ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/ssdePaper/nptSSD.tex
Revision: 759
Committed: Wed Sep 10 22:42:57 2003 UTC (20 years, 9 months ago) by chrisfen
Content type: application/x-tex
File size: 48932 byte(s)
Log Message:
Adjustments to the introduction.  Minor additions to the methods section.

Chris

File Contents

# User Rev Content
1 chrisfen 743 \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 chrisfen 759 \title{On the temperature dependent properties of the soft sticky dipole (SSD) and related single point water models}
27 chrisfen 743
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 the SSD water model, both with and without the use of reaction
40     field. The constant pressure simulations of the melting of both $I_h$
41     and $I_c$ ice showed a density maximum near 260 K. In most cases, the
42     calculated densities were significantly lower than the densities
43     calculated in simulations of other water models. Analysis of particle
44     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 of
150     SSD, it was shown that it actually improves upon the prediction of
151     water's dynamical properties 3 and 4-point models.\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 chrisfen 757 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 chrisfen 759 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 chrisfen 757
179 chrisfen 743 \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} For the constant pressure
213     simulations, the \emph{Q} parameter for the was set to 5.0 amu
214     \(\cdot\)\AA\(^{2}\), and the relaxation time (\(\tau\))\ was set at
215     100 ps.
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 actually quite sensitive to errors in the
224     equations of motion. The original implementation of this code utilized
225     quaternions for rotational motion propagation; however, a detailed
226     investigation showed that they resulted in a steady drift in the total
227     energy, 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 chrisfen 759 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 chrisfen 743
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. On average, a 1000 SSD
248 chrisfen 759 particle simulation shows a 7\% increase in computation time using the
249 chrisfen 743 symplectic step method in place of quaternions. This cost is more than
250     justified when comparing the energy conservation of the two methods as
251 chrisfen 759 illustrated in figure \ref{timestep}.
252 chrisfen 743
253     \begin{figure}
254     \includegraphics[width=61mm, angle=-90]{timeStep.epsi}
255     \caption{Energy conservation using quaternion based integration versus
256     the symplectic step method proposed by Dullweber \emph{et al.} with
257     increasing time step. For each time step, the dotted line is total
258     energy using the symplectic step integrator, and the solid line comes
259     from the quaternion integrator. The larger time step plots are shifted
260     up from the true energy baseline for clarity.}
261     \label{timestep}
262     \end{figure}
263    
264     In figure \ref{timestep}, the resulting energy drift at various time
265     steps for both the symplectic step and quaternion integration schemes
266     is compared. All of the 1000 SSD particle simulations started with the
267     same configuration, and the only difference was the method for
268     handling rotational motion. At time steps of 0.1 and 0.5 fs, both
269     methods for propagating particle rotation conserve energy fairly well,
270     with the quaternion method showing a slight energy drift over time in
271     the 0.5 fs time step simulation. At time steps of 1 and 2 fs, the
272     energy conservation benefits of the symplectic step method are clearly
273 chrisfen 759 demonstrated. Thus, while maintaining the same degree of energy
274     conservation, one can take considerably longer time steps, leading to
275     an overall reduction in computation time.
276 chrisfen 743
277     Energy drift in these SSD particle simulations was unnoticeable for
278     time steps up to three femtoseconds. A slight energy drift on the
279     order of 0.012 kcal/mol per nanosecond was observed at a time step of
280     four femtoseconds, and as expected, this drift increases dramatically
281     with increasing time step. To insure accuracy in the constant energy
282     simulations, time steps were set at 2 fs and kept at this value for
283     constant pressure simulations as well.
