ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/ssdePaper/nptSSD.tex
Revision: 861
Committed: Wed Nov 12 13:37:15 2003 UTC (20 years, 7 months ago) by chrisfen
Content type: application/x-tex
File size: 50011 byte(s)
Log Message:
Massive changes to content. Declared as first draft.

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 water very well in both the normal and super-cooled
46 liquid 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. Compared to the density
49 corrected version of SSD (SSD1), these modified models were shown to
50 maintain or improve upon the structural and transport properties.
51 \end{abstract}
52
53 \maketitle
54
55 %\narrowtext
56
57
58 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
59 % BODY OF TEXT
60 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
61
62 \section{Introduction}
63
64 One of the most important tasks in simulations of biochemical systems
65 is the proper depiction of water and water solvation. In fact, the
66 bulk of the calculations performed in solvated simulations are of
67 interactions with or between solvent molecules. Thus, the outcomes of
68 these types of simulations are highly dependent on the physical
69 properties of water, both as individual molecules and in
70 groups/bulk. Due to the fact that explicit solvent accounts for a
71 massive portion of the calculations, it necessary to simplify the
72 solvent to some extent in order to complete simulations in a
73 reasonable amount of time. In the case of simulating water in
74 bio-molecular studies, the balance between accurate properties and
75 computational efficiency is especially delicate, and it has resulted
76 in a variety of different water
77 models.\cite{Jorgensen83,Berendsen87,Jorgensen00} Many of these models
78 get specific properties correct or better than their predecessors, but
79 this is often at a cost of some other properties or of computer
80 time. As an example, compare TIP3P or TIP4P to TIP5P. TIP5P succeeds
81 in improving the structural and transport properties over its
82 predecessors, yet this comes at a greater than 50\% increase in
83 computational cost.\cite{Jorgensen01,Jorgensen00} One recently
84 developed model that succeeds in both retaining accuracy of system
85 properties and simplifying calculations to increase computational
86 efficiency is the Soft Sticky Dipole water model.\cite{Ichiye96}
87
88 The Soft Sticky Dipole (SSD)\ water model was developed by Ichiye
89 \emph{et al.} as a modified form of the hard-sphere water model
90 proposed by Bratko, Blum, and Luzar.\cite{Bratko85,Bratko95} SSD
91 consists of a single point dipole with a Lennard-Jones core and a
92 sticky potential that directs the particles to assume the proper
93 hydrogen bond orientation in the first solvation shell. Thus, the
94 interaction between two SSD water molecules \emph{i} and \emph{j} is
95 given by the potential
96 \begin{equation}
97 u_{ij} = u_{ij}^{LJ} (r_{ij})\ + u_{ij}^{dp}
98 (\mathbf{r}_{ij},\boldsymbol{\Omega}_i,\boldsymbol{\Omega}_j)\ +
99 u_{ij}^{sp}
100 (\mathbf{r}_{ij},\boldsymbol{\Omega}_i,\boldsymbol{\Omega}_j),
101 \end{equation}
102 where the $\mathbf{r}_{ij}$ is the position vector between molecules
103 \emph{i} and \emph{j} with magnitude equal to the distance $r_ij$, and
104 $\boldsymbol{\Omega}_i$ and $\boldsymbol{\Omega}_j$ represent the
105 orientations of the respective molecules. The Lennard-Jones, dipole,
106 and sticky parts of the potential are giving by the following
107 equations,
108 \begin{equation}
109 u_{ij}^{LJ}(r_{ij}) = 4\epsilon \left[\left(\frac{\sigma}{r_{ij}}\right)^{12}-\left(\frac{\sigma}{r_{ij}}\right)^{6}\right],
110 \end{equation}
111 \begin{equation}
112 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}\ ,
113 \end{equation}
114 \begin{equation}
115 \begin{split}
116 u_{ij}^{sp}
117 (\mathbf{r}_{ij},\boldsymbol{\Omega}_i,\boldsymbol{\Omega}_j)
118 &=
119 \frac{\nu_0}{2}[s(r_{ij})w(\mathbf{r}_{ij},\boldsymbol{\Omega}_i,\boldsymbol{\Omega}_j)\\
120 & \quad \ +
121 s^\prime(r_{ij})w^\prime(\mathbf{r}_{ij},\boldsymbol{\Omega}_i,\boldsymbol{\Omega}_j)]\ ,
122 \end{split}
123 \end{equation}
124 where $\boldsymbol{\mu}_i$ and $\boldsymbol{\mu}_j$ are the dipole
125 unit vectors of particles \emph{i} and \emph{j} with magnitude 2.35 D,
126 $\nu_0$ scales the strength of the overall sticky potential, $s$ and
127 $s^\prime$ are cubic switching functions. The $w$ and $w^\prime$
128 functions take the following forms,
129 \begin{equation}
130 w(\mathbf{r}_{ij},\boldsymbol{\Omega}_i,\boldsymbol{\Omega}_j)=\sin\theta_{ij}\sin2\theta_{ij}\cos2\phi_{ij},
131 \end{equation}
132 \begin{equation}
133 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,
134 \end{equation}
135 where $w^0 = 0.07715$. The $w$ function is the tetrahedral attractive
136 term that promotes hydrogen bonding orientations within the first
137 solvation shell, and $w^\prime$ is a dipolar repulsion term that
138 repels unrealistic dipolar arrangements within the first solvation
139 shell. A more detailed description of the functional parts and
140 variables in this potential can be found in other
141 articles.\cite{Ichiye96,Ichiye99}
142
143 Being that this is a one-site point dipole model, the actual force
144 calculations are simplified significantly. In the original Monte Carlo
145 simulations using this model, Ichiye \emph{et al.} reported a
146 calculation speed up of up to an order of magnitude over other
147 comparable models, while maintaining the structural behavior of
148 water.\cite{Ichiye96} In the original molecular dynamics studies, it
149 was shown that SSD improves on the prediction of many of water's
150 dynamical properties over TIP3P and SPC/E.\cite{Ichiye99} This
151 attractive combination of speed and accurate depiction of solvent
152 properties makes SSD a model of interest for the simulation of large
153 scale biological systems, such as membrane phase behavior.
