ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/ssdePaper/nptSSD.tex
Revision: 863
Committed: Thu Nov 13 15:55:20 2003 UTC (20 years, 7 months ago) by chrisfen
Content type: application/x-tex
File size: 49961 byte(s)
Log Message:
Changed ice 0 to ice i.

File Contents

# User Rev Content
1 chrisfen 861 %\documentclass[prb,aps,times,twocolumn,tabularx]{revtex4}
2 chrisfen 862 \documentclass[11pt]{article}
3     \usepackage{endfloat}
4 chrisfen 743 \usepackage{amsmath}
5 chrisfen 862 \usepackage{epsf}
6     \usepackage{berkeley}
7     \usepackage{setspace}
8     \usepackage{tabularx}
9 chrisfen 743 \usepackage{graphicx}
10 chrisfen 862 \usepackage[ref]{overcite}
11 chrisfen 743 %\usepackage{berkeley}
12     %\usepackage{curves}
13 chrisfen 862 \pagestyle{plain}
14     \pagenumbering{arabic}
15     \oddsidemargin 0.0cm \evensidemargin 0.0cm
16     \topmargin -21pt \headsep 10pt
17     \textheight 9.0in \textwidth 6.5in
18     \brokenpenalty=10000
19     \renewcommand{\baselinestretch}{1.2}
20     \renewcommand\citemid{\ } % no comma in optional reference note
21 chrisfen 743
22     \begin{document}
23    
24 chrisfen 759 \title{On the temperature dependent properties of the soft sticky dipole (SSD) and related single point water models}
25 chrisfen 743
26 chrisfen 862 \author{Christopher J. Fennell and J. Daniel Gezelter\footnote{Corresponding author. \ Electronic mail: gezelter@nd.edu} \\
27     Department of Chemistry and Biochemistry\\ University of Notre Dame\\
28 chrisfen 743 Notre Dame, Indiana 46556}
29    
30     \date{\today}
31    
32 chrisfen 862 \maketitle
33    
34 chrisfen 743 \begin{abstract}
35     NVE and NPT molecular dynamics simulations were performed in order to
36     investigate the density maximum and temperature dependent transport
37 chrisfen 856 for SSD and related water models, both with and without the use of
38     reaction field. The constant pressure simulations of the melting of
39     both $I_h$ and $I_c$ ice showed a density maximum near 260 K. In most
40     cases, the calculated densities were significantly lower than the
41     densities calculated in simulations of other water models. Analysis of
42     particle diffusion showed SSD to capture the transport properties of
43 chrisfen 861 experimental water very well in both the normal and super-cooled
44     liquid regimes. In order to correct the density behavior, SSD was
45 chrisfen 743 reparameterized for use both with and without a long-range interaction
46 chrisfen 861 correction, SSD/RF and SSD/E respectively. Compared to the density
47     corrected version of SSD (SSD1), these modified models were shown to
48     maintain or improve upon the structural and transport properties.
49 chrisfen 743 \end{abstract}
50    
51 chrisfen 862 \newpage
52 chrisfen 743
53     %\narrowtext
54    
55    
56     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
57     % BODY OF TEXT
58     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
59    
60     \section{Introduction}
61    
62 chrisfen 862 One of the most important tasks in the simulation of biochemical
63     systems is the proper depiction of water and water solvation. In fact,
64     the bulk of the calculations performed in solvated simulations are of
65 chrisfen 743 interactions with or between solvent molecules. Thus, the outcomes of
66     these types of simulations are highly dependent on the physical
67 chrisfen 862 properties of water, both as individual molecules and in clusters or
68     bulk. Due to the fact that explicit solvent accounts for a massive
69     portion of the calculations, it necessary to simplify the solvent to
70     some extent in order to complete simulations in a reasonable amount of
71     time. In the case of simulating water in biomolecular studies, the
72     balance between accurate properties and computational efficiency is
73     especially delicate, and it has resulted in a variety of different
74     water models.\cite{Jorgensen83,Berendsen87,Jorgensen00} Many of these
75     models predict specific properties more accurately than their
76     predecessors, but often at the cost of other properties or of computer
77     time. As an example, compare TIP3P or TIP4P to TIP5P. TIP5P improves
78     upon the structural and transport properties of water relative to the
79     previous TIP models, yet this comes at a greater than 50\% increase in
80 chrisfen 743 computational cost.\cite{Jorgensen01,Jorgensen00} One recently
81 chrisfen 862 developed model that succeeds in both retaining the accuracy of system
82 chrisfen 743 properties and simplifying calculations to increase computational
83     efficiency is the Soft Sticky Dipole water model.\cite{Ichiye96}
84    
85     The Soft Sticky Dipole (SSD)\ water model was developed by Ichiye
86     \emph{et al.} as a modified form of the hard-sphere water model
87     proposed by Bratko, Blum, and Luzar.\cite{Bratko85,Bratko95} SSD
88     consists of a single point dipole with a Lennard-Jones core and a
89     sticky potential that directs the particles to assume the proper
90     hydrogen bond orientation in the first solvation shell. Thus, the
91     interaction between two SSD water molecules \emph{i} and \emph{j} is
92     given by the potential
93     \begin{equation}
94     u_{ij} = u_{ij}^{LJ} (r_{ij})\ + u_{ij}^{dp}
95     (\mathbf{r}_{ij},\boldsymbol{\Omega}_i,\boldsymbol{\Omega}_j)\ +
96     u_{ij}^{sp}
97     (\mathbf{r}_{ij},\boldsymbol{\Omega}_i,\boldsymbol{\Omega}_j),
98     \end{equation}
99     where the $\mathbf{r}_{ij}$ is the position vector between molecules
100 chrisfen 862 \emph{i} and \emph{j} with magnitude equal to the distance $r_{ij}$, and
101 chrisfen 743 $\boldsymbol{\Omega}_i$ and $\boldsymbol{\Omega}_j$ represent the
102     orientations of the respective molecules. The Lennard-Jones, dipole,
103     and sticky parts of the potential are giving by the following
104 chrisfen 862 equations:
105 chrisfen 743 \begin{equation}
106     u_{ij}^{LJ}(r_{ij}) = 4\epsilon \left[\left(\frac{\sigma}{r_{ij}}\right)^{12}-\left(\frac{\sigma}{r_{ij}}\right)^{6}\right],
107     \end{equation}
108     \begin{equation}
109     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}\ ,
110     \end{equation}
111     \begin{equation}
112     u_{ij}^{sp}
113 chrisfen 862 (\mathbf{r}_{ij},\boldsymbol{\Omega}_i,\boldsymbol{\Omega}_j) =
114     \frac{\nu_0}{2}[s(r_{ij})w(\mathbf{r}_{ij},\boldsymbol{\Omega}_i,\boldsymbol{\Omega}_j) + s^\prime(r_{ij})w^\prime(\mathbf{r}_{ij},\boldsymbol{\Omega}_i,\boldsymbol{\Omega}_j)]\ ,
115 chrisfen 743 \end{equation}
116     where $\boldsymbol{\mu}_i$ and $\boldsymbol{\mu}_j$ are the dipole
117     unit vectors of particles \emph{i} and \emph{j} with magnitude 2.35 D,
118 chrisfen 862 $\nu_0$ scales the strength of the overall sticky potential, and $s$
119     and $s^\prime$ are cubic switching functions. The $w$ and $w^\prime$
120     functions take the following forms:
121 chrisfen 743 \begin{equation}
122     w(\mathbf{r}_{ij},\boldsymbol{\Omega}_i,\boldsymbol{\Omega}_j)=\sin\theta_{ij}\sin2\theta_{ij}\cos2\phi_{ij},
123     \end{equation}
124     \begin{equation}
125     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,
126     \end{equation}
127     where $w^0 = 0.07715$. The $w$ function is the tetrahedral attractive
128     term that promotes hydrogen bonding orientations within the first
129     solvation shell, and $w^\prime$ is a dipolar repulsion term that
130     repels unrealistic dipolar arrangements within the first solvation
131     shell. A more detailed description of the functional parts and
132     variables in this potential can be found in other
133     articles.\cite{Ichiye96,Ichiye99}
134    
135     Being that this is a one-site point dipole model, the actual force
136     calculations are simplified significantly. In the original Monte Carlo
137 chrisfen 862 simulations using this model, Ichiye \emph{et al.} reported an
138     increase in calculation efficiency of up to an order of magnitude over
139     other comparable models, while maintaining the structural behavior of
140 chrisfen 777 water.\cite{Ichiye96} In the original molecular dynamics studies, it
141     was shown that SSD improves on the prediction of many of water's
142     dynamical properties over TIP3P and SPC/E.\cite{Ichiye99} This
143 chrisfen 743 attractive combination of speed and accurate depiction of solvent
144     properties makes SSD a model of interest for the simulation of large
145 chrisfen 861 scale biological systems, such as membrane phase behavior.
