9.0in \textwidth 6.5in
20 %\brokenpenalty=10000
21 %\renewcommand{\baselinestretch}{1.2}
22 %\renewcommand\citemid{\ } % no comma in optional reference note
23
24 \begin{document}
25
26 \title{On the temperature dependent properties of the soft sticky dipole (SSD) and related single point water models}
27
28 \author{Christopher J. Fennell and J. Daniel Gezelter{\thefootnote}
29 \footnote[1]{Corresponding author. \ Electronic mail: gezelter@nd.edu}}
30
31 \address{Department of Chemistry and Biochemistry\\ University of Notre Dame\\
32 Notre Dame, Indiana 46556}
33
34 \date{\today}
35
36 \begin{abstract}
37 NVE and NPT molecular dynamics simulations were performed in order to
38 investigate the density maximum and temperature dependent transport
39 for SSD and related water models, both with and without the use of
40 reaction field. The constant pressure simulations of the melting of
41 both $I_h$ and $I_c$ ice showed a density maximum near 260 K. In most
42 cases, the calculated densities were significantly lower than the
43 densities calculated in simulations of other water models. Analysis of
44 particle diffusion showed SSD to capture the transport properties of
45 experimental very well in both the normal and super-cooled liquid
46 regimes. In order to correct the density behavior, SSD was
47 reparameterized for use both with and without a long-range interaction
48 correction, SSD/RF and SSD/E respectively. In addition to correcting
49 the abnormally low densities, these new versions were show to maintain
50 or improve upon the transport and structural features of the original
51 water model.
52 \end{abstract}
53
54 \maketitle
55
56 %\narrowtext
57
58
59 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
60 % BODY OF TEXT
61 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
62
63 \section{Introduction}
64
65 One of the most important tasks in simulations of biochemical systems
66 is the proper depiction of water and water solvation. In fact, the
67 bulk of the calculations performed in solvated simulations are of
68 interactions with or between solvent molecules. Thus, the outcomes of
69 these types of simulations are highly dependent on the physical
70 properties of water, both as individual molecules and in
71 groups/bulk. Due to the fact that explicit solvent accounts for a
72 massive portion of the calculations, it necessary to simplify the
73 solvent to some extent in order to complete simulations in a
74 reasonable amount of time. In the case of simulating water in
75 bio-molecular studies, the balance between accurate properties and
76 computational efficiency is especially delicate, and it has resulted
77 in a variety of different water
78 models.\cite{Jorgensen83,Berendsen87,Jorgensen00} Many of these models
79 get specific properties correct or better than their predecessors, but
80 this is often at a cost of some other properties or of computer
81 time. As an example, compare TIP3P or TIP4P to TIP5P. TIP5P succeeds
82 in improving the structural and transport properties over its
83 predecessors, yet this comes at a greater than 50\% increase in
84 computational cost.\cite{Jorgensen01,Jorgensen00} One recently
85 developed model that succeeds in both retaining accuracy of system
86 properties and simplifying calculations to increase computational
87 efficiency is the Soft Sticky Dipole water model.\cite{Ichiye96}
88
89 The Soft Sticky Dipole (SSD)\ water model was developed by Ichiye
90 \emph{et al.} as a modified form of the hard-sphere water model
91 proposed by Bratko, Blum, and Luzar.\cite{Bratko85,Bratko95} SSD
92 consists of a single point dipole with a Lennard-Jones core and a
93 sticky potential that directs the particles to assume the proper
94 hydrogen bond orientation in the first solvation shell. Thus, the
95 interaction between two SSD water molecules \emph{i} and \emph{j} is
96 given by the potential
97 \begin{equation}
98 u_{ij} = u_{ij}^{LJ} (r_{ij})\ + u_{ij}^{dp}
99 (\mathbf{r}_{ij},\boldsymbol{\Omega}_i,\boldsymbol{\Omega}_j)\ +
100 u_{ij}^{sp}
101 (\mathbf{r}_{ij},\boldsymbol{\Omega}_i,\boldsymbol{\Omega}_j),
102 \end{equation}
103 where the $\mathbf{r}_{ij}$ is the position vector between molecules
104 \emph{i} and \emph{j} with magnitude equal to the distance $r_ij$, and
105 $\boldsymbol{\Omega}_i$ and $\boldsymbol{\Omega}_j$ represent the
106 orientations of the respective molecules. The Lennard-Jones, dipole,
107 and sticky parts of the potential are giving by the following
108 equations,
109 \begin{equation}
110 u_{ij}^{LJ}(r_{ij}) = 4\epsilon \left[\left(\frac{\sigma}{r_{ij}}\right)^{12}-\left(\frac{\sigma}{r_{ij}}\right)^{6}\right],
111 \end{equation}
112 \begin{equation}
113 u_{ij}^{dp} = \frac{\boldsymbol{\mu}_i\cdot\boldsymbol{\mu}_j}{r_{ij}^3}-\frac{3(\boldsymbol{\mu}_i\cdot\mathbf{r}_{ij})(\boldsymbol{\mu}_j\cdot\mathbf{r}_{ij})}{r_{ij}^5}\ ,
114 \end{equation}
115 \begin{equation}
116 \begin{split}
117 u_{ij}^{sp}
118 (\mathbf{r}_{ij},\boldsymbol{\Omega}_i,\boldsymbol{\Omega}_j)
119 &=
120 \frac{\nu_0}{2}[s(r_{ij})w(\mathbf{r}_{ij},\boldsymbol{\Omega}_i,\boldsymbol{\Omega}_j)\\
121 & \quad \ +
122 s^\prime(r_{ij})w^\prime(\mathbf{r}_{ij},\boldsymbol{\Omega}_i,\boldsymbol{\Omega}_j)]\ ,
123 \end{split}
124 \end{equation}
125 where $\boldsymbol{\mu}_i$ and $\boldsymbol{\mu}_j$ are the dipole
126 unit vectors of particles \emph{i} and \emph{j} with magnitude 2.35 D,
127 $\nu_0$ scales the strength of the overall sticky potential, $s$ and
128 $s^\prime$ are cubic switching functions. The $w$ and $w^\prime$
129 functions take the following forms,
130 \begin{equation}
131 w(\mathbf{r}_{ij},\boldsymbol{\Omega}_i,\boldsymbol{\Omega}_j)=\sin\theta_{ij}\sin2\theta_{ij}\cos2\phi_{ij},
132 \end{equation}
133 \begin{equation}
134 w^\prime(\mathbf{r}_{ij},\boldsymbol{\Omega}_i,\boldsymbol{\Omega}_j) = (\cos\theta_{ij}-0.6)^2(\cos\theta_{ij}+0.8)^2-w^0,
135 \end{equation}
136 where $w^0 = 0.07715$. The $w$ function is the tetrahedral attractive
137 term that promotes hydrogen bonding orientations within the first
138 solvation shell, and $w^\prime$ is a dipolar repulsion term that
139 repels unrealistic dipolar arrangements within the first solvation
140 shell. A more detailed description of the functional parts and
141 variables in this potential can be found in other
142 articles.\cite{Ichiye96,Ichiye99}
143
144 Being that this is a one-site point dipole model, the actual force
145 calculations are simplified significantly. In the original Monte Carlo
146 simulations using this model, Ichiye \emph{et al.} reported a
147 calculation speed up of up to an order of magnitude over other
148 comparable models while maintaining the structural behavior of
149 water.\cite{Ichiye96} In the original molecular dynamics studies, it
150 was shown that SSD improves on the prediction of many of water's
151 dynamical properties over TIP3P and SPC/E.\cite{Ichiye99} This
152 attractive combination of speed and accurate depiction of solvent
153 properties makes SSD a model of interest for the simulation of large
154 scale biological systems, such as membrane phase behavior, a specific
155 interest within our group.