154
155 One of the key limitations of this water model, however, is that it
156 has been parameterized for use with the Ewald Sum technique for the
157 handling of long-ranged interactions. When studying very large
158 systems, the Ewald summation and even particle-mesh Ewald become
159 computational burdens with their respective ideal $N^\frac{3}{2}$ and
160 $N\log N$ calculation scaling orders for $N$ particles.\cite{Darden99}
161 In applying this water model in these types of systems, it would be
162 useful to know its properties and behavior with the more
163 computationally efficient reaction field (RF) technique, and even with
164 a cutoff that lacks any form of long range correction. This study
165 addresses these issues by looking at the structural and transport
166 behavior of SSD over a variety of temperatures, with the purpose of
167 utilizing the RF correction technique. Toward the end, we suggest
168 alterations to the parameters that result in more water-like
169 behavior. It should be noted that in a recent publication, some the
170 original investigators of the SSD water model have put forth
171 adjustments to the SSD water model to address abnormal density
172 behavior (also observed here), calling the corrected model
173 SSD1.\cite{Ichiye03} This study will make comparisons with this new
174 model's behavior with the goal of improving upon the depiction of
175 water under conditions without the Ewald Sum.
176
177 \section{Methods}
178
179 As stated previously, in this study the long-range dipole-dipole
180 interactions were accounted for using the reaction field method. The
181 magnitude of the reaction field acting on dipole \emph{i} is given by
182 \begin{equation}
183 \mathcal{E}_{i} = \frac{2(\varepsilon_{s} - 1)}{2\varepsilon_{s} + 1}
184 \frac{1}{r_{c}^{3}} \sum_{j\in{\mathcal{R}}} \boldsymbol{\mu}_{j} f(r_{ij})\ ,
185 \label{rfequation}
186 \end{equation}
187 where $\mathcal{R}$ is the cavity defined by the cutoff radius
188 ($r_{c}$), $\varepsilon_{s}$ is the dielectric constant imposed on the
189 system (80 in this case), $\boldsymbol{\mu}_{j}$ is the dipole moment
190 vector of particle \emph{j}, and $f(r_{ij})$ is a cubic switching
191 function.\cite{AllenTildesley} The reaction field contribution to the
192 total energy by particle \emph{i} is given by
193 $-\frac{1}{2}\boldsymbol{\mu}_{i}\cdot\mathcal{E}_{i}$ and the torque
194 on dipole \emph{i} by
195 $\boldsymbol{\mu}_{i}\times\mathcal{E}_{i}$.\cite{AllenTildesley} Use
196 of reaction field is known to alter the orientational dynamic
197 properties, such as the dielectric relaxation time, based on changes
198 in the length of the cutoff radius.\cite{Berendsen98} This variable
199 behavior makes reaction field a less attractive method than other
200 methods, like the Ewald summation; however, for the simulation of
201 large-scale system, the computational cost benefit of reaction field
202 is dramatic. To address some of the dynamical property alterations due
203 to the use of reaction field, simulations were also performed without
204 a surrounding dielectric and suggestions are proposed on how to make
205 SSD more accurate both with and without a reaction field.
206
207 Simulations were performed in both the isobaric-isothermal and
208 microcanonical ensembles. The constant pressure simulations were
209 implemented using an integral thermostat and barostat as outlined by
210 Hoover.\cite{Hoover85,Hoover86} All particles were treated as
211 non-linear rigid bodies. Vibrational constraints are not necessary in
212 simulations of SSD, because there are no explicit hydrogen atoms, and
213 thus no molecular vibrational modes need to be considered.
214
215 Integration of the equations of motion was carried out using the
216 symplectic splitting method proposed by Dullweber \emph{et
217 al.}.\cite{Dullweber1997} The reason for this integrator selection
218 deals with poor energy conservation of rigid body systems using
219 quaternions. While quaternions work well for orientational motion in
220 alternate ensembles, the microcanonical ensemble has a constant energy
221 requirement that is quite sensitive to errors in the equations of
222 motion. The original implementation of this code utilized quaternions
223 for rotational motion propagation; however, a detailed investigation
224 showed that they resulted in a steady drift in the total energy,
225 something that has been observed by others.\cite{Laird97}
226
227 The key difference in the integration method proposed by Dullweber
228 \emph{et al.} is that the entire rotation matrix is propagated from
229 one time step to the next. In the past, this would not have been as
230 feasible a option, being that the rotation matrix for a single body is
231 nine elements long as opposed to 3 or 4 elements for Euler angles and
232 quaternions respectively. System memory has become much less of an
233 issue in recent times, and this has resulted in substantial benefits
234 in energy conservation. There is still the issue of 5 or 6 additional
235 elements for describing the orientation of each particle, which will
236 increase dump files substantially. Simply translating the rotation
237 matrix into its component Euler angles or quaternions for storage
238 purposes relieves this burden.
239
240 The symplectic splitting method allows for Verlet style integration of
241 both linear and angular motion of rigid bodies. In the integration
242 method, the orientational propagation involves a sequence of matrix
243 evaluations to update the rotation matrix.\cite{Dullweber1997} These
244 matrix rotations end up being more costly computationally than the
245 simpler arithmetic quaternion propagation. With the same time step, a
246 1000 SSD particle simulation shows an average 7\% increase in
247 computation time using the symplectic step method in place of
248 quaternions. This cost is more than justified when comparing the
249 energy conservation of the two methods as illustrated in figure
250 \ref{timestep}.