146 chrisfen 743
147 chrisfen 757 One of the key limitations of this water model, however, is that it
148     has been parameterized for use with the Ewald Sum technique for the
149     handling of long-ranged interactions. When studying very large
150     systems, the Ewald summation and even particle-mesh Ewald become
151 chrisfen 862 computational burdens, with their respective ideal $N^\frac{3}{2}$ and
152 chrisfen 757 $N\log N$ calculation scaling orders for $N$ particles.\cite{Darden99}
153 chrisfen 759 In applying this water model in these types of systems, it would be
154     useful to know its properties and behavior with the more
155     computationally efficient reaction field (RF) technique, and even with
156 chrisfen 862 a cutoff that lacks any form of long-range correction. This study
157 chrisfen 759 addresses these issues by looking at the structural and transport
158 chrisfen 862 behavior of SSD over a variety of temperatures with the purpose of
159     utilizing the RF correction technique. We then suggest alterations to
160     the parameters that result in more water-like behavior. It should be
161     noted that in a recent publication, some of the original investigators of
162     the SSD water model have put forth adjustments to the SSD water model
163     to address abnormal density behavior (also observed here), calling the
164     corrected model SSD1.\cite{Ichiye03} This study will make comparisons
165     with SSD1's behavior with the goal of improving upon the
166     depiction of water under conditions without the Ewald Sum.
167 chrisfen 757
168 chrisfen 743 \section{Methods}
169    
170 chrisfen 862 As stated previously, the long-range dipole-dipole interactions were
171     accounted for in this study by using the reaction field method. The
172 chrisfen 743 magnitude of the reaction field acting on dipole \emph{i} is given by
173     \begin{equation}
174     \mathcal{E}_{i} = \frac{2(\varepsilon_{s} - 1)}{2\varepsilon_{s} + 1}
175     \frac{1}{r_{c}^{3}} \sum_{j\in{\mathcal{R}}} \boldsymbol{\mu}_{j} f(r_{ij})\ ,
176     \label{rfequation}
177     \end{equation}
178     where $\mathcal{R}$ is the cavity defined by the cutoff radius
179     ($r_{c}$), $\varepsilon_{s}$ is the dielectric constant imposed on the
180     system (80 in this case), $\boldsymbol{\mu}_{j}$ is the dipole moment
181     vector of particle \emph{j}, and $f(r_{ij})$ is a cubic switching
182     function.\cite{AllenTildesley} The reaction field contribution to the
183     total energy by particle \emph{i} is given by
184     $-\frac{1}{2}\boldsymbol{\mu}_{i}\cdot\mathcal{E}_{i}$ and the torque
185     on dipole \emph{i} by
186     $\boldsymbol{\mu}_{i}\times\mathcal{E}_{i}$.\cite{AllenTildesley} Use
187     of reaction field is known to alter the orientational dynamic
188     properties, such as the dielectric relaxation time, based on changes
189     in the length of the cutoff radius.\cite{Berendsen98} This variable
190     behavior makes reaction field a less attractive method than other
191     methods, like the Ewald summation; however, for the simulation of
192 chrisfen 862 large-scale systems, the computational cost benefit of reaction field
193 chrisfen 743 is dramatic. To address some of the dynamical property alterations due
194     to the use of reaction field, simulations were also performed without
195 chrisfen 862 a surrounding dielectric and suggestions are presented on how to make
196 chrisfen 861 SSD more accurate both with and without a reaction field.
197 chrisfen 777
198 chrisfen 743 Simulations were performed in both the isobaric-isothermal and
199     microcanonical ensembles. The constant pressure simulations were
200     implemented using an integral thermostat and barostat as outlined by
201 chrisfen 777 Hoover.\cite{Hoover85,Hoover86} All particles were treated as
202     non-linear rigid bodies. Vibrational constraints are not necessary in
203     simulations of SSD, because there are no explicit hydrogen atoms, and
204     thus no molecular vibrational modes need to be considered.
205 chrisfen 743
206     Integration of the equations of motion was carried out using the
207     symplectic splitting method proposed by Dullweber \emph{et
208 chrisfen 862 al.}\cite{Dullweber1997} The reason for this integrator selection
209 chrisfen 743 deals with poor energy conservation of rigid body systems using
210     quaternions. While quaternions work well for orientational motion in
211     alternate ensembles, the microcanonical ensemble has a constant energy
212 chrisfen 777 requirement that is quite sensitive to errors in the equations of
213     motion. The original implementation of this code utilized quaternions
214     for rotational motion propagation; however, a detailed investigation
215     showed that they resulted in a steady drift in the total energy,
216     something that has been observed by others.\cite{Laird97}
217 chrisfen 743
218     The key difference in the integration method proposed by Dullweber
219     \emph{et al.} is that the entire rotation matrix is propagated from
220     one time step to the next. In the past, this would not have been as
221 chrisfen 862 feasible an option, being that the rotation matrix for a single body is
222 chrisfen 743 nine elements long as opposed to 3 or 4 elements for Euler angles and
223     quaternions respectively. System memory has become much less of an
224     issue in recent times, and this has resulted in substantial benefits
225 chrisfen 759 in energy conservation. There is still the issue of 5 or 6 additional
226     elements for describing the orientation of each particle, which will
227     increase dump files substantially. Simply translating the rotation
228     matrix into its component Euler angles or quaternions for storage
229     purposes relieves this burden.