284    
285     Ice crystals in both the $I_h$ and $I_c$ lattices were generated as
286     starting points for all the simulations. The $I_h$ crystals were
287     formed by first arranging the center of masses of the SSD particles
288     into a ``hexagonal'' ice lattice of 1024 particles. Because of the
289     crystal structure of $I_h$ ice, the simulation box assumed a
290     rectangular shape with a edge length ratio of approximately
291     1.00$\times$1.06$\times$1.23. The particles were then allowed to
292     orient freely about fixed positions with angular momenta randomized at
293     400 K for varying times. The rotational temperature was then scaled
294     down in stages to slowly cool the crystals down to 25 K. The particles
295     were then allowed translate with fixed orientations at a constant
296     pressure of 1 atm for 50 ps at 25 K. Finally, all constraints were
297     removed and the ice crystals were allowed to equilibrate for 50 ps at
298     25 K and a constant pressure of 1 atm. This procedure resulted in
299     structurally stable $I_h$ ice crystals that obey the Bernal-Fowler
300     rules\cite{Bernal33,Rahman72}. This method was also utilized in the
301     making of diamond lattice $I_c$ ice crystals, with each cubic
302     simulation box consisting of either 512 or 1000 particles. Only
303     isotropic volume fluctuations were performed under constant pressure,
304     so the ratio of edge lengths remained constant throughout the
305     simulations.
306    
307     \section{Results and discussion}
308    
309     Melting studies were performed on the randomized ice crystals using
310     constant pressure and temperature dynamics. This involved an initial
311     randomization of velocities about the starting temperature of 25 K for
312     varying amounts of time. The systems were all equilibrated for 100 ps
313     prior to a 200 ps data collection run at each temperature setting,
314     ranging from 25 to 400 K, with a maximum degree increment of 25 K. For
315     regions of interest along this stepwise progression, the temperature
316     increment was decreased from 25 K to 10 and then 5 K. The above
317     equilibration and production times were sufficient in that the system
318     volume fluctuations dampened out in all but the very cold simulations
319     (below 225 K). In order to further improve statistics, five separate
320     simulation progressions were performed, and the averaged results from
321     the $I_h$ melting simulations are shown in figure \ref{dense1}.
322    
323     \begin{figure}
324     \includegraphics[width=65mm, angle=-90]{1hdense.epsi}
325     \caption{Average density of SSD water at increasing temperatures
326     starting from ice $I_h$ lattice.}
327     \label{dense1}
328     \end{figure}
329    
330     \subsection{Density Behavior}
331     In the initial average density versus temperature plot, the density
332     maximum clearly appears between 255 and 265 K. The calculated
333     densities within this range were nearly indistinguishable, as can be
334     seen in the zoom of this region of interest, shown in figure
335     \ref{dense1}. The greater certainty of the average value at 260 K makes
336     a good argument for the actual density maximum residing at this
337     midpoint value. Figure \ref{dense1} was constructed using ice $I_h$
338     crystals for the initial configuration; and though not pictured, the
339     simulations starting from ice $I_c$ crystal configurations showed
340     similar results, with a liquid-phase density maximum in this same
341     region (between 255 and 260 K). In addition, the $I_c$ crystals are
342     more fragile than the $I_h$ crystals, leading them to deform into a
343     dense glassy state at lower temperatures. This resulted in an overall
344     low temperature density maximum at 200 K, but they still retained a
345     common liquid state density maximum with the $I_h$ simulations.
346    
347     \begin{figure}
348     \includegraphics[width=65mm,angle=-90]{dense2.eps}
349     \caption{Density versus temperature for TIP4P\cite{Jorgensen98b},
350     TIP3P\cite{Jorgensen98b}, SPC/E\cite{Clancy94}, SSD without Reaction
351     Field, SSD, and Experiment\cite{CRC80}. }
352     \label{dense2}
353     \end{figure}
354    
355     The density maximum for SSD actually compares quite favorably to other
356     simple water models. Figure \ref{dense2} shows a plot of these
357     findings with the density progression of several other models and
358     experiment obtained from other
359     sources.\cite{Jorgensen98b,Clancy94,CRC80} Of the listed simple water
360     models, SSD has results closest to the experimentally observed water
361     density maximum. Of the listed water models, TIP4P has a density
362     maximum behavior most like that seen in SSD. Though not shown, it is
363     useful to note that TIP5P has a water density maximum nearly identical
364     to experiment.