251
252 \begin{figure}
253 \includegraphics[width=61mm, angle=-90]{timeStep.epsi}
254 \caption{Energy conservation using quaternion based integration versus
255 the symplectic step method proposed by Dullweber \emph{et al.} with
256 increasing time step. For each time step, the dotted line is total
257 energy using the symplectic step integrator, and the solid line comes
258 from the quaternion integrator. The larger time step plots are shifted
259 up from the true energy baseline for clarity.}
260 \label{timestep}
261 \end{figure}
262
263 In figure \ref{timestep}, the resulting energy drift at various time
264 steps for both the symplectic step and quaternion integration schemes
265 is compared. All of the 1000 SSD particle simulations started with the
266 same configuration, and the only difference was the method for
267 handling rotational motion. At time steps of 0.1 and 0.5 fs, both
268 methods for propagating particle rotation conserve energy fairly well,
269 with the quaternion method showing a slight energy drift over time in
270 the 0.5 fs time step simulation. At time steps of 1 and 2 fs, the
271 energy conservation benefits of the symplectic step method are clearly
272 demonstrated. Thus, while maintaining the same degree of energy
273 conservation, one can take considerably longer time steps, leading to
274 an overall reduction in computation time.
275
276 Energy drift in these SSD particle simulations was unnoticeable for
277 time steps up to three femtoseconds. A slight energy drift on the
278 order of 0.012 kcal/mol per nanosecond was observed at a time step of
279 four femtoseconds, and as expected, this drift increases dramatically
280 with increasing time step. To insure accuracy in the constant energy
281 simulations, time steps were set at 2 fs and kept at this value for
282 constant pressure simulations as well.
283
284 Ice crystals in both the $I_h$ and $I_c$ lattices were generated as
285 starting points for all the simulations. The $I_h$ crystals were
286 formed by first arranging the center of masses of the SSD particles
287 into a ``hexagonal'' ice lattice of 1024 particles. Because of the
288 crystal structure of $I_h$ ice, the simulation box assumed a
289 rectangular shape with a edge length ratio of approximately
290 1.00$\times$1.06$\times$1.23. The particles were then allowed to
291 orient freely about fixed positions with angular momenta randomized at
292 400 K for varying times. The rotational temperature was then scaled
293 down in stages to slowly cool the crystals down to 25 K. The particles
294 were then allowed translate with fixed orientations at a constant
295 pressure of 1 atm for 50 ps at 25 K. Finally, all constraints were
296 removed and the ice crystals were allowed to equilibrate for 50 ps at
297 25 K and a constant pressure of 1 atm. This procedure resulted in
298 structurally stable $I_h$ ice crystals that obey the Bernal-Fowler
299 rules\cite{Bernal33,Rahman72}. This method was also utilized in the
300 making of diamond lattice $I_c$ ice crystals, with each cubic
301 simulation box consisting of either 512 or 1000 particles. Only
302 isotropic volume fluctuations were performed under constant pressure,
303 so the ratio of edge lengths remained constant throughout the
304 simulations.
305
306 \section{Results and discussion}
307
308 Melting studies were performed on the randomized ice crystals using
309 constant pressure and temperature dynamics. By performing melting
310 simulations, the melting transition can be determined by monitoring
311 the heat capacity, in addition to determining the density maximum -
312 provided that the density maximum occurs in the liquid and not the
313 supercooled regime. An ensemble average from five separate melting
314 simulations was acquired, each starting from different ice crystals
315 generated as described previously. All simulations were equilibrated
316 for 100 ps prior to a 200 ps data collection run at each temperature
317 setting. The temperature range of study spanned from 25 to 400 K, with
318 a maximum degree increment of 25 K. For regions of interest along this
319 stepwise progression, the temperature increment was decreased from 25
320 K to 10 and 5 K. The above equilibration and production times were
321 sufficient in that the system volume fluctuations dampened out in all
322 but the very cold simulations (below 225 K).
323
324 \subsection{Density Behavior}
325 Initial simulations focused on the original SSD water model, and an
326 average density versus temperature plot is shown in figure
327 \ref{dense1}. Note that the density maximum when using a reaction
328 field appears between 255 and 265 K, where the calculated densities
329 within this range were nearly indistinguishable. The greater certainty
330 of the average value at 260 K makes a good argument for the actual
331 density maximum residing at this midpoint value. Figure \ref{dense1}
332 was constructed using ice $I_h$ crystals for the initial
333 configuration; and though not pictured, the simulations starting from
334 ice $I_c$ crystal configurations showed similar results, with a
335 liquid-phase density maximum in this same region (between 255 and 260
336 K). In addition, the $I_c$ crystals are more fragile than the $I_h$
337 crystals, leading them to deform into a dense glassy state at lower
338 temperatures. This resulted in an overall low temperature density
339 maximum at 200 K, but they still retained a common liquid state
340 density maximum with the $I_h$ simulations.
341
342 \begin{figure}
343 \includegraphics[width=65mm,angle=-90]{dense2.eps}
344 \caption{Density versus temperature for TIP4P\cite{Jorgensen98b},
345 TIP3P\cite{Jorgensen98b}, SPC/E\cite{Clancy94}, SSD without Reaction
346 Field, SSD, and Experiment\cite{CRC80}. The arrows indicate the
347 change in densities observed when turning off the reaction field. The
348 the lower than expected densities for the SSD model were what
349 prompted the original reparameterization.\cite{Ichiye03}}
350 \label{dense1}
351 \end{figure}
352
353 The density maximum for SSD actually compares quite favorably to other
354 simple water models. Figure \ref{dense1} also shows calculated
355 densities of several other models and experiment obtained from other
356 sources.\cite{Jorgensen98b,Clancy94,CRC80} Of the listed simple water
357 models, SSD has results closest to the experimentally observed water
358 density maximum. Of the listed water models, TIP4P has a density
359 maximum behavior most like that seen in SSD. Though not included in
360 this particular plot, it is useful to note that TIP5P has a water
361 density maximum nearly identical to experiment.