230 chrisfen 743
231     The symplectic splitting method allows for Verlet style integration of
232 chrisfen 862 both linear and angular motion of rigid bodies. In this integration
233 chrisfen 743 method, the orientational propagation involves a sequence of matrix
234     evaluations to update the rotation matrix.\cite{Dullweber1997} These
235 chrisfen 862 matrix rotations are more costly computationally than the simpler
236     arithmetic quaternion propagation. With the same time step, a 1000 SSD
237     particle simulation shows an average 7\% increase in computation time
238     using the symplectic step method in place of quaternions. This cost is
239     more than justified when comparing the energy conservation of the two
240     methods as illustrated in figure \ref{timestep}.
241 chrisfen 743
242     \begin{figure}
243 chrisfen 862 \begin{center}
244     \epsfxsize=6in
245     \epsfbox{timeStep.epsi}
246 chrisfen 743 \caption{Energy conservation using quaternion based integration versus
247     the symplectic step method proposed by Dullweber \emph{et al.} with
248 chrisfen 862 increasing time step. The larger time step plots are shifted up from
249     the true energy baseline (that of $\Delta t$ = 0.1 fs) for clarity.}
250 chrisfen 743 \label{timestep}
251 chrisfen 862 \end{center}
252 chrisfen 743 \end{figure}
253    
254     In figure \ref{timestep}, the resulting energy drift at various time
255     steps for both the symplectic step and quaternion integration schemes
256     is compared. All of the 1000 SSD particle simulations started with the
257 chrisfen 862 same configuration, and the only difference was the method used to
258     handle rotational motion. At time steps of 0.1 and 0.5 fs, both
259 chrisfen 743 methods for propagating particle rotation conserve energy fairly well,
260     with the quaternion method showing a slight energy drift over time in
261     the 0.5 fs time step simulation. At time steps of 1 and 2 fs, the
262     energy conservation benefits of the symplectic step method are clearly
263 chrisfen 759 demonstrated. Thus, while maintaining the same degree of energy
264     conservation, one can take considerably longer time steps, leading to
265     an overall reduction in computation time.
266 chrisfen 743
267 chrisfen 862 Energy drift in the symplectic step simulations was unnoticeable for
268 chrisfen 743 time steps up to three femtoseconds. A slight energy drift on the
269     order of 0.012 kcal/mol per nanosecond was observed at a time step of
270     four femtoseconds, and as expected, this drift increases dramatically
271     with increasing time step. To insure accuracy in the constant energy
272     simulations, time steps were set at 2 fs and kept at this value for
273     constant pressure simulations as well.
274    
275     Ice crystals in both the $I_h$ and $I_c$ lattices were generated as
276 chrisfen 862 starting points for all simulations. The $I_h$ crystals were formed by
277     first arranging the centers of mass of the SSD particles into a
278     ``hexagonal'' ice lattice of 1024 particles. Because of the crystal
279     structure of $I_h$ ice, the simulation box assumed a rectangular shape
280     with an edge length ratio of approximately
281 chrisfen 743 1.00$\times$1.06$\times$1.23. The particles were then allowed to
282     orient freely about fixed positions with angular momenta randomized at
283     400 K for varying times. The rotational temperature was then scaled
284 chrisfen 862 down in stages to slowly cool the crystals to 25 K. The particles were
285     then allowed to translate with fixed orientations at a constant
286 chrisfen 743 pressure of 1 atm for 50 ps at 25 K. Finally, all constraints were
287     removed and the ice crystals were allowed to equilibrate for 50 ps at
288     25 K and a constant pressure of 1 atm. This procedure resulted in
289     structurally stable $I_h$ ice crystals that obey the Bernal-Fowler
290 chrisfen 862 rules.\cite{Bernal33,Rahman72} This method was also utilized in the
291 chrisfen 743 making of diamond lattice $I_c$ ice crystals, with each cubic
292     simulation box consisting of either 512 or 1000 particles. Only
293     isotropic volume fluctuations were performed under constant pressure,
294     so the ratio of edge lengths remained constant throughout the
295     simulations.
296    
297     \section{Results and discussion}
298    
299     Melting studies were performed on the randomized ice crystals using
300 chrisfen 862 constant pressure and temperature dynamics. During melting
301     simulations, the melting transition and the density maximum can both
302     be observed, provided that the density maximum occurs in the liquid
303     and not the supercooled regime. An ensemble average from five separate
304     melting simulations was acquired, each starting from different ice
305     crystals generated as described previously. All simulations were
306     equilibrated for 100 ps prior to a 200 ps data collection run at each
307     temperature setting. The temperature range of study spanned from 25 to
308     400 K, with a maximum degree increment of 25 K. For regions of
309     interest along this stepwise progression, the temperature increment
310     was decreased from 25 K to 10 and 5 K. The above equilibration and
311     production times were sufficient in that the system volume
312     fluctuations dampened out in all but the very cold simulations (below
313     225 K).
314 chrisfen 743
315     \subsection{Density Behavior}
316 chrisfen 861 Initial simulations focused on the original SSD water model, and an
317     average density versus temperature plot is shown in figure
318     \ref{dense1}. Note that the density maximum when using a reaction
319     field appears between 255 and 265 K, where the calculated densities
320     within this range were nearly indistinguishable. The greater certainty
321     of the average value at 260 K makes a good argument for the actual
322     density maximum residing at this midpoint value. Figure \ref{dense1}
323     was constructed using ice $I_h$ crystals for the initial
324 chrisfen 862 configuration; though not pictured, the simulations starting from ice
325     $I_c$ crystal configurations showed similar results, with a
326 chrisfen 861 liquid-phase density maximum in this same region (between 255 and 260
327     K). In addition, the $I_c$ crystals are more fragile than the $I_h$
328 chrisfen 862 crystals, leading to deformation into a dense glassy state at lower
329 chrisfen 861 temperatures. This resulted in an overall low temperature density
330 chrisfen 862 maximum at 200 K, while still retaining a liquid state density maximum
331     in common with the $I_h$ simulations.
332 chrisfen 743
333     \begin{figure}
334 chrisfen 862 \begin{center}
335     \epsfxsize=6in
336     \epsfbox{denseSSD.eps}
337     \caption{Density versus temperature for TIP4P,\cite{Jorgensen98b}
338     TIP3P,\cite{Jorgensen98b} SPC/E,\cite{Clancy94} SSD without Reaction
339     Field, SSD, and experiment.\cite{CRC80} The arrows indicate the
340 chrisfen 856 change in densities observed when turning off the reaction field. The
341     the lower than expected densities for the SSD model were what
342     prompted the original reparameterization.\cite{Ichiye03}}
343 chrisfen 861 \label{dense1}
344 chrisfen 862 \end{center}
345 chrisfen 743 \end{figure}
346    
347     The density maximum for SSD actually compares quite favorably to other
348 chrisfen 861 simple water models. Figure \ref{dense1} also shows calculated
349     densities of several other models and experiment obtained from other
350 chrisfen 743 sources.\cite{Jorgensen98b,Clancy94,CRC80} Of the listed simple water
351     models, SSD has results closest to the experimentally observed water
352     density maximum. Of the listed water models, TIP4P has a density
353 chrisfen 861 maximum behavior most like that seen in SSD. Though not included in
354     this particular plot, it is useful to note that TIP5P has a water
355     density maximum nearly identical to experiment.