365    
366     Possibly of more importance is the density scaling of SSD relative to
367     other common models at any given temperature (Fig. \ref{dense2}). Note
368     that the SSD model assumes a lower density than any of the other
369     listed models at the same pressure, behavior which is especially
370     apparent at temperatures greater than 300 K. Lower than expected
371     densities have been observed for other systems with the use of a
372     reaction field for long-range electrostatic interactions, so the most
373     likely reason for these significantly lower densities in these
374     simulations is the presence of the reaction field.\cite{Berendsen98}
375     In order to test the effect of the reaction field on the density of
376     the systems, the simulations were repeated for the temperature region
377     of interest without a reaction field present. The results of these
378     simulations are also displayed in figure \ref{dense2}. Without
379     reaction field, these densities increase considerably to more
380     experimentally reasonable values, especially around the freezing point
381     of liquid water. The shape of the curve is similar to the curve
382     produced from SSD simulations using reaction field, specifically the
383     rapidly decreasing densities at higher temperatures; however, a slight
384     shift in the density maximum location, down to 245 K, is
385     observed. This is probably a more accurate comparison to the other
386     listed water models in that no long range corrections were applied in
387     those simulations.\cite{Clancy94,Jorgensen98b}
388    
389     It has been observed that densities are dependent on the cutoff radius
390     used for a variety of water models in simulations both with and
391     without the use of reaction field.\cite{Berendsen98} In order to
392     address the possible affect of cutoff radius, simulations were
393     performed with a dipolar cutoff radius of 12.0 \AA\ to compliment the
394     previous SSD simulations, all performed with a cutoff of 9.0 \AA. All
395     the resulting densities overlapped within error and showed no
396     significant trend in lower or higher densities as a function of cutoff
397     radius, both for simulations with and without reaction field. These
398     results indicate that there is no major benefit in choosing a longer
399     cutoff radius in simulations using SSD. This is comforting in that the
400     use of a longer cutoff radius results in a near doubling of the time
401     required to compute a single trajectory.
402    
403     \subsection{Transport Behavior}
404     Of importance in these types of studies are the transport properties
405     of the particles and how they change when altering the environmental
406     conditions. In order to probe transport, constant energy simulations
407     were performed about the average density uncovered by the constant
408     pressure simulations. Simulations started with randomized velocities
409     and underwent 50 ps of temperature scaling and 50 ps of constant
410     energy equilibration before obtaining a 200 ps trajectory. Diffusion
411     constants were calculated via root-mean square deviation analysis. The
412     averaged results from 5 sets of these NVE simulations is displayed in
413     figure \ref{diffuse}, alongside experimental, SPC/E, and TIP5P
414     results.\cite{Gillen72,Mills73,Clancy94,Jorgensen01}
415    
416     \begin{figure}
417     \includegraphics[width=65mm, angle=-90]{betterDiffuse.epsi}
418     \caption{Average diffusion coefficient over increasing temperature for
419     SSD, SPC/E\cite{Clancy94}, TIP5P\cite{Jorgensen01}, and Experimental
420     data from Gillen \emph{et al.}\cite{Gillen72}, and from
421     Mills\cite{Mills73}.}
422     \label{diffuse}
423     \end{figure}
424    
425     The observed values for the diffusion constant point out one of the
426     strengths of the SSD model. Of the three experimental models shown,
427     the SSD model has the most accurate depiction of the diffusion trend
428     seen in experiment in both the supercooled and normal regimes. SPC/E
429     does a respectable job by getting similar values as SSD and experiment
430     around 290 K; however, it deviates at both higher and lower
431     temperatures, failing to predict the experimental trend. TIP5P and SSD
432     both start off low at the colder temperatures and tend to diffuse too
433     rapidly at the higher temperatures. This type of trend at the higher
434     temperatures is not surprising in that the densities of both TIP5P and
435     SSD are lower than experimental water at temperatures higher than room
436     temperature. When calculating the diffusion coefficients for SSD at
437     experimental densities, the resulting values fall more in line with
438     experiment at these temperatures, albeit not at standard
439     pressure. Results under these conditions can be found later in this
440     paper.