156
157 One of the key limitations of this water model, however, is that it
158 has been parameterized for use with the Ewald Sum technique for the
159 handling of long-ranged interactions. When studying very large
160 systems, the Ewald summation and even particle-mesh Ewald become
161 computational burdens with their respective ideal $N^\frac{3}{2}$ and
162 $N\log N$ calculation scaling orders for $N$ particles.\cite{Darden99}
163 In applying this water model in these types of systems, it would be
164 useful to know its properties and behavior with the more
165 computationally efficient reaction field (RF) technique, and even with
166 a cutoff that lacks any form of long range correction. This study
167 addresses these issues by looking at the structural and transport
168 behavior of SSD over a variety of temperatures, with the purpose of
169 utilizing the RF correction technique. Towards the end, we suggest
170 alterations to the parameters that result in more water-like
171 behavior. It should be noted that in a recent publication, some the
172 original investigators of the SSD water model have put forth
173 adjustments to the original SSD water model to address abnormal
174 density behavior (also observed here), calling the corrected model
175 SSD1.\cite{Ichiye03} This study will consider this new model's
176 behavior as well, and hopefully improve upon its depiction of water
177 under conditions without the Ewald Sum.
178
179 \section{Methods}
180
181 As stated previously, in this study the long-range dipole-dipole
182 interactions were accounted for using the reaction field method. The
183 magnitude of the reaction field acting on dipole \emph{i} is given by
184 \begin{equation}
185 \mathcal{E}_{i} = \frac{2(\varepsilon_{s} - 1)}{2\varepsilon_{s} + 1}
186 \frac{1}{r_{c}^{3}} \sum_{j\in{\mathcal{R}}} \boldsymbol{\mu}_{j} f(r_{ij})\ ,
187 \label{rfequation}
188 \end{equation}
189 where $\mathcal{R}$ is the cavity defined by the cutoff radius
190 ($r_{c}$), $\varepsilon_{s}$ is the dielectric constant imposed on the
191 system (80 in this case), $\boldsymbol{\mu}_{j}$ is the dipole moment
192 vector of particle \emph{j}, and $f(r_{ij})$ is a cubic switching
193 function.\cite{AllenTildesley} The reaction field contribution to the
194 total energy by particle \emph{i} is given by
195 $-\frac{1}{2}\boldsymbol{\mu}_{i}\cdot\mathcal{E}_{i}$ and the torque
196 on dipole \emph{i} by
197 $\boldsymbol{\mu}_{i}\times\mathcal{E}_{i}$.\cite{AllenTildesley} Use
198 of reaction field is known to alter the orientational dynamic
199 properties, such as the dielectric relaxation time, based on changes
200 in the length of the cutoff radius.\cite{Berendsen98} This variable
201 behavior makes reaction field a less attractive method than other
202 methods, like the Ewald summation; however, for the simulation of
203 large-scale system, the computational cost benefit of reaction field
204 is dramatic. To address some of the dynamical property alterations due
205 to the use of reaction field, simulations were also performed without
206 a surrounding dielectric and suggestions are proposed on how to make
207 SSD more compatible with a reaction field.
208
209 Simulations were performed in both the isobaric-isothermal and
210 microcanonical ensembles. The constant pressure simulations were
211 implemented using an integral thermostat and barostat as outlined by
212 Hoover.\cite{Hoover85,Hoover86} All particles were treated as
213 non-linear rigid bodies. Vibrational constraints are not necessary in
214 simulations of SSD, because there are no explicit hydrogen atoms, and
215 thus no molecular vibrational modes need to be considered.
216
217 Integration of the equations of motion was carried out using the
218 symplectic splitting method proposed by Dullweber \emph{et
219 al.}.\cite{Dullweber1997} The reason for this integrator selection
220 deals with poor energy conservation of rigid body systems using
221 quaternions. While quaternions work well for orientational motion in
222 alternate ensembles, the microcanonical ensemble has a constant energy
223 requirement that is quite sensitive to errors in the equations of
224 motion. The original implementation of this code utilized quaternions
225 for rotational motion propagation; however, a detailed investigation
226 showed that they resulted in a steady drift in the total energy,
227 something that has been observed by others.\cite{Laird97}
228
229 The key difference in the integration method proposed by Dullweber
230 \emph{et al.} is that the entire rotation matrix is propagated from
231 one time step to the next. In the past, this would not have been as
232 feasible a option, being that the rotation matrix for a single body is
233 nine elements long as opposed to 3 or 4 elements for Euler angles and
234 quaternions respectively. System memory has become much less of an
235 issue in recent times, and this has resulted in substantial benefits
236 in energy conservation. There is still the issue of 5 or 6 additional
237 elements for describing the orientation of each particle, which will
238 increase dump files substantially. Simply translating the rotation
239 matrix into its component Euler angles or quaternions for storage
240 purposes relieves this burden.
