ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/ssdePaper/nptSSD.tex
Revision: 1029
Committed: Thu Feb 5 20:47:50 2004 UTC (20 years, 5 months ago) by gezelter
Content type: application/x-tex
File size: 56836 byte(s)
Log Message:
Changes in response to JCP reviewer

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 gezelter 921 \title{On the structural and transport properties of the soft sticky
25 gezelter 1029 dipole ({\sc ssd}) and related single point water models}
26 chrisfen 743
27 chrisfen 862 \author{Christopher J. Fennell and J. Daniel Gezelter\footnote{Corresponding author. \ Electronic mail: gezelter@nd.edu} \\
28     Department of Chemistry and Biochemistry\\ University of Notre Dame\\
29 chrisfen 743 Notre Dame, Indiana 46556}
30    
31     \date{\today}
32    
33 chrisfen 862 \maketitle
34    
35 chrisfen 743 \begin{abstract}
36 gezelter 921 The density maximum and temperature dependence of the self-diffusion
37 gezelter 1029 constant were investigated for the soft sticky dipole ({\sc ssd}) water
38 gezelter 921 model and two related re-parameterizations of this single-point model.
39     A combination of microcanonical and isobaric-isothermal molecular
40     dynamics simulations were used to calculate these properties, both
41     with and without the use of reaction field to handle long-range
42     electrostatics. The isobaric-isothermal (NPT) simulations of the
43     melting of both ice-$I_h$ and ice-$I_c$ showed a density maximum near
44     260 K. In most cases, the use of the reaction field resulted in
45     calculated densities which were were significantly lower than
46     experimental densities. Analysis of self-diffusion constants shows
47 gezelter 1029 that the original {\sc ssd} model captures the transport properties of
48 chrisfen 861 experimental water very well in both the normal and super-cooled
49 gezelter 1029 liquid regimes. We also present our re-parameterized versions of {\sc ssd}
50 gezelter 921 for use both with the reaction field or without any long-range
51 gezelter 1029 electrostatic corrections. These are called the {\sc ssd/rf} and {\sc ssd/e}
52 gezelter 921 models respectively. These modified models were shown to maintain or
53     improve upon the experimental agreement with the structural and
54 gezelter 1029 transport properties that can be obtained with either the original {\sc ssd}
55     or the density corrected version of the original model ({\sc ssd1}).
56 gezelter 921 Additionally, a novel low-density ice structure is presented
57 gezelter 1029 which appears to be the most stable ice structure for the entire {\sc ssd}
58 gezelter 921 family.
59 chrisfen 743 \end{abstract}
60    
61 chrisfen 862 \newpage
62 chrisfen 743
63     %\narrowtext
64    
65    
66     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
67     % BODY OF TEXT
68     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
69    
70     \section{Introduction}
71    
72 chrisfen 862 One of the most important tasks in the simulation of biochemical
73 gezelter 921 systems is the proper depiction of the aqueous environment of the
74     molecules of interest. In some cases (such as in the simulation of
75     phospholipid bilayers), the majority of the calculations that are
76     performed involve interactions with or between solvent molecules.
77     Thus, the properties one may observe in biochemical simulations are
78     going to be highly dependent on the physical properties of the water
79     model that is chosen.
80 chrisfen 743
81 gezelter 921 There is an especially delicate balance between computational
82     efficiency and the ability of the water model to accurately predict
83     the properties of bulk
84     water.\cite{Jorgensen83,Berendsen87,Jorgensen00} For example, the
85     TIP5P model improves on the structural and transport properties of
86     water relative to the previous TIP models, yet this comes at a greater
87     than 50\% increase in computational
88     cost.\cite{Jorgensen01,Jorgensen00}
89    
90     One recently developed model that largely succeeds in retaining the
91     accuracy of bulk properties while greatly reducing the computational
92 gezelter 1029 cost is the Soft Sticky Dipole ({\sc ssd}) water
93     model.\cite{Ichiye96,Ichiye96b,Ichiye99,Ichiye03} The {\sc ssd} model was
94 gezelter 921 developed by Ichiye \emph{et al.} as a modified form of the
95     hard-sphere water model proposed by Bratko, Blum, and
96 gezelter 1029 Luzar.\cite{Bratko85,Bratko95} {\sc ssd} is a {\it single point} model which
97 gezelter 921 has an interaction site that is both a point dipole along with a
98     Lennard-Jones core. However, since the normal aligned and
99     anti-aligned geometries favored by point dipoles are poor mimics of
100     local structure in liquid water, a short ranged ``sticky'' potential
101     is also added. The sticky potential directs the molecules to assume
102     the proper hydrogen bond orientation in the first solvation
103     shell.