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 in that SSD is a simple rigid body model with a fixed
454     dipole.
455    
456     \begin{figure}
457     \includegraphics[width=85mm]{fullContours.eps}
458     \caption{Contour plots of 2D angular g($r$)'s for 512 SSD systems at
459     100 K (A \& B) and 300 K (C \& D). Contour colors are inverted for
460     clarity: dark areas signify peaks while light areas signify
461     depressions. White areas have g(\emph{r}) values below 0.5 and black
462     areas have values above 1.5.}
463     \label{contour}
464     \end{figure}
465    
466     Additional analyses for understanding the melting phase-transition
467     process were performed via two-dimensional structure and dipole angle
468     correlations. Expressions for the correlations are as follows:
469    
470     \begin{figure}
471     \includegraphics[width=45mm]{corrDiag.eps}
472     \caption{Two dimensional illustration of the angles involved in the
473     correlations observed in figure \ref{contour}.}
474     \label{corrAngle}
475     \end{figure}
476    
477     \begin{multline}
478     g_{\text{AB}}(r,\cos\theta) = \\
479     \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\ ,
480     \end{multline}
481     \begin{multline}
482     g_{\text{AB}}(r,\cos\omega) = \\
483     \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\ ,
484     \end{multline}
485     where $\theta$ and $\omega$ refer to the angles shown in the above
486     illustration. By binning over both distance and the cosine of the
487     desired angle between the two dipoles, the g(\emph{r}) can be
488     dissected to determine the common dipole arrangements that constitute
489     the peaks and troughs. Frames A and B of figure \ref{contour} show a
490     relatively crystalline state of an ice $I_c$ simulation. The first
491     peak of the g(\emph{r}) primarily consists of the preferred hydrogen
492     bonding arrangements as dictated by the tetrahedral sticky potential,
493     one peak for the donating and the other for the accepting hydrogen
494     bonds. Due to the high degree of crystallinity of the sample, the
495     second and third solvation shells show a repeated peak arrangement
496     which decays at distances around the fourth solvation shell, near the
497     imposed cutoff for the Lennard-Jones and dipole-dipole interactions.
498     In the higher temperature simulation shown in frames C and D, the
499     repeated peak features are significantly blurred. The first solvation
500     shell still shows the strong effect of the sticky-potential, although
501     it covers a larger area, extending to include a fraction of aligned
502     dipole peaks within the first solvation shell. The latter peaks lose
503     definition as thermal motion and the competing dipole force overcomes
504     the sticky potential's tight tetrahedral structuring of the 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 of it's range (around 4 \AA), and arrangements
522     that seek to ideally satisfy both the sticky and dipole forces locate
523     themselves just beyond this region (around 5 \AA).
524    
525     From these findings, the split second peak is primarily the product of
526     the dipolar repulsion term of the sticky potential. In fact, the
527     leading of the two peaks can be pushed out and merged with the outer
528     split peak just by extending the switching function cutoff
529     ($s^\prime(r_{ij})$) from its normal 4.0 \AA\ to values of 4.5 or even
530     5 \AA. This type of correction is not recommended for improving the
531     liquid structure, because the second solvation shell will still be
532     shifted too far out. In addition, this would have an even more
533     detrimental effect on the system densities, leading to a liquid with a
534     more open structure and a density considerably lower than the normal
535     SSD behavior shown previously. A better correction would be to include
536     the quadrupole-quadrupole interactions for the water particles outside
537     of the first solvation shell, but this reduces the simplicity and
538     speed advantage of SSD, so it is not the most desirable path to take.