241
242 The symplectic splitting method allows for Verlet style integration of
243 both linear and angular motion of rigid bodies. In the integration
244 method, the orientational propagation involves a sequence of matrix
245 evaluations to update the rotation matrix.\cite{Dullweber1997} These
246 matrix rotations end up being more costly computationally than the
247 simpler arithmetic quaternion propagation. With the same time step, a
248 1000 SSD particle simulation shows an average 7\% increase in
249 computation time using the symplectic step method in place of
250 quaternions. This cost is more than justified when comparing the
251 energy conservation of the two methods as illustrated in figure
252 \ref{timestep}.
253
254 \begin{figure}
255 \includegraphics[width=61mm, angle=-90]{timeStep.epsi}
256 \caption{Energy conservation using quaternion based integration versus
257 the symplectic step method proposed by Dullweber \emph{et al.} with
258 increasing time step. For each time step, the dotted line is total
259 energy using the symplectic step integrator, and the solid line comes
260 from the quaternion integrator. The larger time step plots are shifted
261 up from the true energy baseline for clarity.}
262 \label{timestep}
263 \end{figure}
264
265 In figure \ref{timestep}, the resulting energy drift at various time
266 steps for both the symplectic step and quaternion integration schemes
267 is compared. All of the 1000 SSD particle simulations started with the
268 same configuration, and the only difference was the method for
269 handling rotational motion. At time steps of 0.1 and 0.5 fs, both
270 methods for propagating particle rotation conserve energy fairly well,
271 with the quaternion method showing a slight energy drift over time in
272 the 0.5 fs time step simulation. At time steps of 1 and 2 fs, the
273 energy conservation benefits of the symplectic step method are clearly
274 demonstrated. Thus, while maintaining the same degree of energy
275 conservation, one can take considerably longer time steps, leading to
276 an overall reduction in computation time.
277
278 Energy drift in these SSD particle simulations was unnoticeable for
279 time steps up to three femtoseconds. A slight energy drift on the
280 order of 0.012 kcal/mol per nanosecond was observed at a time step of
281 four femtoseconds, and as expected, this drift increases dramatically
282 with increasing time step. To insure accuracy in the constant energy
283 simulations, time steps were set at 2 fs and kept at this value for
284 constant pressure simulations as well.
285
286 Ice crystals in both the $I_h$ and $I_c$ lattices were generated as
287 starting points for all the simulations. The $I_h$ crystals were
288 formed by first arranging the center of masses of the SSD particles
289 into a ``hexagonal'' ice lattice of 1024 particles. Because of the
290 crystal structure of $I_h$ ice, the simulation box assumed a
291 rectangular shape with a edge length ratio of approximately
292 1.00$\times$1.06$\times$1.23. The particles were then allowed to
293 orient freely about fixed positions with angular momenta randomized at
294 400 K for varying times. The rotational temperature was then scaled
295 down in stages to slowly cool the crystals down to 25 K. The particles
296 were then allowed translate with fixed orientations at a constant
297 pressure of 1 atm for 50 ps at 25 K. Finally, all constraints were
298 removed and the ice crystals were allowed to equilibrate for 50 ps at
299 25 K and a constant pressure of 1 atm. This procedure resulted in
300 structurally stable $I_h$ ice crystals that obey the Bernal-Fowler
301 rules\cite{Bernal33,Rahman72}. This method was also utilized in the
302 making of diamond lattice $I_c$ ice crystals, with each cubic
303 simulation box consisting of either 512 or 1000 particles. Only
304 isotropic volume fluctuations were performed under constant pressure,
305 so the ratio of edge lengths remained constant throughout the
306 simulations.
307
308 \section{Results and discussion}
309
310 Melting studies were performed on the randomized ice crystals using
311 constant pressure and temperature dynamics. By performing melting
312 simulations, the melting transition can be determined by monitoring
313 the heat capacity, in addition to determining the density maximum,
314 provided that the density maximum occurs in the liquid and not the
315 supercooled regimes. An ensemble average from five separate melting
316 simulations was acquired, each starting from different ice crystals
317 generated as described previously. All simulations were equilibrated
318 for 100 ps prior to a 200 ps data collection run at each temperature
319 setting, ranging from 25 to 400 K, with a maximum degree increment of
320 25 K. For regions of interest along this stepwise progression, the
321 temperature increment was decreased from 25 K to 10 and then 5 K. The
322 above equilibration and production times were sufficient in that the
323 system volume fluctuations dampened out in all but the very cold
324 simulations (below 225 K).
325
326 \subsection{Density Behavior}
327 In the initial average density versus temperature plot, the density
328 maximum appears between 255 and 265 K. The calculated densities within
329 this range were nearly indistinguishable, as can be seen in the zoom
330 of this region of interest, shown in figure
331 \ref{dense1}. The greater certainty of the average value at 260 K makes
332 a good argument for the actual density maximum residing at this
333 midpoint value. Figure \ref{dense1} was constructed using ice $I_h$
334 crystals for the initial configuration; and though not pictured, the
335 simulations starting from ice $I_c$ crystal configurations showed
336 similar results, with a liquid-phase density maximum in this same
337 region (between 255 and 260 K). In addition, the $I_c$ crystals are
338 more fragile than the $I_h$ crystals, leading them to deform into a
339 dense glassy state at lower temperatures. This resulted in an overall
340 low temperature density maximum at 200 K, but they still retained a
341 common liquid state density maximum with the $I_h$ simulations.
342
343 \begin{figure}
344 \includegraphics[width=65mm,angle=-90]{dense2.eps}
345 \caption{Density versus temperature for TIP4P\cite{Jorgensen98b},
346 TIP3P\cite{Jorgensen98b}, SPC/E\cite{Clancy94}, SSD without Reaction
347 Field, SSD, and Experiment\cite{CRC80}. The arrows indicate the
348 change in densities observed when turning off the reaction field. The
349 the lower than expected densities for the SSD model were what
350 prompted the original reparameterization.\cite{Ichiye03}}
351 \label{dense2}
352 \end{figure}
353
354 The density maximum for SSD actually compares quite favorably to other
355 simple water models. Figure \ref{dense2} shows a plot of these
356 findings with the density progression of several other models and
357 experiment obtained from other
358 sources.\cite{Jorgensen98b,Clancy94,CRC80} Of the listed simple water
359 models, SSD has results closest to the experimentally observed water
360 density maximum. Of the listed water models, TIP4P has a density
361 maximum behavior most like that seen in SSD. Though not shown, it is
362 useful to note that TIP5P has a water density maximum nearly identical
363 to experiment.