104    
105 gezelter 1029 The interaction between two {\sc ssd} water molecules \emph{i} and \emph{j}
106 gezelter 921 is given by the potential
107 chrisfen 743 \begin{equation}
108     u_{ij} = u_{ij}^{LJ} (r_{ij})\ + u_{ij}^{dp}
109 gezelter 921 ({\bf r}_{ij},{\bf \Omega}_i,{\bf \Omega}_j)\ +
110 chrisfen 743 u_{ij}^{sp}
111 gezelter 921 ({\bf r}_{ij},{\bf \Omega}_i,{\bf \Omega}_j),
112 chrisfen 743 \end{equation}
113 gezelter 921 where the ${\bf r}_{ij}$ is the position vector between molecules
114     \emph{i} and \emph{j} with magnitude $r_{ij}$, and
115     ${\bf \Omega}_i$ and ${\bf \Omega}_j$ represent the orientations of
116     the two molecules. The Lennard-Jones and dipole interactions are given
117     by the following familiar forms:
118 chrisfen 743 \begin{equation}
119 gezelter 921 u_{ij}^{LJ}(r_{ij}) = 4\epsilon
120     \left[\left(\frac{\sigma}{r_{ij}}\right)^{12}-\left(\frac{\sigma}{r_{ij}}\right)^{6}\right]
121     \ ,
122 chrisfen 743 \end{equation}
123 gezelter 921 and
124 chrisfen 743 \begin{equation}
125 gezelter 921 u_{ij}^{dp} = \frac{|\mu_i||\mu_j|}{4 \pi \epsilon_0 r_{ij}^3} \left(
126     \hat{\bf u}_i \cdot \hat{\bf u}_j - 3(\hat{\bf u}_i\cdot\hat{\bf
127     r}_{ij})(\hat{\bf u}_j\cdot\hat{\bf r}_{ij}) \right)\ ,
128 chrisfen 743 \end{equation}
129 gezelter 921 where $\hat{\bf u}_i$ and $\hat{\bf u}_j$ are the unit vectors along
130     the dipoles of molecules $i$ and $j$ respectively. $|\mu_i|$ and
131     $|\mu_j|$ are the strengths of the dipole moments, and $\hat{\bf
132     r}_{ij}$ is the unit vector pointing from molecule $j$ to molecule
133     $i$.
134    
135     The sticky potential is somewhat less familiar:
136 chrisfen 743 \begin{equation}
137     u_{ij}^{sp}
138 gezelter 921 ({\bf r}_{ij},{\bf \Omega}_i,{\bf \Omega}_j) =
139     \frac{\nu_0}{2}[s(r_{ij})w({\bf r}_{ij},{\bf \Omega}_i,{\bf \Omega}_j)
140     + s^\prime(r_{ij})w^\prime({\bf r}_{ij},{\bf \Omega}_i,{\bf
141     \Omega}_j)]\ .
142 chrisfen 1017 \label{stickyfunction}
143 chrisfen 743 \end{equation}
144 gezelter 921 Here, $\nu_0$ is a strength parameter for the sticky potential, and
145     $s$ and $s^\prime$ are cubic switching functions which turn off the
146     sticky interaction beyond the first solvation shell. The $w$ function
147     can be thought of as an attractive potential with tetrahedral
148     geometry:
149 chrisfen 743 \begin{equation}
150 gezelter 921 w({\bf r}_{ij},{\bf \Omega}_i,{\bf \Omega}_j)=\sin\theta_{ij}\sin2\theta_{ij}\cos2\phi_{ij},
151 chrisfen 743 \end{equation}
152 gezelter 921 while the $w^\prime$ function counters the normal aligned and
153     anti-aligned structures favored by point dipoles:
154 chrisfen 743 \begin{equation}
155 chrisfen 1017 w^\prime({\bf r}_{ij},{\bf \Omega}_i,{\bf \Omega}_j) = (\cos\theta_{ij}-0.6)^2(\cos\theta_{ij}+0.8)^2-w^\circ,
156 chrisfen 743 \end{equation}
157 gezelter 921 It should be noted that $w$ is proportional to the sum of the $Y_3^2$
158     and $Y_3^{-2}$ spherical harmonics (a linear combination which
159     enhances the tetrahedral geometry for hydrogen bonded structures),
160     while $w^\prime$ is a purely empirical function. A more detailed
161     description of the functional parts and variables in this potential
162 gezelter 1029 can be found in the original {\sc ssd}
163 gezelter 921 articles.\cite{Ichiye96,Ichiye96b,Ichiye99,Ichiye03}
164 chrisfen 743
165 gezelter 1029 Since {\sc ssd} is a single-point {\it dipolar} model, the force
166 gezelter 921 calculations are simplified significantly relative to the standard
167     {\it charged} multi-point models. In the original Monte Carlo
168     simulations using this model, Ichiye {\it et al.} reported that using
169 gezelter 1029 {\sc ssd} decreased computer time by a factor of 6-7 compared to other
170 gezelter 921 models.\cite{Ichiye96} What is most impressive is that this savings
171     did not come at the expense of accurate depiction of the liquid state
172 gezelter 1029 properties. Indeed, {\sc ssd} maintains reasonable agreement with the Soper
173 gezelter 921 data for the structural features of liquid
174     water.\cite{Soper86,Ichiye96} Additionally, the dynamical properties
175 gezelter 1029 exhibited by {\sc ssd} agree with experiment better than those of more
176 gezelter 921 computationally expensive models (like TIP3P and
177     SPC/E).\cite{Ichiye99} The combination of speed and accurate depiction
178 gezelter 1029 of solvent properties makes {\sc ssd} a very attractive model for the
179 gezelter 921 simulation of large scale biochemical simulations.