356 chrisfen 743
357     It has been observed that densities are dependent on the cutoff radius
358     used for a variety of water models in simulations both with and
359     without the use of reaction field.\cite{Berendsen98} In order to
360     address the possible affect of cutoff radius, simulations were
361     performed with a dipolar cutoff radius of 12.0 \AA\ to compliment the
362     previous SSD simulations, all performed with a cutoff of 9.0 \AA. All
363 chrisfen 862 of the resulting densities overlapped within error and showed no
364     significant trend toward lower or higher densities as a function of
365     cutoff radius, for simulations both with and without reaction
366     field. These results indicate that there is no major benefit in
367     choosing a longer cutoff radius in simulations using SSD. This is
368     advantageous in that the use of a longer cutoff radius results in
369     significant increases in the time required to obtain a single
370     trajectory.
371 chrisfen 743
372 chrisfen 862 The key feature to recognize in figure \ref{dense1} is the density
373     scaling of SSD relative to other common models at any given
374 chrisfen 861 temperature. Note that the SSD model assumes a lower density than any
375     of the other listed models at the same pressure, behavior which is
376     especially apparent at temperatures greater than 300 K. Lower than
377 chrisfen 862 expected densities have been observed for other systems using a
378     reaction field for long-range electrostatic interactions, so the most
379     likely reason for the significantly lower densities seen in these
380 chrisfen 861 simulations is the presence of the reaction
381     field.\cite{Berendsen98,Nezbeda02} In order to test the effect of the
382     reaction field on the density of the systems, the simulations were
383     repeated without a reaction field present. The results of these
384     simulations are also displayed in figure \ref{dense1}. Without
385 chrisfen 862 reaction field, the densities increase considerably to more
386 chrisfen 861 experimentally reasonable values, especially around the freezing point
387     of liquid water. The shape of the curve is similar to the curve
388     produced from SSD simulations using reaction field, specifically the
389     rapidly decreasing densities at higher temperatures; however, a shift
390 chrisfen 862 in the density maximum location, down to 245 K, is observed. This is a
391     more accurate comparison to the other listed water models, in that no
392     long range corrections were applied in those
393 chrisfen 861 simulations.\cite{Clancy94,Jorgensen98b} However, even without a
394     reaction field, the density around 300 K is still significantly lower
395     than experiment and comparable water models. This anomalous behavior
396     was what lead Ichiye \emph{et al.} to recently reparameterize SSD and
397     make SSD1.\cite{Ichiye03} In discussing potential adjustments later in
398     this paper, all comparisons were performed with this new model.
399    
400 chrisfen 743 \subsection{Transport Behavior}
401     Of importance in these types of studies are the transport properties
402 chrisfen 862 of the particles and their change in responce to altering
403     environmental conditions. In order to probe transport, constant energy
404     simulations were performed about the average density uncovered by the
405     constant pressure simulations. Simulations started with randomized
406     velocities and underwent 50 ps of temperature scaling and 50 ps of
407     constant energy equilibration before obtaining a 200 ps
408     trajectory. Diffusion constants were calculated via root-mean square
409     deviation analysis. The averaged results from five sets of NVE
410     simulations are displayed in figure \ref{diffuse}, alongside
411     experimental, SPC/E, and TIP5P
412     results.\cite{Gillen72,Mills73,Clancy94,Jorgensen01}
413 chrisfen 743
414     \begin{figure}
415 chrisfen 862 \begin{center}
416     \epsfxsize=6in
417     \epsfbox{betterDiffuse.epsi}
418 chrisfen 743 \caption{Average diffusion coefficient over increasing temperature for
419 chrisfen 862 SSD, SPC/E,\cite{Clancy94} TIP5P,\cite{Jorgensen01} and Experimental
420     data.\cite{Gillen72,Mills73} Of the three water models shown, SSD has
421     the least deviation from the experimental values. The rapidly
422     increasing diffusion constants for TIP5P and SSD correspond to
423     significant decrease in density at the higher temperatures.}
424 chrisfen 743 \label{diffuse}
425 chrisfen 862 \end{center}
426 chrisfen 743 \end{figure}
427    
428     The observed values for the diffusion constant point out one of the
429     strengths of the SSD model. Of the three experimental models shown,
430     the SSD model has the most accurate depiction of the diffusion trend
431 chrisfen 862 seen in experiment in both the supercooled and liquid temperature
432     regimes. SPC/E does a respectable job by producing values similar to
433     SSD and experiment around 290 K; however, it deviates at both higher
434     and lower temperatures, failing to predict the experimental
435     trend. TIP5P and SSD both start off low at colder temperatures and
436     tend to diffuse too rapidly at higher temperatures. This trend at
437     higher temperatures is not surprising in that the densities of both
438     TIP5P and SSD are lower than experimental water at these higher
439     temperatures. When calculating the diffusion coefficients for SSD at
440 chrisfen 743 experimental densities, the resulting values fall more in line with
441 chrisfen 861 experiment at these temperatures, albeit not at standard pressure.
442 chrisfen 743
443     \subsection{Structural Changes and Characterization}
444     By starting the simulations from the crystalline state, the melting
445     transition and the ice structure can be studied along with the liquid
446 chrisfen 862 phase behavior beyond the melting point. The constant pressure heat
447     capacity (C$_\text{p}$) was monitored to locate the melting transition
448     in each of the simulations. In the melting simulations of the 1024
449     particle ice $I_h$ simulations, a large spike in C$_\text{p}$ occurs
450     at 245 K, indicating a first order phase transition for the melting of
451     these ice crystals. When the reaction field is turned off, the melting
452     transition occurs at 235 K. These melting transitions are
453     considerably lower than the experimental value, but this is not a
454     surprise considering the simplicity of the SSD model.