364
365 Possibly of more importance is the density scaling of SSD relative to
366 other common models at any given temperature (Fig. \ref{dense2}). Note
367 that the SSD model assumes a lower density than any of the other
368 listed models at the same pressure, behavior which is especially
369 apparent at temperatures greater than 300 K. Lower than expected
370 densities have been observed for other systems with the use of a
371 reaction field for long-range electrostatic interactions, so the most
372 likely reason for these significantly lower densities in these
373 simulations is the presence of the reaction field.\cite{Berendsen98}
374 In order to test the effect of the reaction field on the density of
375 the systems, the simulations were repeated for the temperature region
376 of interest without a reaction field present. The results of these
377 simulations are also displayed in figure \ref{dense2}. Without
378 reaction field, these densities increase considerably to more
379 experimentally reasonable values, especially around the freezing point
380 of liquid water. The shape of the curve is similar to the curve
381 produced from SSD simulations using reaction field, specifically the
382 rapidly decreasing densities at higher temperatures; however, a slight
383 shift in the density maximum location, down to 245 K, is
384 observed. This is probably a more accurate comparison to the other
385 listed water models in that no long range corrections were applied in
386 those simulations.\cite{Clancy94,Jorgensen98b}
387
388 It has been observed that densities are dependent on the cutoff radius
389 used for a variety of water models in simulations both with and
390 without the use of reaction field.\cite{Berendsen98} In order to
391 address the possible affect of cutoff radius, simulations were
392 performed with a dipolar cutoff radius of 12.0 \AA\ to compliment the
393 previous SSD simulations, all performed with a cutoff of 9.0 \AA. All
394 the resulting densities overlapped within error and showed no
395 significant trend in lower or higher densities as a function of cutoff
396 radius, both for simulations with and without reaction field. These
397 results indicate that there is no major benefit in choosing a longer
398 cutoff radius in simulations using SSD. This is comforting in that the
399 use of a longer cutoff radius results in a near doubling of the time
400 required to compute a single trajectory.
401
402 \subsection{Transport Behavior}
403 Of importance in these types of studies are the transport properties
404 of the particles and how they change when altering the environmental
405 conditions. In order to probe transport, constant energy simulations
406 were performed about the average density uncovered by the constant
407 pressure simulations. Simulations started with randomized velocities
408 and underwent 50 ps of temperature scaling and 50 ps of constant
409 energy equilibration before obtaining a 200 ps trajectory. Diffusion
410 constants were calculated via root-mean square deviation analysis. The
411 averaged results from 5 sets of these NVE simulations is displayed in
412 figure \ref{diffuse}, alongside experimental, SPC/E, and TIP5P
413 results.\cite{Gillen72,Mills73,Clancy94,Jorgensen01}
414
415 \begin{figure}
416 \includegraphics[width=65mm, angle=-90]{betterDiffuse.epsi}
417 \caption{Average diffusion coefficient over increasing temperature for
418 SSD, SPC/E\cite{Clancy94}, TIP5P\cite{Jorgensen01}, and Experimental
419 data from Gillen \emph{et al.}\cite{Gillen72}, and from
420 Mills\cite{Mills73}.}
421 \label{diffuse}
422 \end{figure}
423
424 The observed values for the diffusion constant point out one of the
425 strengths of the SSD model. Of the three experimental models shown,
426 the SSD model has the most accurate depiction of the diffusion trend
427 seen in experiment in both the supercooled and normal regimes. SPC/E
428 does a respectable job by getting similar values as SSD and experiment
429 around 290 K; however, it deviates at both higher and lower
430 temperatures, failing to predict the experimental trend. TIP5P and SSD
431 both start off low at the colder temperatures and tend to diffuse too
432 rapidly at the higher temperatures. This type of trend at the higher
433 temperatures is not surprising in that the densities of both TIP5P and
434 SSD are lower than experimental water at temperatures higher than room
435 temperature. When calculating the diffusion coefficients for SSD at
436 experimental densities, the resulting values fall more in line with
437 experiment at these temperatures, albeit not at standard
438 pressure. Results under these conditions can be found later in this
439 paper.
440
441 \subsection{Structural Changes and Characterization}
442 By starting the simulations from the crystalline state, the melting
443 transition and the ice structure can be studied along with the liquid
444 phase behavior beyond the melting point. To locate the melting
445 transition, the constant pressure heat capacity (C$_\text{p}$) was
446 monitored in each of the simulations. In the melting simulations of
447 the 1024 particle ice $I_h$ simulations, a large spike in C$_\text{p}$
448 occurs at 245 K, indicating a first order phase transition for the
449 melting of these ice crystals. When the reaction field is turned off,
450 the melting transition occurs at 235 K. These melting transitions are
451 considerably lower than the experimental value, but this is not
452 surprising in that SSD is a simple rigid body model with a fixed
453 dipole.