180 chrisfen 743
181 gezelter 1029 One feature of the {\sc ssd} model is that it was parameterized for use with
182 gezelter 921 the Ewald sum to handle long-range interactions. This would normally
183     be the best way of handling long-range interactions in systems that
184     contain other point charges. However, our group has recently become
185     interested in systems with point dipoles as mimics for neutral, but
186     polarized regions on molecules (e.g. the zwitterionic head group
187     regions of phospholipids). If the system of interest does not contain
188     point charges, the Ewald sum and even particle-mesh Ewald become
189     computational bottlenecks. Their respective ideal $N^\frac{3}{2}$ and
190     $N\log N$ calculation scaling orders for $N$ particles can become
191     prohibitive when $N$ becomes large.\cite{Darden99} In applying this
192     water model in these types of systems, it would be useful to know its
193     properties and behavior under the more computationally efficient
194     reaction field (RF) technique, or even with a simple cutoff. This
195     study addresses these issues by looking at the structural and
196 gezelter 1029 transport behavior of {\sc ssd} over a variety of temperatures with the
197 gezelter 921 purpose of utilizing the RF correction technique. We then suggest
198     modifications to the parameters that result in more realistic bulk
199     phase behavior. It should be noted that in a recent publication, some
200 gezelter 1029 of the original investigators of the {\sc ssd} water model have suggested
201     adjustments to the {\sc ssd} water model to address abnormal density
202 gezelter 921 behavior (also observed here), calling the corrected model
203 gezelter 1029 {\sc ssd1}.\cite{Ichiye03} In what follows, we compare our
204     reparamaterization of {\sc ssd} with both the original {\sc ssd} and {\sc ssd1} models
205     with the goal of improving the bulk phase behavior of an {\sc ssd}-derived
206 gezelter 921 model in simulations utilizing the Reaction Field.
207 chrisfen 757
208 chrisfen 743 \section{Methods}
209    
210 gezelter 921 Long-range dipole-dipole interactions were accounted for in this study
211     by using either the reaction field method or by resorting to a simple
212 chrisfen 1019 cubic switching function at a cutoff radius. The reaction field
213     method was actually first used in Monte Carlo simulations of liquid
214     water.\cite{Barker73} Under this method, the magnitude of the reaction
215     field acting on dipole $i$ is
216 chrisfen 743 \begin{equation}
217     \mathcal{E}_{i} = \frac{2(\varepsilon_{s} - 1)}{2\varepsilon_{s} + 1}
218 gezelter 1029 \frac{1}{r_{c}^{3}} \sum_{j\in{\mathcal{R}}} {\bf \mu}_{j} s(r_{ij}),
219 chrisfen 743 \label{rfequation}
220     \end{equation}
221     where $\mathcal{R}$ is the cavity defined by the cutoff radius
222     ($r_{c}$), $\varepsilon_{s}$ is the dielectric constant imposed on the
223 gezelter 921 system (80 in the case of liquid water), ${\bf \mu}_{j}$ is the dipole
224 gezelter 1029 moment vector of particle $j$, and $s(r_{ij})$ is a cubic switching
225 chrisfen 743 function.\cite{AllenTildesley} The reaction field contribution to the
226 gezelter 921 total energy by particle $i$ is given by $-\frac{1}{2}{\bf
227     \mu}_{i}\cdot\mathcal{E}_{i}$ and the torque on dipole $i$ by ${\bf
228     \mu}_{i}\times\mathcal{E}_{i}$.\cite{AllenTildesley} Use of the reaction
229     field is known to alter the bulk orientational properties, such as the
230     dielectric relaxation time. There is particular sensitivity of this
231     property on changes in the length of the cutoff
232     radius.\cite{Berendsen98} This variable behavior makes reaction field
233     a less attractive method than the Ewald sum. However, for very large
234     systems, the computational benefit of reaction field is dramatic.
235    
236     We have also performed a companion set of simulations {\it without} a
237     surrounding dielectric (i.e. using a simple cubic switching function
238 chrisfen 1022 at the cutoff radius), and as a result we have two reparamaterizations
239 gezelter 1029 of {\sc ssd} which could be used either with or without the reaction field
240 gezelter 921 turned on.
241 chrisfen 777
242 gezelter 1029 Simulations to obtain the preferred densities of the models were
243     performed in the isobaric-isothermal (NPT) ensemble, while all
244     dynamical properties were obtained from microcanonical (NVE)
245     simulations done at densities matching the NPT density for a
246     particular target temperature. The constant pressure simulations were
247     implemented using an integral thermostat and barostat as outlined by
248     Hoover.\cite{Hoover85,Hoover86} All molecules were treated as
249     non-linear rigid bodies. Vibrational constraints are not necessary in
250     simulations of {\sc ssd}, because there are no explicit hydrogen atoms, and
251     thus no molecular vibrational modes need to be considered.
252 chrisfen 743
253     Integration of the equations of motion was carried out using the
254 chrisfen 1027 symplectic splitting method proposed by Dullweber, Leimkuhler, and
255 gezelter 1029 McLachlan ({\sc dlm}).\cite{Dullweber1997} Our reason for selecting this
256 chrisfen 1027 integrator centers on poor energy conservation of rigid body dynamics
257     using traditional quaternion integration.\cite{Evans77,Evans77b} In
258     typical microcanonical ensemble simulations, the energy drift when
259 gezelter 1029 using quaternions was substantially greater than when using the {\sc dlm}
260 chrisfen 1027 method (fig. \ref{timestep}). This steady drift in the total energy
261     has also been observed by Kol {\it et al.}\cite{Laird97}
262 chrisfen 743
263     The key difference in the integration method proposed by Dullweber
264     \emph{et al.} is that the entire rotation matrix is propagated from
265 gezelter 921 one time step to the next. The additional memory required by the
266     algorithm is inconsequential on modern computers, and translating the
267     rotation matrix into quaternions for storage purposes makes trajectory
268     data quite compact.
269 chrisfen 743
270 gezelter 1029 The {\sc dlm} method allows for Verlet style integration of both
271 chrisfen 1027 translational and orientational motion of rigid bodies. In this
272 gezelter 921 integration method, the orientational propagation involves a sequence
273     of matrix evaluations to update the rotation
274     matrix.\cite{Dullweber1997} These matrix rotations are more costly
275     than the simpler arithmetic quaternion propagation. With the same time
276 gezelter 1029 step, a 1000 {\sc ssd} particle simulation shows an average 7\% increase in
277     computation time using the {\sc dlm} method in place of quaternions. The
278 chrisfen 1027 additional expense per step is justified when one considers the
279 gezelter 1029 ability to use time steps that are nearly twice as large under {\sc dlm}
280 chrisfen 1027 than would be usable under quaternion dynamics. The energy
281     conservation of the two methods using a number of different time steps
282     is illustrated in figure
283 gezelter 921 \ref{timestep}.