362
363 It has been observed that densities are dependent on the cutoff radius
364 used for a variety of water models in simulations both with and
365 without the use of reaction field.\cite{Berendsen98} In order to
366 address the possible affect of cutoff radius, simulations were
367 performed with a dipolar cutoff radius of 12.0 \AA\ to compliment the
368 previous SSD simulations, all performed with a cutoff of 9.0 \AA. All
369 the resulting densities overlapped within error and showed no
370 significant trend in lower or higher densities as a function of cutoff
371 radius, both for simulations with and without reaction field. These
372 results indicate that there is no major benefit in choosing a longer
373 cutoff radius in simulations using SSD. This is comforting in that the
374 use of a longer cutoff radius results in significant increases in the
375 time required to obtain a single trajectory.
376
377 The most important thing to recognize in figure \ref{dense1} is the
378 density scaling of SSD relative to other common models at any given
379 temperature. Note that the SSD model assumes a lower density than any
380 of the other listed models at the same pressure, behavior which is
381 especially apparent at temperatures greater than 300 K. Lower than
382 expected densities have been observed for other systems with the use
383 of a reaction field for long-range electrostatic interactions, so the
384 most likely reason for these significantly lower densities in these
385 simulations is the presence of the reaction
386 field.\cite{Berendsen98,Nezbeda02} In order to test the effect of the
387 reaction field on the density of the systems, the simulations were
388 repeated without a reaction field present. The results of these
389 simulations are also displayed in figure \ref{dense1}. Without
390 reaction field, these densities increase considerably to more
391 experimentally reasonable values, especially around the freezing point
392 of liquid water. The shape of the curve is similar to the curve
393 produced from SSD simulations using reaction field, specifically the
394 rapidly decreasing densities at higher temperatures; however, a shift
395 in the density maximum location, down to 245 K, is observed. This is
396 probably a more accurate comparison to the other listed water models,
397 in that no long range corrections were applied in those
398 simulations.\cite{Clancy94,Jorgensen98b} However, even without a
399 reaction field, the density around 300 K is still significantly lower
400 than experiment and comparable water models. This anomalous behavior
401 was what lead Ichiye \emph{et al.} to recently reparameterize SSD and
402 make SSD1.\cite{Ichiye03} In discussing potential adjustments later in
403 this paper, all comparisons were performed with this new model.
404
405 \subsection{Transport Behavior}
406 Of importance in these types of studies are the transport properties
407 of the particles and how they change when altering the environmental
408 conditions. In order to probe transport, constant energy simulations
409 were performed about the average density uncovered by the constant
410 pressure simulations. Simulations started with randomized velocities
411 and underwent 50 ps of temperature scaling and 50 ps of constant
412 energy equilibration before obtaining a 200 ps trajectory. Diffusion
413 constants were calculated via root-mean square deviation analysis. The
414 averaged results from 5 sets of these NVE simulations is displayed in
415 figure \ref{diffuse}, alongside experimental, SPC/E, and TIP5P
416 results.\cite{Gillen72,Mills73,Clancy94,Jorgensen01}
417
418 \begin{figure}
419 \includegraphics[width=65mm, angle=-90]{betterDiffuse.epsi}
420 \caption{Average diffusion coefficient over increasing temperature for
421 SSD, SPC/E\cite{Clancy94}, TIP5P\cite{Jorgensen01}, and Experimental
422 data from Gillen \emph{et al.}\cite{Gillen72}, and from
423 Mills\cite{Mills73}.}
424 \label{diffuse}
425 \end{figure}
426
427 The observed values for the diffusion constant point out one of the
428 strengths of the SSD model. Of the three experimental models shown,
429 the SSD model has the most accurate depiction of the diffusion trend
430 seen in experiment in both the supercooled and normal regimes. SPC/E
431 does a respectable job by getting similar values as SSD and experiment
432 around 290 K; however, it deviates at both higher and lower
433 temperatures, failing to predict the experimental trend. TIP5P and SSD
434 both start off low at the colder temperatures and tend to diffuse too
435 rapidly at the higher temperatures. This type of trend at the higher
436 temperatures is not surprising in that the densities of both TIP5P and
437 SSD are lower than experimental water at temperatures higher than room
438 temperature. When calculating the diffusion coefficients for SSD at
439 experimental densities, the resulting values fall more in line with
440 experiment at these temperatures, albeit not at standard pressure.
441
442 \subsection{Structural Changes and Characterization}
443 By starting the simulations from the crystalline state, the melting
444 transition and the ice structure can be studied along with the liquid
445 phase behavior beyond the melting point. To locate the melting
446 transition, the constant pressure heat capacity (C$_\text{p}$) was
447 monitored in each of the simulations. In the melting simulations of
448 the 1024 particle ice $I_h$ simulations, a large spike in C$_\text{p}$
449 occurs at 245 K, indicating a first order phase transition for the
450 melting of these ice crystals. When the reaction field is turned off,
451 the melting transition occurs at 235 K. These melting transitions are
452 considerably lower than the experimental value, but this is not
453 surprising when considering the simplicity of the SSD model.
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 \begin{figure}
466 \includegraphics[width=45mm]{corrDiag.eps}
467 \caption{Two dimensional illustration of the angles involved in the
468 correlations observed in figure \ref{contour}.}
469 \label{corrAngle}
470 \end{figure}
471
472 Additional analysis of the melting phase-transition process was
473 performed by using two-dimensional structure and dipole angle
474 correlations. Expressions for these correlations are as follows:
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 figure
485 \ref{corrAngle}. 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}) consists primarily 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, these
498 longer-ranged repeated peak features deteriorate rapidly. The first
499 solvation shell still shows the strong effect of the sticky-potential,
500 although it covers a larger area, extending to include a fraction of
501 aligned dipole peaks within the first solvation shell. The latter
502 peaks lose definition as thermal motion and the competing dipole force
503 overcomes the sticky potential's tight tetrahedral structuring of the
504 fluid.