454
455 \begin{figure}
456 \includegraphics[width=85mm]{fullContours.eps}
457 \caption{Contour plots of 2D angular g($r$)'s for 512 SSD systems at
458 100 K (A \& B) and 300 K (C \& D). Contour colors are inverted for
459 clarity: dark areas signify peaks while light areas signify
460 depressions. White areas have g(\emph{r}) values below 0.5 and black
461 areas have values above 1.5.}
462 \label{contour}
463 \end{figure}
464
465 Additional analyses for understanding the melting phase-transition
466 process were performed via two-dimensional structure and dipole angle
467 correlations. Expressions for the correlations are as follows:
468
469 \begin{figure}
470 \includegraphics[width=45mm]{corrDiag.eps}
471 \caption{Two dimensional illustration of the angles involved in the
472 correlations observed in figure \ref{contour}.}
473 \label{corrAngle}
474 \end{figure}
475
476 \begin{multline}
477 g_{\text{AB}}(r,\cos\theta) = \\
478 \frac{V}{N_\text{A}N_\text{B}}\langle\sum_{i\in\text{A}}\sum_{j\in\text{B}}\delta(\cos\theta-\cos\theta_{ij})\delta(r-\left|\mathbf{r}_{ij}\right|)\rangle\ ,
479 \end{multline}
480 \begin{multline}
481 g_{\text{AB}}(r,\cos\omega) = \\
482 \frac{V}{N_\text{A}N_\text{B}}\langle\sum_{i\in\text{A}}\sum_{j\in\text{B}}\delta(\cos\omega-\cos\omega_{ij})\delta(r-\left|\mathbf{r}_{ij}\right|)\rangle\ ,
483 \end{multline}
484 where $\theta$ and $\omega$ refer to the angles shown in the above
485 illustration. By binning over both distance and the cosine of the
486 desired angle between the two dipoles, the g(\emph{r}) can be
487 dissected to determine the common dipole arrangements that constitute
488 the peaks and troughs. Frames A and B of figure \ref{contour} show a
489 relatively crystalline state of an ice $I_c$ simulation. The first
490 peak of the g(\emph{r}) primarily consists of the preferred hydrogen
491 bonding arrangements as dictated by the tetrahedral sticky potential,
492 one peak for the donating and the other for the accepting hydrogen
493 bonds. Due to the high degree of crystallinity of the sample, the
494 second and third solvation shells show a repeated peak arrangement
495 which decays at distances around the fourth solvation shell, near the
496 imposed cutoff for the Lennard-Jones and dipole-dipole interactions.
497 In the higher temperature simulation shown in frames C and D, the
498 repeated peak features are significantly blurred. The first solvation
499 shell still shows the strong effect of the sticky-potential, although
500 it covers a larger area, extending to include a fraction of aligned
501 dipole peaks within the first solvation shell. The latter peaks lose
502 definition as thermal motion and the competing dipole force overcomes
503 the sticky potential's tight tetrahedral structuring of the fluid.
504
505 This complex interplay between dipole and sticky interactions was
506 remarked upon as a possible reason for the split second peak in the
507 oxygen-oxygen g(\emph{r}).\cite{Ichiye96} At low temperatures, the
508 second solvation shell peak appears to have two distinct parts that
509 blend together to form one observable peak. At higher temperatures,
510 this split character alters to show the leading 4 \AA\ peak dominated
511 by equatorial anti-parallel dipole orientations, and there is tightly
512 bunched group of axially arranged dipoles that most likely consist of
513 the smaller fraction aligned dipole pairs. The trailing part of the
514 split peak at 5 \AA\ is dominated by aligned dipoles that range
515 primarily within the axial to the chief hydrogen bond arrangements
516 similar to those seen in the first solvation shell. This evidence
517 indicates that the dipole pair interaction begins to dominate outside
518 of the range of the dipolar repulsion term, with the primary
519 energetically favorable dipole arrangements populating the region
520 immediately outside of it's range (around 4 \AA), and arrangements
521 that seek to ideally satisfy both the sticky and dipole forces locate
522 themselves just beyond this region (around 5 \AA).
523
524 From these findings, the split second peak is primarily the product of
525 the dipolar repulsion term of the sticky potential. In fact, the
526 leading of the two peaks can be pushed out and merged with the outer
527 split peak just by extending the switching function cutoff
528 ($s^\prime(r_{ij})$) from its normal 4.0 \AA\ to values of 4.5 or even
529 5 \AA. This type of correction is not recommended for improving the
530 liquid structure, because the second solvation shell will still be
531 shifted too far out. In addition, this would have an even more
532 detrimental effect on the system densities, leading to a liquid with a
533 more open structure and a density considerably lower than the normal
534 SSD behavior shown previously. A better correction would be to include
535 the quadrupole-quadrupole interactions for the water particles outside
536 of the first solvation shell, but this reduces the simplicity and
537 speed advantage of SSD, so it is not the most desirable path to take.
538
539 \subsection{Adjusted Potentials: SSD/E and SSD/RF}
540 The propensity of SSD to adopt lower than expected densities under
541 varying conditions is troubling, especially at higher temperatures. In
542 order to correct this behavior, it's necessary to adjust the force
543 field parameters for the primary intermolecular interactions. In
544 undergoing a reparameterization, it is important not to focus on just
545 one property and neglect the other important properties. In this case,
546 it would be ideal to correct the densities while maintaining the
547 accurate transport properties.
548
549 The possible parameters for tuning include the $\sigma$ and $\epsilon$
550 Lennard-Jones parameters, the dipole strength ($\mu$), and the sticky
551 attractive and dipole repulsive terms with their respective
552 cutoffs. To alter the attractive and repulsive terms of the sticky
553 potential independently, it is necessary to separate the terms as
554 follows:
555 \begin{equation}
556 \begin{split}
557 u_{ij}^{sp}
558 (\mathbf{r}_{ij},\boldsymbol{\Omega}_i,\boldsymbol{\Omega}_j) &=
559 \frac{\nu_0}{2}[s(r_{ij})w(\mathbf{r}_{ij},\boldsymbol{\Omega}_i,\boldsymbol{\Omega}_j)]\\
560 & \quad \ + \frac{\nu_0^\prime}{2}
561 [s^\prime(r_{ij})w^\prime(\mathbf{r}_{ij},\boldsymbol{\Omega}_i,\boldsymbol{\Omega}_j)],
562 \end{split}
563 \end{equation}
564
565 where $\nu_0$ scales the strength of the tetrahedral attraction and
566 $\nu_0^\prime$ acts in an identical fashion on the dipole repulsion
567 term. For purposes of the reparameterization, the separation was
568 performed, but the final parameters were adjusted so that it is
569 unnecessary to separate the terms when implementing the adjusted water
570 potentials. The results of the reparameterizations are shown in table
571 \ref{params}. Note that both the tetrahedral attractive and dipolar
572 repulsive don't share the same lower cutoff ($r_l$) in the newly
573 parameterized potentials - soft sticky dipole enhanced (SSD/E) and
574 soft sticky dipole reaction field (SSD/RF).