455 chrisfen 743
456 chrisfen 862 \begin{figure}
457     \begin{center}
458     \epsfxsize=6in
459     \epsfbox{corrDiag.eps}
460     \caption{Two dimensional illustration of angles involved in the
461     correlations observed in figure \ref{contour}.}
462     \label{corrAngle}
463     \end{center}
464     \end{figure}
465    
466     \begin{figure}
467     \begin{center}
468     \epsfxsize=6in
469     \epsfbox{fullContours.eps}
470 chrisfen 743 \caption{Contour plots of 2D angular g($r$)'s for 512 SSD systems at
471     100 K (A \& B) and 300 K (C \& D). Contour colors are inverted for
472     clarity: dark areas signify peaks while light areas signify
473     depressions. White areas have g(\emph{r}) values below 0.5 and black
474     areas have values above 1.5.}
475     \label{contour}
476 chrisfen 862 \end{center}
477 chrisfen 743 \end{figure}
478    
479 chrisfen 861 Additional analysis of the melting phase-transition process was
480     performed by using two-dimensional structure and dipole angle
481     correlations. Expressions for these correlations are as follows:
482    
483 chrisfen 862 \begin{equation}
484     g_{\text{AB}}(r,\cos\theta) = \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\ ,
485     \end{equation}
486     \begin{equation}
487     g_{\text{AB}}(r,\cos\omega) =
488 chrisfen 743 \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\ ,
489 chrisfen 862 \end{equation}
490 chrisfen 861 where $\theta$ and $\omega$ refer to the angles shown in figure
491     \ref{corrAngle}. By binning over both distance and the cosine of the
492 chrisfen 743 desired angle between the two dipoles, the g(\emph{r}) can be
493     dissected to determine the common dipole arrangements that constitute
494     the peaks and troughs. Frames A and B of figure \ref{contour} show a
495     relatively crystalline state of an ice $I_c$ simulation. The first
496 chrisfen 861 peak of the g(\emph{r}) consists primarily of the preferred hydrogen
497     bonding arrangements as dictated by the tetrahedral sticky potential -
498 chrisfen 743 one peak for the donating and the other for the accepting hydrogen
499     bonds. Due to the high degree of crystallinity of the sample, the
500     second and third solvation shells show a repeated peak arrangement
501     which decays at distances around the fourth solvation shell, near the
502     imposed cutoff for the Lennard-Jones and dipole-dipole interactions.
503 chrisfen 861 In the higher temperature simulation shown in frames C and D, these
504     longer-ranged repeated peak features deteriorate rapidly. The first
505     solvation shell still shows the strong effect of the sticky-potential,
506     although it covers a larger area, extending to include a fraction of
507     aligned dipole peaks within the first solvation shell. The latter
508     peaks lose definition as thermal motion and the competing dipole force
509     overcomes the sticky potential's tight tetrahedral structuring of the
510     fluid.
511 chrisfen 743
512     This complex interplay between dipole and sticky interactions was
513     remarked upon as a possible reason for the split second peak in the
514     oxygen-oxygen g(\emph{r}).\cite{Ichiye96} At low temperatures, the
515 chrisfen 862 second solvation shell peak appears to have two distinct components
516     that blend together to form one observable peak. At higher
517     temperatures, this split character alters to show the leading 4 \AA\
518     peak dominated by equatorial anti-parallel dipole orientations. There
519     is also a tightly bunched group of axially arranged dipoles that most
520     likely consist of the smaller fraction of aligned dipole pairs. The
521     trailing component of the split peak at 5 \AA\ is dominated by aligned
522     dipoles that assume hydrogen bond arrangements similar to those seen
523     in the first solvation shell. This evidence indicates that the dipole
524     pair interaction begins to dominate outside of the range of the
525     dipolar repulsion term. Primary energetically favorable dipole
526     arrangements populate the region immediately outside this repulsion
527     region (around 4 \AA), while arrangements that seek to ideally satisfy
528     both the sticky and dipole forces locate themselves just beyond this
529     initial buildup (around 5 \AA).
530 chrisfen 743
531     From these findings, the split second peak is primarily the product of
532 chrisfen 861 the dipolar repulsion term of the sticky potential. In fact, the inner
533     peak can be pushed out and merged with the outer split peak just by
534     extending the switching function cutoff ($s^\prime(r_{ij})$) from its
535     normal 4.0 \AA\ to values of 4.5 or even 5 \AA. This type of
536     correction is not recommended for improving the liquid structure,
537 chrisfen 862 since the second solvation shell would still be shifted too far
538 chrisfen 861 out. In addition, this would have an even more detrimental effect on
539     the system densities, leading to a liquid with a more open structure
540     and a density considerably lower than the normal SSD behavior shown
541     previously. A better correction would be to include the
542     quadrupole-quadrupole interactions for the water particles outside of
543     the first solvation shell, but this reduces the simplicity and speed
544     advantage of SSD.
545 chrisfen 743
546 chrisfen 861 \subsection{Adjusted Potentials: SSD/RF and SSD/E}
547 chrisfen 743 The propensity of SSD to adopt lower than expected densities under
548     varying conditions is troubling, especially at higher temperatures. In
549 chrisfen 861 order to correct this model for use with a reaction field, it is
550     necessary to adjust the force field parameters for the primary
551     intermolecular interactions. In undergoing a reparameterization, it is
552     important not to focus on just one property and neglect the other
553     important properties. In this case, it would be ideal to correct the
554     densities while maintaining the accurate transport properties.
555 chrisfen 743
556 chrisfen 862 The parameters available for tuning include the $\sigma$ and $\epsilon$
557 chrisfen 743 Lennard-Jones parameters, the dipole strength ($\mu$), and the sticky
558     attractive and dipole repulsive terms with their respective
559     cutoffs. To alter the attractive and repulsive terms of the sticky
560     potential independently, it is necessary to separate the terms as
561     follows:
562     \begin{equation}
563     u_{ij}^{sp}
564 chrisfen 862 (\mathbf{r}_{ij},\boldsymbol{\Omega}_i,\boldsymbol{\Omega}_j) =
565     \frac{\nu_0}{2}[s(r_{ij})w(\mathbf{r}_{ij},\boldsymbol{\Omega}_i,\boldsymbol{\Omega}_j)] + \frac{\nu_0^\prime}{2} [s^\prime(r_{ij})w^\prime(\mathbf{r}_{ij},\boldsymbol{\Omega}_i,\boldsymbol{\Omega}_j)],
566 chrisfen 743 \end{equation}
567    
568     where $\nu_0$ scales the strength of the tetrahedral attraction and
569     $\nu_0^\prime$ acts in an identical fashion on the dipole repulsion
570 chrisfen 862 term. The separation was performed for purposes of the
571     reparameterization, but the final parameters were adjusted so that it
572     is unnecessary to separate the terms when implementing the adjusted
573     water potentials. The results of the reparameterizations are shown in
574     table \ref{params}. Note that the tetrahedral attractive and dipolar
575     repulsive terms do not share the same lower cutoff ($r_l$) in the
576     newly parameterized potentials - soft sticky dipole reaction field
577     (SSD/RF - for use with a reaction field) and soft sticky dipole
578     enhanced (SSD/E - an attempt to improve the liquid structure in
579     simulations without a long-range correction).