505
506 This complex interplay between dipole and sticky interactions was
507 remarked upon as a possible reason for the split second peak in the
508 oxygen-oxygen g(\emph{r}).\cite{Ichiye96} At low temperatures, the
509 second solvation shell peak appears to have two distinct parts that
510 blend together to form one observable peak. At higher temperatures,
511 this split character alters to show the leading 4 \AA\ peak dominated
512 by equatorial anti-parallel dipole orientations, and there is tightly
513 bunched group of axially arranged dipoles that most likely consist of
514 the smaller fraction aligned dipole pairs. The trailing part of the
515 split peak at 5 \AA\ is dominated by aligned dipoles that range
516 primarily within the axial to the chief hydrogen bond arrangements
517 similar to those seen in the first solvation shell. This evidence
518 indicates that the dipole pair interaction begins to dominate outside
519 of the range of the dipolar repulsion term, with the primary
520 energetically favorable dipole arrangements populating the region
521 immediately outside this repulsion region (around 4 \AA), and
522 arrangements that seek to ideally satisfy both the sticky and dipole
523 forces locate themselves just beyond this initial buildup (around 5
524 \AA).
525
526 From these findings, the split second peak is primarily the product of
527 the dipolar repulsion term of the sticky potential. In fact, the inner
528 peak can be pushed out and merged with the outer split peak just by
529 extending the switching function cutoff ($s^\prime(r_{ij})$) from its
530 normal 4.0 \AA\ to values of 4.5 or even 5 \AA. This type of
531 correction is not recommended for improving the liquid structure,
532 because the second solvation shell will still be shifted too far
533 out. In addition, this would have an even more detrimental effect on
534 the system densities, leading to a liquid with a more open structure
535 and a density considerably lower than the normal SSD behavior shown
536 previously. A better correction would be to include the
537 quadrupole-quadrupole interactions for the water particles outside of
538 the first solvation shell, but this reduces the simplicity and speed
539 advantage of SSD.
540
541 \subsection{Adjusted Potentials: SSD/RF and SSD/E}
542 The propensity of SSD to adopt lower than expected densities under
543 varying conditions is troubling, especially at higher temperatures. In
544 order to correct this model for use with a reaction field, it is
545 necessary to adjust the force field parameters for the primary
546 intermolecular interactions. In undergoing a reparameterization, it is
547 important not to focus on just one property and neglect the other
548 important properties. In this case, it would be ideal to correct the
549 densities while maintaining the accurate transport properties.
550
551 The possible parameters for tuning include the $\sigma$ and $\epsilon$
552 Lennard-Jones parameters, the dipole strength ($\mu$), and the sticky
553 attractive and dipole repulsive terms with their respective
554 cutoffs. To alter the attractive and repulsive terms of the sticky
555 potential independently, it is necessary to separate the terms as
556 follows:
557 \begin{equation}
558 \begin{split}
559 u_{ij}^{sp}
560 (\mathbf{r}_{ij},\boldsymbol{\Omega}_i,\boldsymbol{\Omega}_j) &=
561 \frac{\nu_0}{2}[s(r_{ij})w(\mathbf{r}_{ij},\boldsymbol{\Omega}_i,\boldsymbol{\Omega}_j)]\\
562 & \quad \ + \frac{\nu_0^\prime}{2}
563 [s^\prime(r_{ij})w^\prime(\mathbf{r}_{ij},\boldsymbol{\Omega}_i,\boldsymbol{\Omega}_j)],
564 \end{split}
565 \end{equation}
566
567 where $\nu_0$ scales the strength of the tetrahedral attraction and
568 $\nu_0^\prime$ acts in an identical fashion on the dipole repulsion
569 term. For purposes of the reparameterization, the separation was
570 performed, but the final parameters were adjusted so that it is
571 unnecessary to separate the terms when implementing the adjusted water
572 potentials. The results of the reparameterizations are shown in table
573 \ref{params}. Note that both the tetrahedral attractive and dipolar
574 repulsive don't share the same lower cutoff ($r_l$) in the newly
575 parameterized potentials - soft sticky dipole reaction field (SSD/RF -
576 for use with a reaction field) and soft sticky dipole enhanced (SSD/E
577 - an attempt to improve the liquid structure in simulations without a
578 long-range correction).