575
576 \begin{table}
577 \caption{Parameters for the original and adjusted models}
578 \begin{tabular}{ l c c c c }
579 \hline \\[-3mm]
580 \ \ \ Parameters\ \ \ & \ \ \ SSD$^\dagger$ \ \ \ & \ SSD1$^\ddagger$\ \ & \ SSD/E\ \ & \ SSD/RF \\
581 \hline \\[-3mm]
582 \ \ \ $\sigma$ (\AA) & 3.051 & 3.016 & 3.035 & 3.019\\
583 \ \ \ $\epsilon$ (kcal/mol) & 0.152 & 0.152 & 0.152 & 0.152\\
584 \ \ \ $\mu$ (D) & 2.35 & 2.35 & 2.42 & 2.48\\
585 \ \ \ $\nu_0$ (kcal/mol) & 3.7284 & 3.6613 & 3.90 & 3.90\\
586 \ \ \ $r_l$ (\AA) & 2.75 & 2.75 & 2.40 & 2.40\\
587 \ \ \ $r_u$ (\AA) & 3.35 & 3.35 & 3.80 & 3.80\\
588 \ \ \ $\nu_0^\prime$ (kcal/mol) & 3.7284 & 3.6613 & 3.90 & 3.90\\
589 \ \ \ $r_l^\prime$ (\AA) & 2.75 & 2.75 & 2.75 & 2.75\\
590 \ \ \ $r_u^\prime$ (\AA) & 4.00 & 4.00 & 3.35 & 3.35\\
591 \\[-2mm]$^\dagger$ ref. \onlinecite{Ichiye96}
592 \\$^\ddagger$ ref. \onlinecite{Ichiye03}
593 \end{tabular}
594 \label{params}
595 \end{table}
596
597 \begin{figure}
598 \includegraphics[width=85mm]{GofRCompare.epsi}
599 \caption{Plots comparing experiment\cite{Head-Gordon00_1} with SSD/E
600 and SSD1 without reaction field (top), as well as SSD/RF and SSD1 with
601 reaction field turned on (bottom). The insets show the respective
602 first peaks in detail. Solid Line - experiment, dashed line - SSD/E
603 and SSD/RF, and dotted line - SSD1 (with and without reaction field).}
604 \label{grcompare}
605 \end{figure}
606
607 \begin{figure}
608 \includegraphics[width=85mm]{dualsticky.ps}
609 \caption{Isosurfaces of the sticky potential for SSD1 (left) and SSD/E \&
610 SSD/RF (right). Light areas correspond to the tetrahedral attractive
611 part, and the darker areas correspond to the dipolar repulsive part.}
612 \label{isosurface}
613 \end{figure}
614
615 In the paper detailing the development of SSD, Liu and Ichiye placed
616 particular emphasis on an accurate description of the first solvation
617 shell. This resulted in a somewhat tall and sharp first peak that
618 integrated to give similar coordination numbers to the experimental
619 data obtained by Soper and Phillips.\cite{Ichiye96,Soper86} New
620 experimental x-ray scattering data from the Head-Gordon lab indicates
621 a slightly lower and shifted first peak in the g$_\mathrm{OO}(r)$, so
622 adjustments to SSD were made while taking into consideration the new
623 experimental findings.\cite{Head-Gordon00_1} Figure \ref{grcompare}
624 shows the relocation of the first peak of the oxygen-oxygen
625 g(\emph{r}) by comparing the original SSD (with and without reaction
626 field), SSD-E, and SSD-RF to the new experimental results. Both the
627 modified water models have shorter peaks that are brought in more
628 closely to the experimental peak (as seen in the insets of figure
629 \ref{grcompare}). This structural alteration was accomplished by a
630 reduction in the Lennard-Jones $\sigma$ variable as well as adjustment
631 of the sticky potential strength and cutoffs. The cutoffs for the
632 tetrahedral attractive and dipolar repulsive terms were nearly swapped
633 with each other. Isosurfaces of the original and modified sticky
634 potentials are shown in figure \cite{isosurface}. In these
635 isosurfaces, it is easy to see how altering the cutoffs changes the
636 repulsive and attractive character of the particles. With a reduced
637 repulsive surface (the darker region), the particles can move closer
638 to one another, increasing the density for the overall system. This
639 change in interaction cutoff also results in a more gradual
640 orientational motion by allowing the particles to maintain preferred
641 dipolar arrangements before they begin to feel the pull of the
642 tetrahedral restructuring. Upon moving closer together, the dipolar
643 repulsion term becomes active and excludes the unphysical
644 arrangements. This compares with the original SSD's excluding dipolar
645 before the particles feel the pull of the ``hydrogen bonds''. Aside
646 from improving the shape of the first peak in the g(\emph{r}), this
647 improves the densities considerably by allowing the persistence of
648 full dipolar character below the previous 4.0 \AA\ cutoff.
649
650 While adjusting the location and shape of the first peak of
651 g(\emph{r}) improves the densities to some degree, these changes alone
652 are insufficient to bring the system densities up to the values
653 observed experimentally. To finish bringing up the densities, the
654 dipole moments were increased in both the adjusted models. Being a
655 dipole based model, the structure and transport are very sensitive to
656 changes in the dipole moment. The original SSD simply used the dipole
657 moment calculated from the TIP3P water model, which at 2.35 D is
658 significantly greater than the experimental gas phase value of 1.84
659 D. The larger dipole moment is a more realistic value and improve the
660 dielectric properties of the fluid. Both theoretical and experimental
661 measurements indicate a liquid phase dipole moment ranging from 2.4 D
662 to values as high as 3.11 D, so there is quite a range available for
663 adjusting the dipole
664 moment.\cite{Sprik91,Kusalik02,Badyal00,Barriol64} The increasing of
665 the dipole moments to 2.418 and 2.48 D for SSD/E and SSD/RF
666 respectively is moderate in the range of the experimental values;
667 however, it leads to significant changes in the density and transport
668 of the water models.
669
670 In order to demonstrate the benefits of this reparameterization, a
671 series of NPT and NVE simulations were performed to probe the density
672 and transport properties of the adapted models and compare the results
673 to the original SSD model. This comparison involved full NPT melting
674 sequences for both SSD/E and SSD/RF, as well as NVE transport
675 calculations at both self-consistent and experimental
676 densities. Again, the results come from five separate simulations of
677 1024 particle systems, and the melting sequences were started from
678 different ice $I_h$ crystals constructed as stated previously. Like
679 before, all of the NPT simulations were equilibrated for 100 ps before
680 a 200 ps data collection run, and they used the previous temperature's
681 final configuration as a starting point. All of the NVE simulations
682 had the same thermalization, equilibration, and data collection times
683 stated earlier in this paper.