284 chrisfen 743
285     \begin{figure}
286 chrisfen 862 \begin{center}
287     \epsfxsize=6in
288     \epsfbox{timeStep.epsi}
289 gezelter 1029 \caption{Energy conservation using both quaternion-based integration and
290     the {\sc dlm} method with increasing time step. The larger time step plots
291     are shifted from the true energy baseline (that of $\Delta t$ = 0.1
292     fs) for clarity.}
293 chrisfen 743 \label{timestep}
294 chrisfen 862 \end{center}
295 chrisfen 743 \end{figure}
296    
297     In figure \ref{timestep}, the resulting energy drift at various time
298 gezelter 1029 steps for both the {\sc dlm} and quaternion integration schemes is compared.
299     All of the 1000 {\sc ssd} particle simulations started with the same
300 chrisfen 1027 configuration, and the only difference was the method used to handle
301     orientational motion. At time steps of 0.1 and 0.5 fs, both methods
302     for propagating the orientational degrees of freedom conserve energy
303     fairly well, with the quaternion method showing a slight energy drift
304     over time in the 0.5 fs time step simulation. At time steps of 1 and 2
305 gezelter 1029 fs, the energy conservation benefits of the {\sc dlm} method are clearly
306 chrisfen 1027 demonstrated. Thus, while maintaining the same degree of energy
307     conservation, one can take considerably longer time steps, leading to
308     an overall reduction in computation time.
309 chrisfen 743
310 gezelter 1029 Energy drift in the simulations using {\sc dlm} integration was unnoticeable
311 chrisfen 1027 for time steps up to 3 fs. A slight energy drift on the order of 0.012
312     kcal/mol per nanosecond was observed at a time step of 4 fs, and as
313     expected, this drift increases dramatically with increasing time
314     step. To insure accuracy in our microcanonical simulations, time steps
315     were set at 2 fs and kept at this value for constant pressure
316     simulations as well.
317 chrisfen 743
318 gezelter 921 Proton-disordered ice crystals in both the $I_h$ and $I_c$ lattices
319     were generated as starting points for all simulations. The $I_h$
320 gezelter 1029 crystals were formed by first arranging the centers of mass of the {\sc ssd}
321 gezelter 921 particles into a ``hexagonal'' ice lattice of 1024 particles. Because
322     of the crystal structure of $I_h$ ice, the simulation box assumed an
323     orthorhombic shape with an edge length ratio of approximately
324 chrisfen 743 1.00$\times$1.06$\times$1.23. The particles were then allowed to
325     orient freely about fixed positions with angular momenta randomized at
326     400 K for varying times. The rotational temperature was then scaled
327 chrisfen 862 down in stages to slowly cool the crystals to 25 K. The particles were
328     then allowed to translate with fixed orientations at a constant
329 chrisfen 743 pressure of 1 atm for 50 ps at 25 K. Finally, all constraints were
330     removed and the ice crystals were allowed to equilibrate for 50 ps at
331     25 K and a constant pressure of 1 atm. This procedure resulted in
332     structurally stable $I_h$ ice crystals that obey the Bernal-Fowler
333 chrisfen 862 rules.\cite{Bernal33,Rahman72} This method was also utilized in the
334 chrisfen 743 making of diamond lattice $I_c$ ice crystals, with each cubic
335     simulation box consisting of either 512 or 1000 particles. Only
336     isotropic volume fluctuations were performed under constant pressure,
337     so the ratio of edge lengths remained constant throughout the
338     simulations.
339    
340     \section{Results and discussion}
341    
342     Melting studies were performed on the randomized ice crystals using
343 gezelter 921 isobaric-isothermal (NPT) dynamics. During melting simulations, the
344     melting transition and the density maximum can both be observed,
345     provided that the density maximum occurs in the liquid and not the
346     supercooled regime. An ensemble average from five separate melting
347     simulations was acquired, each starting from different ice crystals
348     generated as described previously. All simulations were equilibrated
349     for 100 ps prior to a 200 ps data collection run at each temperature
350     setting. The temperature range of study spanned from 25 to 400 K, with
351     a maximum degree increment of 25 K. For regions of interest along this
352     stepwise progression, the temperature increment was decreased from 25
353     K to 10 and 5 K. The above equilibration and production times were
354     sufficient in that fluctuations in the volume autocorrelation function
355     were damped out in all simulations in under 20 ps.
356 chrisfen 743
357     \subsection{Density Behavior}
358    
359 gezelter 1029 Our initial simulations focused on the original {\sc ssd} water model, and
360 gezelter 921 an average density versus temperature plot is shown in figure
361     \ref{dense1}. Note that the density maximum when using a reaction
362     field appears between 255 and 265 K. There were smaller fluctuations
363     in the density at 260 K than at either 255 or 265, so we report this
364     value as the location of the density maximum. Figure \ref{dense1} was
365     constructed using ice $I_h$ crystals for the initial configuration;
366     though not pictured, the simulations starting from ice $I_c$ crystal
367     configurations showed similar results, with a liquid-phase density
368     maximum in this same region (between 255 and 260 K).