539    
540     \subsection{Adjusted Potentials: SSD/E and SSD/RF}
541     The propensity of SSD to adopt lower than expected densities under
542     varying conditions is troubling, especially at higher temperatures. In
543     order to correct this behavior, it's necessary to adjust the force
544     field parameters for the primary intermolecular interactions. In
545     undergoing a reparameterization, it is important not to focus on just
546     one property and neglect the other important properties. In this case,
547     it would be ideal to correct the densities while maintaining the
548     accurate transport properties.
549    
550     The possible parameters for tuning include the $\sigma$ and $\epsilon$
551     Lennard-Jones parameters, the dipole strength ($\mu$), and the sticky
552     attractive and dipole repulsive terms with their respective
553     cutoffs. To alter the attractive and repulsive terms of the sticky
554     potential independently, it is necessary to separate the terms as
555     follows:
556     \begin{equation}
557     \begin{split}
558     u_{ij}^{sp}
559     (\mathbf{r}_{ij},\boldsymbol{\Omega}_i,\boldsymbol{\Omega}_j) &=
560     \frac{\nu_0}{2}[s(r_{ij})w(\mathbf{r}_{ij},\boldsymbol{\Omega}_i,\boldsymbol{\Omega}_j)]\\
561     & \quad \ + \frac{\nu_0^\prime}{2}
562     [s^\prime(r_{ij})w^\prime(\mathbf{r}_{ij},\boldsymbol{\Omega}_i,\boldsymbol{\Omega}_j)],
563     \end{split}
564     \end{equation}
565    
566     where $\nu_0$ scales the strength of the tetrahedral attraction and
567     $\nu_0^\prime$ acts in an identical fashion on the dipole repulsion
568     term. For purposes of the reparameterization, the separation was
569     performed, but the final parameters were adjusted so that it is
570     unnecessary to separate the terms when implementing the adjusted water
571     potentials. The results of the reparameterizations are shown in table
572     \ref{params}. Note that both the tetrahedral attractive and dipolar
573     repulsive don't share the same lower cutoff ($r_l$) in the newly
574     parameterized potentials - soft sticky dipole enhanced (SSD/E) and
575     soft sticky dipole reaction field (SSD/RF).
576    
577     \begin{table}
578     \caption{Parameters for the original and adjusted models}
579     \begin{tabular}{ l c c c }
580     \hline \\[-3mm]
581     \ Parameters & \ \ \ SSD$^\dagger$\ \ \ \ & \ SSD/E\ \ & \ SSD/RF\ \ \\
582     \hline \\[-3mm]
583     \ \ \ $\sigma$ (\AA) & 3.051 & 3.035 & 3.019\\
584     \ \ \ $\epsilon$ (kcal/mol)\ \ & 0.152 & 0.152 & 0.152\\
585     \ \ \ $\mu$ (D) & 2.35 & 2.418 & 2.480\\
586     \ \ \ $\nu_0$ (kcal/mol)\ \ & 3.7284 & 3.90 & 3.90\\
587     \ \ \ $r_l$ (\AA) & 2.75 & 2.40 & 2.40\\
588     \ \ \ $r_u$ (\AA) & 3.35 & 3.80 & 3.80\\
589     \ \ \ $\nu_0^\prime$ (kcal/mol)\ \ & 3.7284 & 3.90 & 3.90\\
590     \ \ \ $r_l^\prime$ (\AA) & 2.75 & 2.75 & 2.75\\
591     \ \ \ $r_u^\prime$ (\AA) & 4.00 & 3.35 & 3.35\\
592     \\[-2mm]$^\dagger$ ref. \onlinecite{Ichiye96}
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 SSD without reaction field (top), as well as SSD/RF and SSD 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 - SSD (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 SSD (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=85mm]{ssdecompare.epsi}
687     \caption{Comparison of densities calculated with SSD/E to SSD without a
688     reaction field, TIP4P\cite{Jorgensen98b}, TIP3P\cite{Jorgensen98b},
689     SPC/E\cite{Clancy94}, and Experiment\cite{CRC80}. The upper plot
690     includes error bars, and the calculated results from the other
691     references were removed for clarity.}
692     \label{ssdedense}
693     \end{figure}
694    
695     Figure \ref{ssdedense} shows the density profile for the SSD/E water
696     model in comparison to the original SSD without a reaction field,
697     experiment, and the other common water models considered
698     previously. The calculated densities have increased significantly over
699     the original SSD model and match the experimental value just below 298
700     K. At 298 K, the density of SSD/E is 0.995$\pm$0.001 g/cm$^3$, which
701     compares well with the experimental value of 0.997 g/cm$^3$ and is
702     considerably better than the SSD value of 0.967$\pm$0.003
703     g/cm$^3$. The increased dipole moment in SSD/E has helped to flatten
704     out the curve at higher temperatures, only the improvement is marginal
705     at best. This steep drop in densities is due to the dipolar rather
706     than charge based interactions which decay more rapidly at longer
707     distances.