579
580 \begin{table}
581 \caption{Parameters for the original and adjusted models}
582 \begin{tabular}{ l c c c c }
583 \hline \\[-3mm]
584 \ \ \ Parameters\ \ \ & \ \ \ SSD$^\dagger$ \ \ \ & \ SSD1$^\ddagger$\ \ & \ SSD/E\ \ & \ SSD/RF \\
585 \hline \\[-3mm]
586 \ \ \ $\sigma$ (\AA) & 3.051 & 3.016 & 3.035 & 3.019\\
587 \ \ \ $\epsilon$ (kcal/mol) & 0.152 & 0.152 & 0.152 & 0.152\\
588 \ \ \ $\mu$ (D) & 2.35 & 2.35 & 2.42 & 2.48\\
589 \ \ \ $\nu_0$ (kcal/mol) & 3.7284 & 3.6613 & 3.90 & 3.90\\
590 \ \ \ $r_l$ (\AA) & 2.75 & 2.75 & 2.40 & 2.40\\
591 \ \ \ $r_u$ (\AA) & 3.35 & 3.35 & 3.80 & 3.80\\
592 \ \ \ $\nu_0^\prime$ (kcal/mol) & 3.7284 & 3.6613 & 3.90 & 3.90\\
593 \ \ \ $r_l^\prime$ (\AA) & 2.75 & 2.75 & 2.75 & 2.75\\
594 \ \ \ $r_u^\prime$ (\AA) & 4.00 & 4.00 & 3.35 & 3.35\\
595 \\[-2mm]$^\dagger$ ref. \onlinecite{Ichiye96}
596 \\$^\ddagger$ ref. \onlinecite{Ichiye03}
597 \end{tabular}
598 \label{params}
599 \end{table}
600
601 \begin{figure}
602 \includegraphics[width=85mm]{GofRCompare.epsi}
603 \caption{Plots comparing experiment\cite{Head-Gordon00_1} with SSD/E
604 and SSD1 without reaction field (top), as well as SSD/RF and SSD1 with
605 reaction field turned on (bottom). The insets show the respective
606 first peaks in detail. Solid Line - experiment, dashed line - SSD/E
607 and SSD/RF, and dotted line - SSD1 (with and without reaction field).}
608 \label{grcompare}
609 \end{figure}
610
611 \begin{figure}
612 \includegraphics[width=85mm]{dualsticky.ps}
613 \caption{Isosurfaces of the sticky potential for SSD1 (left) and SSD/E \&
614 SSD/RF (right). Light areas correspond to the tetrahedral attractive
615 part, and the darker areas correspond to the dipolar repulsive part.}
616 \label{isosurface}
617 \end{figure}
618
619 In the paper detailing the development of SSD, Liu and Ichiye placed
620 particular emphasis on an accurate description of the first solvation
621 shell. This resulted in a somewhat tall and sharp first peak that
622 integrated to give similar coordination numbers to the experimental
623 data obtained by Soper and Phillips.\cite{Ichiye96,Soper86} New
624 experimental x-ray scattering data from the Head-Gordon lab indicates
625 a slightly lower and shifted first peak in the g$_\mathrm{OO}(r)$, so
626 adjustments to SSD were made while taking into consideration the new
627 experimental findings.\cite{Head-Gordon00_1} Figure \ref{grcompare}
628 shows the relocation of the first peak of the oxygen-oxygen
629 g(\emph{r}) by comparing the revised SSD model (SSD1), SSD-E, and
630 SSD-RF to the new experimental results. Both the modified water models
631 have shorter peaks that are brought in more closely to the
632 experimental peak (as seen in the insets of figure \ref{grcompare}).
633 This structural alteration was accomplished by the combined reduction
634 in the Lennard-Jones $\sigma$ variable and adjustment of the sticky
635 potential strength and cutoffs. As can be seen in table \ref{params},
636 the cutoffs for the tetrahedral attractive and dipolar repulsive terms
637 were nearly swapped with each other. Isosurfaces of the original and
638 modified sticky potentials are shown in figure \cite{isosurface}. In
639 these isosurfaces, it is easy to see how altering the cutoffs changes
640 the repulsive and attractive character of the particles. With a
641 reduced repulsive surface (the darker region), the particles can move
642 closer to one another, increasing the density for the overall
643 system. This change in interaction cutoff also results in a more
644 gradual orientational motion by allowing the particles to maintain
645 preferred dipolar arrangements before they begin to feel the pull of
646 the tetrahedral restructuring. Upon moving closer together, the
647 dipolar repulsion term becomes active and excludes unphysical
648 nearest-neighbor arrangements. This compares with how SSD and SSD1
649 exclude preferred dipole alignments before the particles feel the pull
650 of the ``hydrogen bonds''. Aside from improving the shape of the first
651 peak in the g(\emph{r}), this improves the densities considerably by
652 allowing the persistence of full dipolar character below the previous
653 4.0 \AA\ cutoff.
654
655 While adjusting the location and shape of the first peak of
656 g(\emph{r}) improves the densities, these changes alone are
657 insufficient to bring the system densities up to the values observed
658 experimentally. To finish bringing up the densities, the dipole
659 moments were increased in both the adjusted models. Being a dipole
660 based model, the structure and transport are very sensitive to changes
661 in the dipole moment. The original SSD simply used the dipole moment
662 calculated from the TIP3P water model, which at 2.35 D is
663 significantly greater than the experimental gas phase value of 1.84
664 D. The larger dipole moment is a more realistic value and improves the
665 dielectric properties of the fluid. Both theoretical and experimental
666 measurements indicate a liquid phase dipole moment ranging from 2.4 D
667 to values as high as 3.11 D, so there is quite a range of available
668 values for a reasonable dipole
669 moment.\cite{Sprik91,Kusalik02,Badyal00,Barriol64} The increasing of
670 the dipole moments to 2.42 and 2.48 D for SSD/E and SSD/RF
671 respectively is moderate in this range; however, it leads to
672 significant changes in the density and transport of the water models.
673
674 In order to demonstrate the benefits of these reparameterizations, a
675 series of NPT and NVE simulations were performed to probe the density
676 and transport properties of the adapted models and compare the results
677 to the original SSD model. This comparison involved full NPT melting
678 sequences for both SSD/E and SSD/RF, as well as NVE transport
679 calculations at the calculated self-consistent densities. Again, the
680 results come from five separate simulations of 1024 particle systems,
681 and the melting sequences were started from different ice $I_h$
682 crystals constructed as stated earlier. Like before, each NPT
683 simulation was equilibrated for 100 ps before a 200 ps data collection
684 run at each temperature step, and they used the final configuration
685 from the previous temperature simulation as a starting point. All of
686 the NVE simulations had the same thermalization, equilibration, and
687 data collection times stated earlier in this paper.