369    
370 chrisfen 743 \begin{figure}
371 chrisfen 862 \begin{center}
372     \epsfxsize=6in
373     \epsfbox{denseSSD.eps}
374 gezelter 921 \caption{Density versus temperature for TIP4P [Ref. \citen{Jorgensen98b}],
375 gezelter 1029 TIP3P [Ref. \citen{Jorgensen98b}], SPC/E [Ref. \citen{Clancy94}], {\sc ssd}
376     without Reaction Field, {\sc ssd}, and experiment [Ref. \citen{CRC80}]. The
377 gezelter 921 arrows indicate the change in densities observed when turning off the
378 gezelter 1029 reaction field. The the lower than expected densities for the {\sc ssd}
379     model were what prompted the original reparameterization of {\sc ssd1}
380 gezelter 921 [Ref. \citen{Ichiye03}].}
381 chrisfen 861 \label{dense1}
382 chrisfen 862 \end{center}
383 chrisfen 743 \end{figure}
384    
385 gezelter 1029 The density maximum for {\sc ssd} compares quite favorably to other simple
386 gezelter 921 water models. Figure \ref{dense1} also shows calculated densities of
387     several other models and experiment obtained from other
388 chrisfen 743 sources.\cite{Jorgensen98b,Clancy94,CRC80} Of the listed simple water
389 gezelter 1029 models, {\sc ssd} has a temperature closest to the experimentally observed
390 gezelter 921 density maximum. Of the {\it charge-based} models in
391     Fig. \ref{dense1}, TIP4P has a density maximum behavior most like that
392 gezelter 1029 seen in {\sc ssd}. Though not included in this plot, it is useful
393 gezelter 921 to note that TIP5P has a density maximum nearly identical to the
394     experimentally measured temperature.
395 chrisfen 743
396 gezelter 921 It has been observed that liquid state densities in water are
397     dependent on the cutoff radius used both with and without the use of
398     reaction field.\cite{Berendsen98} In order to address the possible
399     effect of cutoff radius, simulations were performed with a dipolar
400 gezelter 1029 cutoff radius of 12.0 \AA\ to complement the previous {\sc ssd} simulations,
401 gezelter 921 all performed with a cutoff of 9.0 \AA. All of the resulting densities
402     overlapped within error and showed no significant trend toward lower
403     or higher densities as a function of cutoff radius, for simulations
404     both with and without reaction field. These results indicate that
405     there is no major benefit in choosing a longer cutoff radius in
406 gezelter 1029 simulations using {\sc ssd}. This is advantageous in that the use of a
407 gezelter 921 longer cutoff radius results in a significant increase in the time
408     required to obtain a single trajectory.
409 chrisfen 743
410 chrisfen 862 The key feature to recognize in figure \ref{dense1} is the density
411 gezelter 1029 scaling of {\sc ssd} relative to other common models at any given
412     temperature. {\sc ssd} assumes a lower density than any of the other listed
413 gezelter 921 models at the same pressure, behavior which is especially apparent at
414     temperatures greater than 300 K. Lower than expected densities have
415     been observed for other systems using a reaction field for long-range
416     electrostatic interactions, so the most likely reason for the
417     significantly lower densities seen in these simulations is the
418     presence of the reaction field.\cite{Berendsen98,Nezbeda02} In order
419     to test the effect of the reaction field on the density of the
420     systems, the simulations were repeated without a reaction field
421     present. The results of these simulations are also displayed in figure
422     \ref{dense1}. Without the reaction field, the densities increase
423     to more experimentally reasonable values, especially around the
424     freezing point of liquid water. The shape of the curve is similar to
425 gezelter 1029 the curve produced from {\sc ssd} simulations using reaction field,
426 gezelter 921 specifically the rapidly decreasing densities at higher temperatures;
427     however, a shift in the density maximum location, down to 245 K, is
428     observed. This is a more accurate comparison to the other listed water
429     models, in that no long range corrections were applied in those
430     simulations.\cite{Clancy94,Jorgensen98b} However, even without the
431 chrisfen 861 reaction field, the density around 300 K is still significantly lower
432     than experiment and comparable water models. This anomalous behavior
433 chrisfen 1027 was what lead Tan {\it et al.} to recently reparameterize
434 gezelter 1029 {\sc ssd}.\cite{Ichiye03} Throughout the remainder of the paper our
435     reparamaterizations of {\sc ssd} will be compared with their newer {\sc ssd1}
436     model.
437 chrisfen 861
438 chrisfen 743 \subsection{Transport Behavior}
439    
440 gezelter 921 Accurate dynamical properties of a water model are particularly
441     important when using the model to study permeation or transport across
442     biological membranes. In order to probe transport in bulk water,
443     constant energy (NVE) simulations were performed at the average
444     density obtained by the NPT simulations at an identical target
445     temperature. Simulations started with randomized velocities and
446     underwent 50 ps of temperature scaling and 50 ps of constant energy
447     equilibration before a 200 ps data collection run. Diffusion constants
448     were calculated via linear fits to the long-time behavior of the
449     mean-square displacement as a function of time. The averaged results
450     from five sets of NVE simulations are displayed in figure
451     \ref{diffuse}, alongside experimental, SPC/E, and TIP5P
452 chrisfen 1022 results.\cite{Gillen72,Holz00,Clancy94,Jorgensen01}
453 gezelter 921
454 chrisfen 743 \begin{figure}
455 chrisfen 862 \begin{center}
456     \epsfxsize=6in
457     \epsfbox{betterDiffuse.epsi}
458 gezelter 921 \caption{Average self-diffusion constant as a function of temperature for
459 gezelter 1029 {\sc ssd}, SPC/E [Ref. \citen{Clancy94}], and TIP5P
460     [Ref. \citen{Jorgensen01}] compared with experimental data
461     [Refs. \citen{Gillen72} and \citen{Holz00}]. Of the three water models
462     shown, {\sc ssd} has the least deviation from the experimental values. The
463     rapidly increasing diffusion constants for TIP5P and {\sc ssd} correspond to
464     significant decreases in density at the higher temperatures.}
465 chrisfen 743 \label{diffuse}
466 chrisfen 862 \end{center}
467 chrisfen 743 \end{figure}
468    
469     The observed values for the diffusion constant point out one of the
470 gezelter 1029 strengths of the {\sc ssd} model. Of the three models shown, the {\sc ssd} model
471 gezelter 921 has the most accurate depiction of self-diffusion in both the
472     supercooled and liquid regimes. SPC/E does a respectable job by
473     reproducing values similar to experiment around 290 K; however, it
474     deviates at both higher and lower temperatures, failing to predict the
475 gezelter 1029 correct thermal trend. TIP5P and {\sc ssd} both start off low at colder
476 gezelter 921 temperatures and tend to diffuse too rapidly at higher temperatures.