708    
709     By monitoring C$\text{p}$ throughout these simulations, the melting
710     transition for SSD/E was observed at 230 K, about 5 degrees lower than
711     SSD. The resulting density maximum is located at 240 K, again about 5
712     degrees lower than the SSD value of 245 K. Though there is a decrease
713     in both of these values, the corrected densities near room temperature
714     justify the modifications taken.
715    
716     \begin{figure}
717     \includegraphics[width=85mm]{ssdrfcompare.epsi}
718     \caption{Comparison of densities calculated with SSD/RF to SSD with a
719     reaction field, TIP4P\cite{Jorgensen98b}, TIP3P\cite{Jorgensen98b},
720     SPC/E\cite{Clancy94}, and Experiment\cite{CRC80}. The upper plot
721     includes error bars, and the calculated results from the other
722     references were removed for clarity.}
723     \label{ssdrfdense}
724     \end{figure}
725    
726     Figure \ref{ssdrfdense} shows a density comparison between SSD/RF and
727     SSD with an active reaction field. Like in the simulations of SSD/E,
728     the densities show a dramatic increase over normal SSD. At 298 K,
729     SSD/RF has a density of 0.997$\pm$0.001 g/cm$^3$, right in line with
730     experiment and considerably better than the SSD value of
731     0.941$\pm$0.001 g/cm$^3$. The melting point is observed at 240 K,
732     which is 5 degrees lower than SSD with a reaction field, and the
733     density maximum at 255 K, again 5 degrees lower than SSD. The density
734     at higher temperature still drops off more rapidly than the charge
735     based models but is in better agreement than SSD/E.
736    
737     The reparameterization of the SSD water model, both for use with and
738     without an applied long-range correction, brought the densities up to
739     what is expected for simulating liquid water. In addition to improving
740     the densities, it is important that particle transport be maintained
741     or improved. Figure \ref{ssdediffuse} compares the temperature
742     dependence of the diffusion constant of SSD/E to SSD without an active
743     reaction field, both at the densities calculated at 1 atm and at the
744     experimentally calculated densities for super-cooled and liquid
745     water. In the upper plot, the diffusion constant for SSD/E is
746     consistently a little faster than experiment, while SSD starts off
747     slower than experiment and crosses to merge with SSD/E at high
748     temperatures. Both models follow the experimental trend well, but
749     diffuse too rapidly at higher temperatures. This abnormally fast
750     diffusion is caused by the decreased system density. Since the
751     densities of SSD/E don't deviate as much from experiment as those of
752     SSD, it follows the experimental trend more closely. This observation
753     is backed up by looking at the lower plot. The diffusion constants for
754     SSD/E track with the experimental values while SSD deviates on the low
755     side of the trend with increasing temperature. This is again a product
756     of SSD/E having densities closer to experiment, and not deviating to
757     lower densities with increasing temperature as rapidly.