580 chrisfen 743
581     \begin{table}
582 chrisfen 862 \begin{center}
583 chrisfen 743 \caption{Parameters for the original and adjusted models}
584 chrisfen 856 \begin{tabular}{ l c c c c }
585 chrisfen 743 \hline \\[-3mm]
586 chrisfen 862 \ \ \ Parameters\ \ \ & \ \ \ SSD\cite{Ichiye96} \ \ \ & \ SSD1\cite{Ichiye03}\ \ & \ SSD/E\ \ & \ SSD/RF \\
587 chrisfen 743 \hline \\[-3mm]
588 chrisfen 856 \ \ \ $\sigma$ (\AA) & 3.051 & 3.016 & 3.035 & 3.019\\
589     \ \ \ $\epsilon$ (kcal/mol) & 0.152 & 0.152 & 0.152 & 0.152\\
590     \ \ \ $\mu$ (D) & 2.35 & 2.35 & 2.42 & 2.48\\
591     \ \ \ $\nu_0$ (kcal/mol) & 3.7284 & 3.6613 & 3.90 & 3.90\\
592     \ \ \ $r_l$ (\AA) & 2.75 & 2.75 & 2.40 & 2.40\\
593     \ \ \ $r_u$ (\AA) & 3.35 & 3.35 & 3.80 & 3.80\\
594     \ \ \ $\nu_0^\prime$ (kcal/mol) & 3.7284 & 3.6613 & 3.90 & 3.90\\
595     \ \ \ $r_l^\prime$ (\AA) & 2.75 & 2.75 & 2.75 & 2.75\\
596     \ \ \ $r_u^\prime$ (\AA) & 4.00 & 4.00 & 3.35 & 3.35\\
597 chrisfen 743 \end{tabular}
598     \label{params}
599 chrisfen 862 \end{center}
600 chrisfen 743 \end{table}
601    
602 chrisfen 862 \begin{figure}
603     \begin{center}
604     \epsfxsize=5in
605     \epsfbox{GofRCompare.epsi}
606 chrisfen 743 \caption{Plots comparing experiment\cite{Head-Gordon00_1} with SSD/E
607 chrisfen 856 and SSD1 without reaction field (top), as well as SSD/RF and SSD1 with
608 chrisfen 743 reaction field turned on (bottom). The insets show the respective
609 chrisfen 862 first peaks in detail. Note how the changes in parameters have lowered
610     and broadened the first peak of SSD/E and SSD/RF.}
611 chrisfen 743 \label{grcompare}
612 chrisfen 862 \end{center}
613 chrisfen 743 \end{figure}
614    
615 chrisfen 862 \begin{figure}
616     \begin{center}
617     \epsfxsize=6in
618     \epsfbox{dualsticky.ps}
619 chrisfen 856 \caption{Isosurfaces of the sticky potential for SSD1 (left) and SSD/E \&
620 chrisfen 743 SSD/RF (right). Light areas correspond to the tetrahedral attractive
621 chrisfen 862 component, and darker areas correspond to the dipolar repulsive
622     component.}
623 chrisfen 743 \label{isosurface}
624 chrisfen 862 \end{center}
625 chrisfen 743 \end{figure}
626    
627     In the paper detailing the development of SSD, Liu and Ichiye placed
628     particular emphasis on an accurate description of the first solvation
629 chrisfen 862 shell. This resulted in a somewhat tall and narrow first peak in the
630     g(\emph{r}) that integrated to give similar coordination numbers to
631     the experimental data obtained by Soper and
632     Phillips.\cite{Ichiye96,Soper86} New experimental x-ray scattering
633     data from the Head-Gordon lab indicates a slightly lower and shifted
634     first peak in the g$_\mathrm{OO}(r)$, so adjustments to SSD were made
635     while taking into consideration the new experimental
636     findings.\cite{Head-Gordon00_1} Figure \ref{grcompare} shows the
637     relocation of the first peak of the oxygen-oxygen g(\emph{r}) by
638     comparing the revised SSD model (SSD1), SSD-E, and SSD-RF to the new
639     experimental results. Both modified water models have shorter peaks
640     that are brought in more closely to the experimental peak (as seen in
641     the insets of figure \ref{grcompare}). This structural alteration was
642     accomplished by the combined reduction in the Lennard-Jones $\sigma$
643     variable and adjustment of the sticky potential strength and
644     cutoffs. As can be seen in table \ref{params}, the cutoffs for the
645     tetrahedral attractive and dipolar repulsive terms were nearly swapped
646     with each other. Isosurfaces of the original and modified sticky
647     potentials are shown in figure \ref{isosurface}. In these isosurfaces,
648     it is easy to see how altering the cutoffs changes the repulsive and
649     attractive character of the particles. With a reduced repulsive
650     surface (darker region), the particles can move closer to one another,
651     increasing the density for the overall system. This change in
652     interaction cutoff also results in a more gradual orientational motion
653     by allowing the particles to maintain preferred dipolar arrangements
654     before they begin to feel the pull of the tetrahedral
655     restructuring. As the particles move closer together, the dipolar
656     repulsion term becomes active and excludes unphysical nearest-neighbor
657     arrangements. This compares with how SSD and SSD1 exclude preferred
658     dipole alignments before the particles feel the pull of the ``hydrogen
659     bonds''. Aside from improving the shape of the first peak in the
660     g(\emph{r}), this modification improves the densities considerably by
661 chrisfen 861 allowing the persistence of full dipolar character below the previous
662     4.0 \AA\ cutoff.
663 chrisfen 743
664     While adjusting the location and shape of the first peak of
665 chrisfen 861 g(\emph{r}) improves the densities, these changes alone are
666     insufficient to bring the system densities up to the values observed
667 chrisfen 862 experimentally. To further increase the densities, the dipole moments
668     were increased in both of the adjusted models. Since SSD is a dipole
669 chrisfen 861 based model, the structure and transport are very sensitive to changes
670     in the dipole moment. The original SSD simply used the dipole moment
671     calculated from the TIP3P water model, which at 2.35 D is
672 chrisfen 743 significantly greater than the experimental gas phase value of 1.84
673 chrisfen 861 D. The larger dipole moment is a more realistic value and improves the
674 chrisfen 743 dielectric properties of the fluid. Both theoretical and experimental
675     measurements indicate a liquid phase dipole moment ranging from 2.4 D
676 chrisfen 862 to values as high as 3.11 D, providing a substantial range of
677     reasonable values for a dipole
678     moment.\cite{Sprik91,Kusalik02,Badyal00,Barriol64} Moderately
679     increasing the dipole moments to 2.42 and 2.48 D for SSD/E and SSD/RF,
680     respectively, leads to significant changes in the density and
681     transport of the water models.
682 chrisfen 743
683 chrisfen 861 In order to demonstrate the benefits of these reparameterizations, a
684 chrisfen 743 series of NPT and NVE simulations were performed to probe the density
685     and transport properties of the adapted models and compare the results
686     to the original SSD model. This comparison involved full NPT melting
687     sequences for both SSD/E and SSD/RF, as well as NVE transport
688 chrisfen 861 calculations at the calculated self-consistent densities. Again, the
689 chrisfen 862 results are obtained from five separate simulations of 1024 particle
690     systems, and the melting sequences were started from different ice
691     $I_h$ crystals constructed as described previously. Each NPT
692 chrisfen 861 simulation was equilibrated for 100 ps before a 200 ps data collection
693 chrisfen 862 run at each temperature step, and the final configuration from the
694     previous temperature simulation was used as a starting point. All NVE
695     simulations had the same thermalization, equilibration, and data
696     collection times as stated earlier in this paper.