477     This behavior at higher temperatures is not particularly surprising
478 gezelter 1029 since the densities of both TIP5P and {\sc ssd} are lower than experimental
479 gezelter 921 water densities at higher temperatures. When calculating the
480 gezelter 1029 diffusion coefficients for {\sc ssd} at experimental densities (instead of
481 gezelter 921 the densities from the NPT simulations), the resulting values fall
482     more in line with experiment at these temperatures.
483 chrisfen 743
484     \subsection{Structural Changes and Characterization}
485 gezelter 921
486 chrisfen 743 By starting the simulations from the crystalline state, the melting
487 gezelter 921 transition and the ice structure can be obtained along with the liquid
488 chrisfen 862 phase behavior beyond the melting point. The constant pressure heat
489     capacity (C$_\text{p}$) was monitored to locate the melting transition
490     in each of the simulations. In the melting simulations of the 1024
491     particle ice $I_h$ simulations, a large spike in C$_\text{p}$ occurs
492     at 245 K, indicating a first order phase transition for the melting of
493     these ice crystals. When the reaction field is turned off, the melting
494     transition occurs at 235 K. These melting transitions are
495 gezelter 921 considerably lower than the experimental value.
496 chrisfen 743
497 chrisfen 862 \begin{figure}
498     \begin{center}
499     \epsfxsize=6in
500     \epsfbox{corrDiag.eps}
501 gezelter 1029 \caption{An illustration of angles involved in the correlations observed in Fig. \ref{contour}.}
502 chrisfen 862 \label{corrAngle}
503     \end{center}
504     \end{figure}
505    
506     \begin{figure}
507     \begin{center}
508     \epsfxsize=6in
509     \epsfbox{fullContours.eps}
510 gezelter 1029 \caption{Contour plots of 2D angular pair correlation functions for
511     512 {\sc ssd} molecules at 100 K (A \& B) and 300 K (C \& D). Dark areas
512     signify regions of enhanced density while light areas signify
513     depletion relative to the bulk density. White areas have pair
514     correlation values below 0.5 and black areas have values above 1.5.}
515 chrisfen 743 \label{contour}
516 chrisfen 862 \end{center}
517 chrisfen 743 \end{figure}
518    
519 gezelter 921 Additional analysis of the melting process was performed using
520     two-dimensional structure and dipole angle correlations. Expressions
521     for these correlations are as follows:
522 chrisfen 861
523 chrisfen 862 \begin{equation}
524 gezelter 921 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|{\bf r}_{ij}\right|)\rangle\ ,
525 chrisfen 862 \end{equation}
526     \begin{equation}
527     g_{\text{AB}}(r,\cos\omega) =
528 gezelter 921 \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|{\bf r}_{ij}\right|)\rangle\ ,
529 chrisfen 862 \end{equation}
530 chrisfen 861 where $\theta$ and $\omega$ refer to the angles shown in figure
531     \ref{corrAngle}. By binning over both distance and the cosine of the
532 gezelter 921 desired angle between the two dipoles, the $g(r)$ can be analyzed to
533     determine the common dipole arrangements that constitute the peaks and
534     troughs in the standard one-dimensional $g(r)$ plots. Frames A and B
535     of figure \ref{contour} show results from an ice $I_c$ simulation. The
536     first peak in the $g(r)$ consists primarily of the preferred hydrogen
537 chrisfen 861 bonding arrangements as dictated by the tetrahedral sticky potential -
538 gezelter 921 one peak for the hydrogen bond donor and the other for the hydrogen
539     bond acceptor. Due to the high degree of crystallinity of the sample,
540     the second and third solvation shells show a repeated peak arrangement
541 chrisfen 743 which decays at distances around the fourth solvation shell, near the
542     imposed cutoff for the Lennard-Jones and dipole-dipole interactions.
543 chrisfen 861 In the higher temperature simulation shown in frames C and D, these
544 gezelter 921 long-range features deteriorate rapidly. The first solvation shell
545     still shows the strong effect of the sticky-potential, although it
546     covers a larger area, extending to include a fraction of aligned
547     dipole peaks within the first solvation shell. The latter peaks lose
548     due to thermal motion and as the competing dipole force overcomes the
549     sticky potential's tight tetrahedral structuring of the crystal.