688
689 \begin{figure}
690 \includegraphics[width=62mm, angle=-90]{ssdeDense.epsi}
691 \caption{Comparison of densities calculated with SSD/E to SSD1 without a
692 reaction field, TIP3P\cite{Jorgensen98b}, TIP5P\cite{Jorgensen00},
693 SPC/E\cite{Clancy94}, and Experiment\cite{CRC80}. The window shows a
694 expansion around 300 K with error bars included to clarify this region
695 of interest. Note that both SSD1 and SSD/E show good agreement with
696 experiment when the long-range correction is neglected.}
697 \label{ssdedense}
698 \end{figure}
699
700 Figure \ref{ssdedense} shows the density profile for the SSD/E model
701 in comparison to SSD1 without a reaction field, experiment, and other
702 common water models. The calculated densities for both SSD/E and SSD1
703 have increased significantly over the original SSD model (see figure
704 \ref{dense1} and are in significantly better agreement with the
705 experimental values. At 298 K, the density of SSD/E and SSD1 without a
706 long-range correction are 0.996$\pm$0.001 g/cm$^3$ and 0.999$\pm$0.001
707 g/cm$^3$ respectively. These both compare well with the experimental
708 value of 0.997 g/cm$^3$, and they are considerably better than the SSD
709 value of 0.967$\pm$0.003 g/cm$^3$. The changes to the dipole moment
710 and sticky switching functions have improved the structuring of the
711 liquid (as seen in figure \ref{grcompare}, but they have shifted the
712 density maximum to much lower temperatures. This comes about via an
713 increase of the liquid disorder through the weakening of the sticky
714 potential and strengthening of the dipolar character. However, this
715 increasing disorder in the SSD/E model has little affect on the
716 melting transition. By monitoring C$\text{p}$ throughout these
717 simulations, the melting transition for SSD/E occurred at 235 K, the
718 same transition temperature observed with SSD and SSD1.
719
720 \begin{figure}
721 \includegraphics[width=62mm, angle=-90]{ssdrfDense.epsi}
722 \caption{Comparison of densities calculated with SSD/RF to SSD1 with a
723 reaction field, TIP3P\cite{Jorgensen98b}, TIP5P\cite{Jorgensen00},
724 SPC/E\cite{Clancy94}, and Experiment\cite{CRC80}. The inset shows the
725 necessity of reparameterization when utilizing a reaction field
726 long-ranged correction - SSD/RF provides significantly more accurate
727 densities than SSD1 when performing room temperature simulations.}
728 \label{ssdrfdense}
729 \end{figure}
730
731 Including the reaction field long-range correction results in a more
732 interesting comparison. A density profile including SSD/RF and SSD1
733 with an active reaction field is shown in figure \ref{ssdrfdense}. As
734 observed in the simulations without a reaction field, the densities of
735 SSD/RF and SSD1 show a dramatic increase over normal SSD (see figure
736 \ref{dense1}). At 298 K, SSD/RF has a density of 0.997$\pm$0.001
737 g/cm$^3$, right in line with experiment and considerably better than
738 the SSD value of 0.941$\pm$0.001 g/cm$^3$ and the SSD1 value of
739 0.972$\pm$0.002 g/cm$^3$. These results further emphasize the
740 importance of reparameterization in order to model the density
741 properly under different simulation conditions. Again, these changes
742 don't have that profound an effect on the melting point which is
743 observed at 245 K for SSD/RF, identical to SSD and only 5 K lower than
744 SSD1 with a reaction field. However, the difference in density maxima
745 is not quite as extreme with SSD/RF showing a density maximum at 255
746 K, fairly close to 260 and 265 K, the density maxima for SSD and SSD1
747 respectively.
748
749 \begin{figure}
750 \includegraphics[width=65mm, angle=-90]{ssdeDiffuse.epsi}
751 \caption{Plots of the diffusion constants calculated from SSD/E and SSD1,
752 both without a reaction field, along with experimental results are
753 from Gillen \emph{et al.}\cite{Gillen72} and Mills\cite{Mills73}. The
754 NVE calculations were performed at the average densities observed in
755 the 1 atm NPT simulations for the respective models. SSD/E is
756 slightly more fluid than experiment at all of the temperatures, but
757 it is closer than SSD1 without a long-range correction.}
758 \label{ssdediffuse}
759 \end{figure}
760
761 The reparameterization of the SSD water model, both for use with and
762 without an applied long-range correction, brought the densities up to
763 what is expected for simulating liquid water. In addition to improving
764 the densities, it is important that particle transport be maintained
765 or improved. Figure \ref{ssdediffuse} compares the temperature
766 dependence of the diffusion constant of SSD/E to SSD1 without an
767 active reaction field, both at the densities calculated at 1 atm and
768 at the experimentally calculated densities for super-cooled and liquid
769 water. In the upper plot, the diffusion constant for SSD/E is
770 consistently a little faster than experiment, while SSD1 remains
771 slower than experiment until relatively high temperatures (greater
772 than 330 K). Both models follow the shape of the experimental trend
773 well below 300 K, but the trend leans toward diffusing too rapidly at
774 higher temperatures, something that is especially apparent with
775 SSD1. This accelerated increasing of diffusion is caused by the
776 rapidly decreasing system density with increasing temperature. Though
777 it is difficult to see in figure \ref{ssdedense}, the densities of SSD1
778 decay more rapidly with temperature than do those of SSD/E, leading to
779 more visible deviation from the experimental diffusion trend. Thus,
780 the changes made to improve the liquid structure may have had an
781 adverse affect on the density maximum, but they improve the transport
782 behavior of the water model.
783
784 \begin{figure}
785 \includegraphics[width=65mm, angle=-90]{ssdrfDiffuse.epsi}
786 \caption{Plots of the diffusion constants calculated from SSD/RF and SSD1,
787 both with an active reaction field, along with experimental results
788 from Gillen \emph{et al.}\cite{Gillen72} and Mills\cite{Mills73}. The
789 NVE calculations were performed at the average densities observed in
790 the 1 atm NPT simulations for both of the models. Note how accurately
791 SSD/RF simulates the diffusion of water throughout this temperature
792 range. The more rapidly increasing diffusion constants at high
793 temperatures for both models is attributed to the significantly lower
794 densities than observed in experiment.}
795 \label{ssdrfdiffuse}
796 \end{figure}
797
798 In figure \ref{ssdrfdiffuse}, the diffusion constants for SSD/RF are
799 compared with SSD1 with an active reaction field. Note that SSD/RF
800 tracks the experimental results incredibly well, identical within
801 error throughout the temperature range shown and only showing a slight
802 increasing trend at higher temperatures. SSD1 tends to diffuse more
803 slowly at low temperatures and deviates to diffuse too rapidly at
804 temperatures greater than 330 K. As was stated in the SSD/E
805 comparisons, this deviation away from the ideal trend is due to a
806 rapid decrease in density at higher temperatures. SSD/RF doesn't
807 suffer from this problem as much as SSD1, because the calculated
808 densities are more true to experiment. These results again emphasize
809 the importance of careful reparameterization when using an altered
810 long-range correction.