697 chrisfen 743
698 chrisfen 862 \begin{figure}
699     \begin{center}
700     \epsfxsize=6in
701     \epsfbox{ssdeDense.epsi}
702 chrisfen 861 \caption{Comparison of densities calculated with SSD/E to SSD1 without a
703 chrisfen 862 reaction field, TIP3P,\cite{Jorgensen98b} TIP5P,\cite{Jorgensen00}
704     SPC/E,\cite{Clancy94} and experiment.\cite{CRC80} The window shows a
705 chrisfen 856 expansion around 300 K with error bars included to clarify this region
706     of interest. Note that both SSD1 and SSD/E show good agreement with
707     experiment when the long-range correction is neglected.}
708 chrisfen 743 \label{ssdedense}
709 chrisfen 862 \end{center}
710 chrisfen 743 \end{figure}
711    
712 chrisfen 861 Figure \ref{ssdedense} shows the density profile for the SSD/E model
713 chrisfen 862 in comparison to SSD1 without a reaction field, other common water
714     models, and experimental results. The calculated densities for both
715     SSD/E and SSD1 have increased significantly over the original SSD
716     model (see figure \ref{dense1}) and are in better agreement with the
717     experimental values. At 298 K, the densities of SSD/E and SSD1 without
718     a long-range correction are 0.996$\pm$0.001 g/cm$^3$ and
719     0.999$\pm$0.001 g/cm$^3$ respectively. These both compare well with
720     the experimental value of 0.997 g/cm$^3$, and they are considerably
721     better than the SSD value of 0.967$\pm$0.003 g/cm$^3$. The changes to
722     the dipole moment and sticky switching functions have improved the
723     structuring of the liquid (as seen in figure \ref{grcompare}, but they
724     have shifted the density maximum to much lower temperatures. This
725     comes about via an increase in the liquid disorder through the
726     weakening of the sticky potential and strengthening of the dipolar
727     character. However, this increasing disorder in the SSD/E model has
728     little effect on the melting transition. By monitoring C$\text{p}$
729     throughout these simulations, the melting transition for SSD/E was
730     shown to occur at 235 K, the same transition temperature observed with
731     SSD and SSD1.
732 chrisfen 743
733 chrisfen 862 \begin{figure}
734     \begin{center}
735     \epsfxsize=6in
736     \epsfbox{ssdrfDense.epsi}
737 chrisfen 861 \caption{Comparison of densities calculated with SSD/RF to SSD1 with a
738 chrisfen 862 reaction field, TIP3P,\cite{Jorgensen98b} TIP5P,\cite{Jorgensen00}
739     SPC/E,\cite{Clancy94} and experiment.\cite{CRC80} The inset shows the
740 chrisfen 856 necessity of reparameterization when utilizing a reaction field
741     long-ranged correction - SSD/RF provides significantly more accurate
742     densities than SSD1 when performing room temperature simulations.}
743 chrisfen 743 \label{ssdrfdense}
744 chrisfen 862 \end{center}
745 chrisfen 743 \end{figure}
746    
747 chrisfen 862 Including the reaction field long-range correction in the simulations
748     results in a more interesting comparison. A density profile including
749     SSD/RF and SSD1 with an active reaction field is shown in figure
750     \ref{ssdrfdense}. As observed in the simulations without a reaction
751     field, the densities of SSD/RF and SSD1 show a dramatic increase over
752     normal SSD (see figure \ref{dense1}). At 298 K, SSD/RF has a density
753     of 0.997$\pm$0.001 g/cm$^3$, directly in line with experiment and
754     considerably better than the SSD value of 0.941$\pm$0.001 g/cm$^3$ and
755     the SSD1 value of 0.972$\pm$0.002 g/cm$^3$. These results further
756     emphasize the importance of reparameterization in order to model the
757     density properly under different simulation conditions. Again, these
758     changes have only a minor effect on the melting point, which observed
759     at 245 K for SSD/RF, is identical to SSD and only 5 K lower than SSD1
760     with a reaction field. Additionally, the difference in density maxima
761     is not as extreme, with SSD/RF showing a density maximum at 255 K,
762     fairly close to the density maxima of 260 K and 265 K, shown by SSD
763     and SSD1 respectively.
764 chrisfen 743
765 chrisfen 862 \begin{figure}
766     \begin{center}
767     \epsfxsize=6in
768     \epsfbox{ssdeDiffuse.epsi}
769 chrisfen 861 \caption{Plots of the diffusion constants calculated from SSD/E and SSD1,
770 chrisfen 862 both without a reaction field, along with experimental
771     results.\cite{Gillen72,Mills73} The NVE calculations were performed
772     at the average densities observed in the 1 atm NPT simulations for
773     the respective models. SSD/E is slightly more fluid than experiment
774     at all of the temperatures, but it is closer than SSD1 without a
775     long-range correction.}
776 chrisfen 861 \label{ssdediffuse}
777 chrisfen 862 \end{center}
778 chrisfen 861 \end{figure}
779    
780 chrisfen 743 The reparameterization of the SSD water model, both for use with and
781     without an applied long-range correction, brought the densities up to
782     what is expected for simulating liquid water. In addition to improving
783     the densities, it is important that particle transport be maintained
784     or improved. Figure \ref{ssdediffuse} compares the temperature
785 chrisfen 861 dependence of the diffusion constant of SSD/E to SSD1 without an
786     active reaction field, both at the densities calculated at 1 atm and
787     at the experimentally calculated densities for super-cooled and liquid
788 chrisfen 862 water. The diffusion constant for SSD/E is consistently a little
789     higher than experiment, while SSD1 remains lower than experiment until
790     relatively high temperatures (greater than 330 K). Both models follow
791     the shape of the experimental curve well below 300 K but tend to
792     diffuse too rapidly at higher temperatures, something that is
793     especially apparent with SSD1. This accelerated increasing of
794     diffusion is caused by the rapidly decreasing system density with
795     increasing temperature. Though it is difficult to see in figure
796     \ref{ssdedense}, the densities of SSD1 decay more rapidly with
797     temperature than do those of SSD/E, leading to more visible deviation
798     from the experimental diffusion trend. Thus, the changes made to
799     improve the liquid structure may have had an adverse affect on the
800     density maximum, but they improve the transport behavior of SSD/E
801     relative to SSD1.