550 chrisfen 743
551     This complex interplay between dipole and sticky interactions was
552     remarked upon as a possible reason for the split second peak in the
553 gezelter 1029 oxygen-oxygen pair correlation function,
554     $g_\mathrm{OO}(r)$.\cite{Ichiye96} At low temperatures, the second
555     solvation shell peak appears to have two distinct components that
556     blend together to form one observable peak. At higher temperatures,
557     this split character alters to show the leading 4 \AA\ peak dominated
558     by equatorial anti-parallel dipole orientations. There is also a
559     tightly bunched group of axially arranged dipoles that most likely
560     consist of the smaller fraction of aligned dipole pairs. The trailing
561     component of the split peak at 5 \AA\ is dominated by aligned dipoles
562     that assume hydrogen bond arrangements similar to those seen in the
563     first solvation shell. This evidence indicates that the dipole pair
564     interaction begins to dominate outside of the range of the dipolar
565     repulsion term. The energetically favorable dipole arrangements
566     populate the region immediately outside this repulsion region (around
567     4 \AA), while arrangements that seek to satisfy both the sticky and
568     dipole forces locate themselves just beyond this initial buildup
569     (around 5 \AA).
570 chrisfen 743
571     From these findings, the split second peak is primarily the product of
572 chrisfen 861 the dipolar repulsion term of the sticky potential. In fact, the inner
573     peak can be pushed out and merged with the outer split peak just by
574 gezelter 921 extending the switching function ($s^\prime(r_{ij})$) from its normal
575     4.0 \AA\ cutoff to values of 4.5 or even 5 \AA. This type of
576 chrisfen 861 correction is not recommended for improving the liquid structure,
577 chrisfen 862 since the second solvation shell would still be shifted too far
578 chrisfen 861 out. In addition, this would have an even more detrimental effect on
579     the system densities, leading to a liquid with a more open structure
580 gezelter 1029 and a density considerably lower than the already low {\sc ssd} density. A
581 gezelter 921 better correction would be to include the quadrupole-quadrupole
582     interactions for the water particles outside of the first solvation
583     shell, but this would remove the simplicity and speed advantage of
584 gezelter 1029 {\sc ssd}.
585 chrisfen 743
586 gezelter 1029 \subsection{Adjusted Potentials: {\sc ssd/rf} and {\sc ssd/e}}
587 gezelter 921
588 gezelter 1029 The propensity of {\sc ssd} to adopt lower than expected densities under
589 chrisfen 743 varying conditions is troubling, especially at higher temperatures. In
590 chrisfen 861 order to correct this model for use with a reaction field, it is
591     necessary to adjust the force field parameters for the primary
592     intermolecular interactions. In undergoing a reparameterization, it is
593     important not to focus on just one property and neglect the other
594     important properties. In this case, it would be ideal to correct the
595 gezelter 921 densities while maintaining the accurate transport behavior.
596 chrisfen 743
597 chrisfen 1017 The parameters available for tuning include the $\sigma$ and
598     $\epsilon$ Lennard-Jones parameters, the dipole strength ($\mu$), the
599 gezelter 1029 strength of the sticky potential ($\nu_0$), and the cutoff distances
600     for the sticky attractive and dipole repulsive cubic switching
601     function cutoffs ($r_l$, $r_u$ and $r_l^\prime$, $r_u^\prime$
602     respectively). The results of the reparameterizations are shown in
603     table \ref{params}. We are calling these reparameterizations the Soft
604     Sticky Dipole / Reaction Field ({\sc ssd/rf} - for use with a reaction
605     field) and Soft Sticky Dipole Extended ({\sc ssd/e} - an attempt to improve
606     the liquid structure in simulations without a long-range correction).
607 chrisfen 743
608     \begin{table}
609 chrisfen 862 \begin{center}
610 chrisfen 743 \caption{Parameters for the original and adjusted models}
611 chrisfen 856 \begin{tabular}{ l c c c c }
612 chrisfen 743 \hline \\[-3mm]
613 gezelter 1029 \ \ \ Parameters\ \ \ & \ \ \ {\sc ssd} [Ref. \citen{Ichiye96}] \ \ \
614     & \ {\sc ssd1} [Ref. \citen{Ichiye03}]\ \ & \ {\sc ssd/e}\ \ & \ {\sc ssd/rf} \\
615 chrisfen 743 \hline \\[-3mm]
616 chrisfen 856 \ \ \ $\sigma$ (\AA) & 3.051 & 3.016 & 3.035 & 3.019\\
617     \ \ \ $\epsilon$ (kcal/mol) & 0.152 & 0.152 & 0.152 & 0.152\\
618     \ \ \ $\mu$ (D) & 2.35 & 2.35 & 2.42 & 2.48\\
619     \ \ \ $\nu_0$ (kcal/mol) & 3.7284 & 3.6613 & 3.90 & 3.90\\
620 chrisfen 1017 \ \ \ $\omega^\circ$ & 0.07715 & 0.07715 & 0.07715 & 0.07715\\
621 chrisfen 856 \ \ \ $r_l$ (\AA) & 2.75 & 2.75 & 2.40 & 2.40\\
622     \ \ \ $r_u$ (\AA) & 3.35 & 3.35 & 3.80 & 3.80\\
623     \ \ \ $r_l^\prime$ (\AA) & 2.75 & 2.75 & 2.75 & 2.75\\
624     \ \ \ $r_u^\prime$ (\AA) & 4.00 & 4.00 & 3.35 & 3.35\\
625 chrisfen 743 \end{tabular}
626     \label{params}
627 chrisfen 862 \end{center}
628 chrisfen 743 \end{table}
629    
630 chrisfen 862 \begin{figure}
631     \begin{center}
632     \epsfxsize=5in
633     \epsfbox{GofRCompare.epsi}
634 gezelter 1029 \caption{Plots comparing experiment [Ref. \citen{Head-Gordon00_1}] with {\sc ssd/e}
635     and {\sc ssd1} without reaction field (top), as well as {\sc ssd/rf} and {\sc ssd1} with