758    
759     \begin{figure}
760     \includegraphics[width=85mm]{ssdediffuse.epsi}
761     \caption{Plots of the diffusion constants calculated from SSD/E and SSD,
762     both without a reaction field along with experimental results from
763     Gillen \emph{et al.}\cite{Gillen72} and Mills\cite{Mills73}. The
764     upper plot is at densities calculated from the NPT simulations at a
765     pressure of 1 atm, while the lower plot is at the experimentally
766     calculated densities.}
767     \label{ssdediffuse}
768     \end{figure}
769    
770     \begin{figure}
771     \includegraphics[width=85mm]{ssdrfdiffuse.epsi}
772     \caption{Plots of the diffusion constants calculated from SSD/RF and SSD,
773     both with an active reaction field along with experimental results
774     from Gillen \emph{et al.}\cite{Gillen72} and Mills\cite{Mills73}. The
775     upper plot is at densities calculated from the NPT simulations at a
776     pressure of 1 atm, while the lower plot is at the experimentally
777     calculated densities.}
778     \label{ssdrfdiffuse}
779     \end{figure}
780    
781     In figure \ref{ssdrfdiffuse}, the diffusion constants for SSD/RF are
782     compared with SSD with an active reaction field. In the upper plot,
783     SSD/RF tracks with the experimental results incredibly well, identical
784     within error throughout the temperature range and only showing a
785     slight increasing trend at higher temperatures. SSD also tracks
786     experiment well, only it tends to diffuse a little more slowly at low
787     temperatures and deviates to diffuse too rapidly at high
788     temperatures. As was stated in the SSD/E comparisons, this deviation
789     away from the ideal trend is due to a rapid decrease in density at
790     higher temperatures. SSD/RF doesn't suffer from this problem as much
791     as SSD, because the calculated densities are more true to
792     experiment. This is again emphasized in the lower plot, where SSD/RF
793     tracks the experimental diffusion exactly while SSD's diffusion
794     constants are slightly too low due to its need for a lower density at
795     the specified temperature.
796    
797     \subsection{Additional Observations}
798    
799     While performing the melting sequences of SSD/E, some interesting
800     observations were made. After melting at 230 K, two of the systems
801     underwent crystallization events near 245 K. As the heating process
802     continued, the two systems remained crystalline until finally melting
803     between 320 and 330 K. These simulations were excluded from the data
804     set shown in figure \ref{ssdedense} and replaced with two additional
805     melting sequences that did not undergo this anomalous phase
806     transition, while this crystallization event was investigated
807     separately.
808    
809     \begin{figure}
810     \includegraphics[width=85mm]{povIce.ps}
811     \caption{Crystal structure of an ice 0 lattice shown from the (001) face.}
812     \label{weirdice}
813     \end{figure}
814    
815     The final configurations of these two melting sequences shows an
816     expanded zeolite-like crystal structure that does not correspond to
817     any known form of ice. For convenience and to help distinguish it from
818     the experimentally observed forms of ice, this crystal structure will
819     henceforth be referred to as ice-zero (ice 0). The crystallinity was
820     extensive enough than a near ideal crystal structure could be
821     obtained. Figure \ref{weirdice} shows the repeating crystal structure
822     of a typical crystal at 5 K. The unit cell contains eight molecules,
823     and figure \ref{unitcell} shows a unit cell built from the water
824     particle center of masses that can be used to construct a repeating
825     lattice, similar to figure \ref{weirdice}. Each molecule is hydrogen
826     bonded to four other water molecules; however, the hydrogen bonds are
827     flexed rather than perfectly straight. This results in a skewed
828     tetrahedral geometry about the central molecule. Looking back at
829     figure \ref{isosurface}, it is easy to see how these flexed hydrogen
830     bonds are allowed in that the attractive regions are conical in shape,
831     with the greatest attraction in the central region. Though not ideal,
832     these flexed hydrogen bonds are favorable enough to stabilize an
833     entire crystal generated around them. In fact, the imperfect ice 0
834     crystals were so stable that they melted at greater than room
835     temperature.