802 chrisfen 743
803 chrisfen 862 \begin{figure}
804     \begin{center}
805     \epsfxsize=6in
806     \epsfbox{ssdrfDiffuse.epsi}
807 chrisfen 856 \caption{Plots of the diffusion constants calculated from SSD/RF and SSD1,
808 chrisfen 862 both with an active reaction field, along with experimental
809     results.\cite{Gillen72,Mills73} The NVE calculations were performed
810     at the average densities observed in the 1 atm NPT simulations for
811     both of the models. Note how accurately SSD/RF simulates the
812     diffusion of water throughout this temperature range. The more
813     rapidly increasing diffusion constants at high temperatures for both
814     models is attributed to the significantly lower densities than
815     observed in experiment.}
816 chrisfen 856 \label{ssdrfdiffuse}
817 chrisfen 862 \end{center}
818 chrisfen 743 \end{figure}
819    
820     In figure \ref{ssdrfdiffuse}, the diffusion constants for SSD/RF are
821 chrisfen 862 compared to SSD1 with an active reaction field. Note that SSD/RF
822 chrisfen 861 tracks the experimental results incredibly well, identical within
823 chrisfen 862 error throughout the temperature range shown and with only a slight
824 chrisfen 861 increasing trend at higher temperatures. SSD1 tends to diffuse more
825     slowly at low temperatures and deviates to diffuse too rapidly at
826 chrisfen 862 temperatures greater than 330 K. As stated in relation to SSD/E, this
827     deviation away from the ideal trend is due to a rapid decrease in
828     density at higher temperatures. SSD/RF does not suffer from this
829     problem as much as SSD1, because the calculated densities are closer
830     to the experimental value. These results again emphasize the
831     importance of careful reparameterization when using an altered
832 chrisfen 861 long-range correction.
833 chrisfen 743
834     \subsection{Additional Observations}
835    
836     \begin{figure}
837 chrisfen 862 \begin{center}
838     \epsfxsize=6in
839     \epsfbox{povIce.ps}
840     \caption{A water lattice built from the crystal structure assumed by
841     SSD/E when undergoing an extremely restricted temperature NPT
842 chrisfen 863 simulation. This form of ice is referred to as ice \emph{i} to
843     emphasize its simulation origins. This image was taken of the (001)
844     face of the crystal.}
845 chrisfen 743 \label{weirdice}
846 chrisfen 862 \end{center}
847 chrisfen 743 \end{figure}
848    
849 chrisfen 861 While performing restricted temperature melting sequences of SSD/E not
850 chrisfen 862 previously discussed, some interesting observations were made. After
851     melting at 235 K, two of five systems underwent crystallization events
852     near 245 K. As the heating process continued, the two systems remained
853     crystalline until finally melting between 320 and 330 K. The final
854     configurations of these two melting sequences show an expanded
855     zeolite-like crystal structure that does not correspond to any known
856     form of ice. For convenience, and to help distinguish it from the
857     experimentally observed forms of ice, this crystal structure will
858 chrisfen 863 henceforth be referred to as ice $\sqrt{\smash[b]{-\text{I}}}$ (ice
859     \emph{i}). The crystallinity was extensive enough that a near ideal
860     crystal structure of ice \emph{i} could be obtained. Figure
861     \ref{weirdice} shows the repeating crystal structure of a typical
862     crystal at 5 K. Each water molecule is hydrogen bonded to four others;
863     however, the hydrogen bonds are flexed rather than perfectly
864     straight. This results in a skewed tetrahedral geometry about the
865     central molecule. Referring to figure \ref{isosurface}, these flexed
866     hydrogen bonds are allowed due to the conical shape of the attractive
867     regions, with the greatest attraction along the direct hydrogen bond
868     configuration. Though not ideal, these flexed hydrogen bonds are
869     favorable enough to stabilize an entire crystal generated around
870     them. In fact, the imperfect ice \emph{i} crystals were so stable that
871 chrisfen 862 they melted at temperatures nearly 100 K greater than both ice I$_c$
872     and I$_h$.
873 chrisfen 743
874 chrisfen 863 These initial simulations indicated that ice \emph{i} is the preferred
875     ice structure for at least the SSD/E model. To verify this, a
876     comparison was made between near ideal crystals of ice $I_h$, ice
877     $I_c$, and ice 0 at constant pressure with SSD/E, SSD/RF, and
878     SSD1. Near ideal versions of the three types of crystals were cooled
879     to 1 K, and the potential energies of each were compared using all
880     three water models. With every water model, ice \emph{i} turned out to
881     have the lowest potential energy: 5\% lower than $I_h$ with SSD1,
882     6.5\% lower with SSD/E, and 7.5\% lower with SSD/RF.
883 chrisfen 743
884 chrisfen 861 In addition to these low temperature comparisons, melting sequences
885 chrisfen 863 were performed with ice \emph{i} as the initial configuration using
886     SSD/E, SSD/RF, and SSD1 both with and without a reaction field. The
887     melting transitions for both SSD/E and SSD1 without a reaction field
888     occurred at temperature in excess of 375 K. SSD/RF and SSD1 with a
889     reaction field showed more reasonable melting transitions near 325
890     K. These melting point observations emphasize the preference for this
891     crystal structure over the most common types of ice when using these
892     single point water models.
893 chrisfen 743
894 chrisfen 863 Recognizing that the above tests show ice \emph{i} to be both the most
895     stable and lowest density crystal structure for these single point
896     water models, it is interesting to speculate on the relative stability
897     of this crystal structure with charge based water models. As a quick
898 chrisfen 743 test, these 3 crystal types were converted from SSD type particles to
899     TIP3P waters and read into CHARMM.\cite{Karplus83} Identical energy
900 chrisfen 862 minimizations were performed on the crystals to compare the system
901 chrisfen 863 energies. Again, ice \emph{i} was observed to have the lowest total
902     system energy. The total energy of ice \emph{i} was ~2\% lower than
903     ice $I_h$, which was in turn ~3\% lower than ice $I_c$. Based on these
904     initial studies, it would not be surprising if results from the other
905     common water models show ice \emph{i} to be the lowest energy crystal
906     structure. A continuation of this work studying ice \emph{i} with
907     multi-point water models will be published in a coming article.
908 chrisfen 743
909     \section{Conclusions}
910     The density maximum and temperature dependent transport for the SSD
911     water model, both with and without the use of reaction field, were
912     studied via a series of NPT and NVE simulations. The constant pressure
913     simulations of the melting of both $I_h$ and $I_c$ ice showed a
914     density maximum near 260 K. In most cases, the calculated densities
915     were significantly lower than the densities calculated in simulations
916     of other water models. Analysis of particle diffusion showed SSD to
917 chrisfen 862 capture the transport properties of experimental water well in both
918     the liquid and super-cooled liquid regimes. In order to correct the
919 chrisfen 861 density behavior, the original SSD model was reparameterized for use
920     both with and without a reaction field (SSD/RF and SSD/E), and
921     comparison simulations were performed with SSD1, the density corrected
922     version of SSD. Both models improve the liquid structure, density
923     values, and diffusive properties under their respective conditions,
924     indicating the necessity of reparameterization when altering the
925 chrisfen 862 long-range correction specifics. When taking into account the
926     appropriate considerations, these simple water models are excellent
927     choices for representing explicit water in large scale simulations of
928     biochemical systems.
929 chrisfen 743
930     \section{Acknowledgments}
931 chrisfen 777 Support for this project was provided by the National Science
932     Foundation under grant CHE-0134881. Computation time was provided by
933     the Notre Dame Bunch-of-Boxes (B.o.B) computer cluster under NSF grant
934     DMR 00 79647.
935 chrisfen 743
936 chrisfen 862
937     \newpage
938    
939 chrisfen 743 \bibliographystyle{jcp}
940     \bibliography{nptSSD}
941    
942     %\pagebreak
943    
944     \end{document}