684
685 \begin{figure}
686 \includegraphics[width=62mm, angle=-90]{ssdeDense.epsi}
687 \caption{Comparison of densities calculated with SSD/E to SSD without a
688 reaction field, TIP3P\cite{Jorgensen98b}, TIP5P\cite{Jorgensen00},
689 SPC/E\cite{Clancy94}, and Experiment\cite{CRC80}. The window shows a
690 expansion around 300 K with error bars included to clarify this region
691 of interest. Note that both SSD1 and SSD/E show good agreement with
692 experiment when the long-range correction is neglected.}
693 \label{ssdedense}
694 \end{figure}
695
696 Figure \ref{ssdedense} shows the density profile for the SSD/E water
697 model in comparison to the original SSD without a reaction field,
698 experiment, and the other common water models considered
699 previously. The calculated densities have increased significantly over
700 the original SSD model and match the experimental value just below 298
701 K. At 298 K, the density of SSD/E is 0.995$\pm$0.001 g/cm$^3$, which
702 compares well with the experimental value of 0.997 g/cm$^3$ and is
703 considerably better than the SSD value of 0.967$\pm$0.003
704 g/cm$^3$. The increased dipole moment in SSD/E has helped to flatten
705 out the curve at higher temperatures, only the improvement is marginal
706 at best. This steep drop in densities is due to the dipolar rather
707 than charge based interactions which decay more rapidly at longer
708 distances.
709
710 By monitoring C$\text{p}$ throughout these simulations, the melting
711 transition for SSD/E was observed at 230 K, about 5 degrees lower than
712 SSD. The resulting density maximum is located at 240 K, again about 5
713 degrees lower than the SSD value of 245 K. Though there is a decrease
714 in both of these values, the corrected densities near room temperature
715 justify the modifications taken.
716
717 \begin{figure}
718 \includegraphics[width=62mm, angle=-90]{ssdrfDense.epsi}
719 \caption{Comparison of densities calculated with SSD/RF to SSD with a
720 reaction field, TIP3P\cite{Jorgensen98b}, TIP5P\cite{Jorgensen00},
721 SPC/E\cite{Clancy94}, and Experiment\cite{CRC80}. The inset shows the
722 necessity of reparameterization when utilizing a reaction field
723 long-ranged correction - SSD/RF provides significantly more accurate
724 densities than SSD1 when performing room temperature simulations.}
725 \label{ssdrfdense}
726 \end{figure}
727
728 Figure \ref{ssdrfdense} shows a density comparison between SSD/RF and
729 SSD with an active reaction field. Like in the simulations of SSD/E,
730 the densities show a dramatic increase over normal SSD. At 298 K,
731 SSD/RF has a density of 0.997$\pm$0.001 g/cm$^3$, right in line with
732 experiment and considerably better than the SSD value of
733 0.941$\pm$0.001 g/cm$^3$. The melting point is observed at 240 K,
734 which is 5 degrees lower than SSD with a reaction field, and the
735 density maximum at 255 K, again 5 degrees lower than SSD. The density
736 at higher temperature still drops off more rapidly than the charge
737 based models but is in better agreement than SSD/E.
738
739 The reparameterization of the SSD water model, both for use with and
740 without an applied long-range correction, brought the densities up to
741 what is expected for simulating liquid water. In addition to improving
742 the densities, it is important that particle transport be maintained
743 or improved. Figure \ref{ssdediffuse} compares the temperature
744 dependence of the diffusion constant of SSD/E to SSD without an active
745 reaction field, both at the densities calculated at 1 atm and at the
746 experimentally calculated densities for super-cooled and liquid
747 water. In the upper plot, the diffusion constant for SSD/E is
748 consistently a little faster than experiment, while SSD starts off
749 slower than experiment and crosses to merge with SSD/E at high
750 temperatures. Both models follow the experimental trend well, but
751 diffuse too rapidly at higher temperatures. This abnormally fast
752 diffusion is caused by the decreased system density. Since the
753 densities of SSD/E don't deviate as much from experiment as those of
754 SSD, it follows the experimental trend more closely. This observation
755 is backed up by looking at the lower plot. The diffusion constants for
756 SSD/E track with the experimental values while SSD deviates on the low
757 side of the trend with increasing temperature. This is again a product
758 of SSD/E having densities closer to experiment, and not deviating to
759 lower densities with increasing temperature as rapidly.
760
761 \begin{figure}
762 \includegraphics[width=65mm, angle=-90]{ssdrfDiffuse.epsi}
763 \caption{Plots of the diffusion constants calculated from SSD/RF and SSD1,
764 both with an active reaction field, along with experimental results
765 from Gillen \emph{et al.}\cite{Gillen72} and Mills\cite{Mills73}. The
766 NVE calculations were performed at the average densities observed in
767 the 1 atm NPT simulations for both of the models. Note how accurately
768 SSD/RF simulates the diffusion of water throughout this temperature
769 range. The more rapidly increasing diffusion constants at high
770 temperatures for both models is attributed to the significantly lower
771 densities than observed in experiment.}
772 \label{ssdrfdiffuse}
773 \end{figure}
774
775 \begin{figure}
776 \includegraphics[width=65mm, angle=-90]{ssdeDiffuse.epsi}
777 \caption{Plots of the diffusion constants calculated from SSD/E and SSD1,
778 both without a reaction field, along with experimental results are
779 from Gillen \emph{et al.}\cite{Gillen72} and Mills\cite{Mills73}. The
780 NVE calculations were performed at the average densities observed in
781 the 1 atm NPT simulations for the respective models. SSD/E is
782 slightly more fluid than experiment at all of the temperatures, but
783 it is closer than SSD1 without a long-range correction.}
784 \label{ssdediffuse}
785 \end{figure}
786
787 In figure \ref{ssdrfdiffuse}, the diffusion constants for SSD/RF are
788 compared with SSD with an active reaction field. In the upper plot,
789 SSD/RF tracks with the experimental results incredibly well, identical
790 within error throughout the temperature range and only showing a
791 slight increasing trend at higher temperatures. SSD also tracks
792 experiment well, only it tends to diffuse a little more slowly at low
793 temperatures and deviates to diffuse too rapidly at high
794 temperatures. As was stated in the SSD/E comparisons, this deviation
795 away from the ideal trend is due to a rapid decrease in density at
796 higher temperatures. SSD/RF doesn't suffer from this problem as much
797 as SSD, because the calculated densities are more true to
798 experiment. This is again emphasized in the lower plot, where SSD/RF
799 tracks the experimental diffusion exactly while SSD's diffusion
800 constants are slightly too low due to its need for a lower density at
801 the specified temperature.