836    
837     \begin{figure}
838     \includegraphics[width=65mm]{ice0cell.eps}
839     \caption{Simple unit cell for constructing ice 0. In this cell, $c$ is
840     equal to $0.4714\times a$, and a typical value for $a$ is 8.25 \AA.}
841     \label{unitcell}
842     \end{figure}
843    
844     The initial simulations indicated that ice 0 is the preferred ice
845     structure for at least SSD/E. To verify this, a comparison was made
846     between near ideal crystals of ice $I_h$, ice $I_c$, and ice 0 at
847     constant pressure with SSD/E, SSD/RF, and SSD. Near ideal versions of
848     the three types of crystals were cooled to ~1 K, and the potential
849     energies of each were compared using all three water models. With
850     every water model, ice 0 turned out to have the lowest potential
851     energy: 5\% lower than $I_h$ with SSD, 6.5\% lower with SSD/E, and
852     7.5\% lower with SSD/RF. In all three of these water models, ice $I_c$
853     was observed to be ~2\% less stable than ice $I_h$. In addition to
854     having the lowest potential energy, ice 0 was the most expanded of the
855     three ice crystals, ~5\% less dense than ice $I_h$ with all of the
856     water models. In all three water models, ice $I_c$ was observed to be
857     ~2\% more dense than ice $I_h$.
858    
859     In addition to the low temperature comparisons, melting sequences were
860     performed with ice 0 as the initial configuration using SSD/E, SSD/RF,
861     and SSD both with and without a reaction field. The melting
862     transitions for both SSD/E and SSD without a reaction field occurred
863     at temperature in excess of 375 K. SSD/RF and SSD with a reaction
864     field had more reasonable melting transitions, down near 325 K. These
865     melting point observations emphasize how preferred this crystal
866     structure is over the most common types of ice when using these single
867     point water models.
868    
869     Recognizing that the above tests show ice 0 to be both the most stable
870     and lowest density crystal structure for these single point water
871     models, it is interesting to speculate on the favorability of this
872     crystal structure with the different charge based models. As a quick
873     test, these 3 crystal types were converted from SSD type particles to
874     TIP3P waters and read into CHARMM.\cite{Karplus83} Identical energy
875     minimizations were performed on all of these crystals to compare the
876     system energies. Again, ice 0 was observed to have the lowest total
877     system energy. The total energy of ice 0 was ~2\% lower than ice
878     $I_h$, which was in turn ~3\% lower than ice $I_c$. From these initial
879     results, we would not be surprised if results from the other common
880     water models show ice 0 to be the lowest energy crystal structure. A
881     continuation on work studing ice 0 with multipoint water models will
882     be published in a coming article.
883    
884     \section{Conclusions}
885     The density maximum and temperature dependent transport for the SSD
886     water model, both with and without the use of reaction field, were
887     studied via a series of NPT and NVE simulations. The constant pressure
888     simulations of the melting of both $I_h$ and $I_c$ ice showed a
889     density maximum near 260 K. In most cases, the calculated densities
890     were significantly lower than the densities calculated in simulations
891     of other water models. Analysis of particle diffusion showed SSD to
892     capture the transport properties of experimental very well in both the
893     normal and super-cooled liquid regimes. In order to correct the
894     density behavior, SSD was reparameterized for use both with and
895     without a long-range interaction correction, SSD/RF and SSD/E
896     respectively. In addition to correcting the abnormally low densities,
897     these new versions were show to maintain or improve upon the transport
898     and structural features of the original water model, all while
899     maintaining the fast performance of the SSD water model. This work
900     shows these simple water models, and in particular SSD/E and SSD/RF,
901     to be excellent choices to represent explicit water in future
902     simulations of biochemical systems.
903    
904     \section{Acknowledgments}
905     The authors would like to thank the National Science Foundation for
906     funding under grant CHE-0134881. Computation time was provided by the
907     Notre Dame Bunch-of-Boxes (B.o.B) computer cluster under NSF grant DMR
908     00 79647.
909    
910     \bibliographystyle{jcp}
911    
912     \bibliography{nptSSD}
913    
914     %\pagebreak
915    
916     \end{document}