811
812 \subsection{Additional Observations}
813
814 \begin{figure}
815 \includegraphics[width=85mm]{povIce.ps}
816 \caption{A water lattice built from the crystal structure that SSD/E
817 assumed when undergoing an extremely restricted temperature NPT
818 simulation. This form of ice is referred to as ice 0 to emphasize its
819 simulation origins. This image was taken of the (001) face of the
820 crystal.}
821 \label{weirdice}
822 \end{figure}
823
824 While performing restricted temperature melting sequences of SSD/E not
825 discussed earlier in this paper, some interesting observations were
826 made. After melting at 235 K, two of five systems underwent
827 crystallization events near 245 K. As the heating process continued,
828 the two systems remained crystalline until finally melting between 320
829 and 330 K. The final configurations of these two melting sequences
830 show an expanded zeolite-like crystal structure that does not
831 correspond to any known form of ice. For convenience and to help
832 distinguish it from the experimentally observed forms of ice, this
833 crystal structure will henceforth be referred to as ice-zero (ice
834 0). The crystallinity was extensive enough that a near ideal crystal
835 structure could be obtained. Figure \ref{weirdice} shows the repeating
836 crystal structure of a typical crystal at 5 K. Each water molecule is
837 hydrogen bonded to four others; however, the hydrogen bonds are flexed
838 rather than perfectly straight. This results in a skewed tetrahedral
839 geometry about the central molecule. Looking back at figure
840 \ref{isosurface}, it is easy to see how these flexed hydrogen bonds
841 are allowed in that the attractive regions are conical in shape, with
842 the greatest attraction in the central region. Though not ideal, these
843 flexed hydrogen bonds are favorable enough to stabilize an entire
844 crystal generated around them. In fact, the imperfect ice 0 crystals
845 were so stable that they melted at temperatures nearly 100 K greater
846 than both ice I$_c$ and I$_h$.
847
848 These initial simulations indicated that ice 0 is the preferred ice
849 structure for at least SSD/E. To verify this, a comparison was made
850 between near ideal crystals of ice $I_h$, ice $I_c$, and ice 0 at
851 constant pressure with SSD/E, SSD/RF, and SSD1. Near ideal versions of
852 the three types of crystals were cooled to 1 K, and the potential
853 energies of each were compared using all three water models. With
854 every water model, ice 0 turned out to have the lowest potential
855 energy: 5\% lower than $I_h$ with SSD1, 6.5\% lower with SSD/E, and
856 7.5\% lower with SSD/RF.
857
858 In addition to these low temperature comparisons, melting sequences
859 were performed with ice 0 as the initial configuration using SSD/E,
860 SSD/RF, and SSD1 both with and without a reaction field. The melting
861 transitions for both SSD/E and SSD1 without a reaction field occurred
862 at temperature in excess of 375 K. SSD/RF and SSD1 with a reaction
863 field had more reasonable melting transitions, down near 325 K. These
864 melting point observations emphasize how preferred this crystal
865 structure is over the most common types of ice when using these single
866 point water models.
867
868 Recognizing that the above tests show ice 0 to be both the most stable
869 and lowest density crystal structure for these single point water
870 models, it is interesting to speculate on the relative stability of
871 this crystal structure with charge based water models. As a quick
872 test, these 3 crystal types were converted from SSD type particles to
873 TIP3P waters and read into CHARMM.\cite{Karplus83} Identical energy
874 minimizations were performed on all of these crystals to compare the
875 system energies. Again, ice 0 was observed to have the lowest total
876 system energy. The total energy of ice 0 was ~2\% lower than ice
877 $I_h$, which was in turn ~3\% lower than ice $I_c$. From these initial
878 results, we would not be surprised if results from the other common
879 water models show ice 0 to be the lowest energy crystal structure. A
880 continuation on work studying ice 0 with multi-point water models will
881 be published in a coming article.
882
883 \section{Conclusions}
884 The density maximum and temperature dependent transport for the SSD
885 water model, both with and without the use of reaction field, were
886 studied via a series of NPT and NVE simulations. The constant pressure
887 simulations of the melting of both $I_h$ and $I_c$ ice showed a
888 density maximum near 260 K. In most cases, the calculated densities
889 were significantly lower than the densities calculated in simulations
890 of other water models. Analysis of particle diffusion showed SSD to
891 capture the transport properties of experimental very well in both the
892 normal and super-cooled liquid regimes. In order to correct the
893 density behavior, the original SSD model was reparameterized for use
894 both with and without a reaction field (SSD/RF and SSD/E), and
895 comparison simulations were performed with SSD1, the density corrected
896 version of SSD. Both models improve the liquid structure, density
897 values, and diffusive properties under their respective conditions,
898 indicating the necessity of reparameterization when altering the
899 long-range correction specifics. When taking the appropriate
900 considerations, these simple water models are excellent choices for
901 representing explicit water in large scale simulations of biochemical
902 systems.
903
904 \section{Acknowledgments}
905 Support for this project was provided by the National Science
906 Foundation under grant CHE-0134881. Computation time was provided by
907 the Notre Dame Bunch-of-Boxes (B.o.B) computer cluster under NSF grant
908 DMR 00 79647.
909
910 \bibliographystyle{jcp}
911
912 \bibliography{nptSSD}
913
914 %\pagebreak
915
916 \end{document}