802
803 \subsection{Additional Observations}
804
805 While performing the melting sequences of SSD/E, some interesting
806 observations were made. After melting at 230 K, two of the systems
807 underwent crystallization events near 245 K. As the heating process
808 continued, the two systems remained crystalline until finally melting
809 between 320 and 330 K. These simulations were excluded from the data
810 set shown in figure \ref{ssdedense} and replaced with two additional
811 melting sequences that did not undergo this anomalous phase
812 transition, while this crystallization event was investigated
813 separately.
814
815 \begin{figure}
816 \includegraphics[width=85mm]{povIce.ps}
817 \caption{Crystal structure of an ice 0 lattice shown from the (001) face.}
818 \label{weirdice}
819 \end{figure}
820
821 The final configurations of these two melting sequences shows an
822 expanded zeolite-like crystal structure that does not correspond to
823 any known form of ice. For convenience and to help distinguish it from
824 the experimentally observed forms of ice, this crystal structure will
825 henceforth be referred to as ice-zero (ice 0). The crystallinity was
826 extensive enough than a near ideal crystal structure could be
827 obtained. Figure \ref{weirdice} shows the repeating crystal structure
828 of a typical crystal at 5 K. The unit cell contains eight molecules,
829 and figure \ref{unitcell} shows a unit cell built from the water
830 particle center of masses that can be used to construct a repeating
831 lattice, similar to figure \ref{weirdice}. Each molecule is hydrogen
832 bonded to four other water molecules; however, the hydrogen bonds are
833 flexed rather than perfectly straight. This results in a skewed
834 tetrahedral geometry about the central molecule. Looking back at
835 figure \ref{isosurface}, it is easy to see how these flexed hydrogen
836 bonds are allowed in that the attractive regions are conical in shape,
837 with the greatest attraction in the central region. Though not ideal,
838 these flexed hydrogen bonds are favorable enough to stabilize an
839 entire crystal generated around them. In fact, the imperfect ice 0
840 crystals were so stable that they melted at greater than room
841 temperature.
842
843 \begin{figure}
844 \includegraphics[width=65mm]{ice0cell.eps}
845 \caption{Simple unit cell for constructing ice 0. In this cell, $c$ is
846 equal to $0.4714\times a$, and a typical value for $a$ is 8.25 \AA.}
847 \label{unitcell}
848 \end{figure}
849
850 The initial simulations indicated that ice 0 is the preferred ice
851 structure for at least SSD/E. To verify this, a comparison was made
852 between near ideal crystals of ice $I_h$, ice $I_c$, and ice 0 at
853 constant pressure with SSD/E, SSD/RF, and SSD. Near ideal versions of
854 the three types of crystals were cooled to ~1 K, and the potential
855 energies of each were compared using all three water models. With
856 every water model, ice 0 turned out to have the lowest potential
857 energy: 5\% lower than $I_h$ with SSD, 6.5\% lower with SSD/E, and
858 7.5\% lower with SSD/RF. In all three of these water models, ice $I_c$
859 was observed to be ~2\% less stable than ice $I_h$. In addition to
860 having the lowest potential energy, ice 0 was the most expanded of the
861 three ice crystals, ~5\% less dense than ice $I_h$ with all of the
862 water models. In all three water models, ice $I_c$ was observed to be
863 ~2\% more dense than ice $I_h$.
864
865 In addition to the low temperature comparisons, melting sequences were
866 performed with ice 0 as the initial configuration using SSD/E, SSD/RF,
867 and SSD both with and without a reaction field. The melting
868 transitions for both SSD/E and SSD without a reaction field occurred
869 at temperature in excess of 375 K. SSD/RF and SSD with a reaction
870 field had more reasonable melting transitions, down near 325 K. These
871 melting point observations emphasize how preferred this crystal
872 structure is over the most common types of ice when using these single
873 point water models.
874
875 Recognizing that the above tests show ice 0 to be both the most stable
876 and lowest density crystal structure for these single point water
877 models, it is interesting to speculate on the favorability of this
878 crystal structure with the different charge based models. As a quick
879 test, these 3 crystal types were converted from SSD type particles to
880 TIP3P waters and read into CHARMM.\cite{Karplus83} Identical energy
881 minimizations were performed on all of these crystals to compare the
882 system energies. Again, ice 0 was observed to have the lowest total
883 system energy. The total energy of ice 0 was ~2\% lower than ice
884 $I_h$, which was in turn ~3\% lower than ice $I_c$. From these initial
885 results, we would not be surprised if results from the other common
886 water models show ice 0 to be the lowest energy crystal structure. A
887 continuation on work studing ice 0 with multipoint water models will
888 be published in a coming article.
889
890 \section{Conclusions}
891 The density maximum and temperature dependent transport for the SSD
892 water model, both with and without the use of reaction field, were
893 studied via a series of NPT and NVE simulations. The constant pressure
894 simulations of the melting of both $I_h$ and $I_c$ ice showed a
895 density maximum near 260 K. In most cases, the calculated densities
896 were significantly lower than the densities calculated in simulations
897 of other water models. Analysis of particle diffusion showed SSD to
898 capture the transport properties of experimental very well in both the
899 normal and super-cooled liquid regimes. In order to correct the
900 density behavior, SSD was reparameterized for use both with and
901 without a long-range interaction correction, SSD/RF and SSD/E
902 respectively. In addition to correcting the abnormally low densities,
903 these new versions were show to maintain or improve upon the transport
904 and structural features of the original water model, all while
905 maintaining the fast performance of the SSD water model. This work
906 shows these simple water models, and in particular SSD/E and SSD/RF,
907 to be excellent choices to represent explicit water in future
908 simulations of biochemical systems.
909
910 \section{Acknowledgments}
911 Support for this project was provided by the National Science
912 Foundation under grant CHE-0134881. Computation time was provided by
913 the Notre Dame Bunch-of-Boxes (B.o.B) computer cluster under NSF grant
914 DMR 00 79647.
915
916 \bibliographystyle{jcp}
917
918 \bibliography{nptSSD}
919
920 %\pagebreak
921
922 \end{document}