636 chrisfen 743 reaction field turned on (bottom). The insets show the respective
637 chrisfen 862 first peaks in detail. Note how the changes in parameters have lowered
638 gezelter 1029 and broadened the first peak of {\sc ssd/e} and {\sc ssd/rf}.}
639 chrisfen 743 \label{grcompare}
640 chrisfen 862 \end{center}
641 chrisfen 743 \end{figure}
642    
643 chrisfen 862 \begin{figure}
644     \begin{center}
645     \epsfxsize=6in
646 chrisfen 1027 \epsfbox{dualsticky_bw.eps}
647 gezelter 1029 \caption{Positive and negative isosurfaces of the sticky potential for
648     {\sc ssd1} (left) and {\sc ssd/e} \& {\sc ssd/rf} (right). Light areas correspond to the
649     tetrahedral attractive component, and darker areas correspond to the
650     dipolar repulsive component.}
651 chrisfen 743 \label{isosurface}
652 chrisfen 862 \end{center}
653 chrisfen 743 \end{figure}
654    
655 gezelter 1029 In the original paper detailing the development of {\sc ssd}, Liu and Ichiye
656 gezelter 921 placed particular emphasis on an accurate description of the first
657     solvation shell. This resulted in a somewhat tall and narrow first
658     peak in $g(r)$ that integrated to give similar coordination numbers to
659 chrisfen 862 the experimental data obtained by Soper and
660     Phillips.\cite{Ichiye96,Soper86} New experimental x-ray scattering
661     data from the Head-Gordon lab indicates a slightly lower and shifted
662 gezelter 1029 first peak in the g$_\mathrm{OO}(r)$, so our adjustments to {\sc ssd} were
663     made after taking into consideration the new experimental
664 chrisfen 862 findings.\cite{Head-Gordon00_1} Figure \ref{grcompare} shows the
665 gezelter 921 relocation of the first peak of the oxygen-oxygen $g(r)$ by comparing
666 gezelter 1029 the revised {\sc ssd} model ({\sc ssd1}), {\sc ssd/e}, and {\sc ssd/rf} to the new
667 chrisfen 862 experimental results. Both modified water models have shorter peaks
668 gezelter 921 that match more closely to the experimental peak (as seen in the
669     insets of figure \ref{grcompare}). This structural alteration was
670 chrisfen 862 accomplished by the combined reduction in the Lennard-Jones $\sigma$
671 gezelter 921 variable and adjustment of the sticky potential strength and cutoffs.
672     As can be seen in table \ref{params}, the cutoffs for the tetrahedral
673     attractive and dipolar repulsive terms were nearly swapped with each
674     other. Isosurfaces of the original and modified sticky potentials are
675     shown in figure \ref{isosurface}. In these isosurfaces, it is easy to
676     see how altering the cutoffs changes the repulsive and attractive
677     character of the particles. With a reduced repulsive surface (darker
678     region), the particles can move closer to one another, increasing the
679     density for the overall system. This change in interaction cutoff also
680     results in a more gradual orientational motion by allowing the
681     particles to maintain preferred dipolar arrangements before they begin
682     to feel the pull of the tetrahedral restructuring. As the particles
683     move closer together, the dipolar repulsion term becomes active and
684     excludes unphysical nearest-neighbor arrangements. This compares with
685 gezelter 1029 how {\sc ssd} and {\sc ssd1} exclude preferred dipole alignments before the
686 gezelter 921 particles feel the pull of the ``hydrogen bonds''. Aside from
687     improving the shape of the first peak in the g(\emph{r}), this
688     modification improves the densities considerably by allowing the
689     persistence of full dipolar character below the previous 4.0 \AA\
690     cutoff.
691 chrisfen 743
692 gezelter 921 While adjusting the location and shape of the first peak of $g(r)$
693     improves the densities, these changes alone are insufficient to bring
694     the system densities up to the values observed experimentally. To
695     further increase the densities, the dipole moments were increased in
696 gezelter 1029 both of our adjusted models. Since {\sc ssd} is a dipole based model, the
697 gezelter 921 structure and transport are very sensitive to changes in the dipole
698 gezelter 1029 moment. The original {\sc ssd} simply used the dipole moment calculated from
699 gezelter 921 the TIP3P water model, which at 2.35 D is significantly greater than
700     the experimental gas phase value of 1.84 D. The larger dipole moment
701     is a more realistic value and improves the dielectric properties of
702     the fluid. Both theoretical and experimental measurements indicate a
703     liquid phase dipole moment ranging from 2.4 D to values as high as
704     3.11 D, providing a substantial range of reasonable values for a
705     dipole moment.\cite{Sprik91,Kusalik02,Badyal00,Barriol64} Moderately
706 gezelter 1029 increasing the dipole moments to 2.42 and 2.48 D for {\sc ssd/e} and {\sc ssd/rf},
707 chrisfen 862 respectively, leads to significant changes in the density and
708     transport of the water models.
709 chrisfen 743
710 chrisfen 861 In order to demonstrate the benefits of these reparameterizations, a
711 chrisfen 743 series of NPT and NVE simulations were performed to probe the density
712     and transport properties of the adapted models and compare the results
713 gezelter 1029 to the original {\sc ssd} model. This comparison involved full NPT melting
714     sequences for both {\sc ssd/e} and {\sc ssd/rf}, as well as NVE transport
715 chrisfen 861 calculations at the calculated self-consistent densities. Again, the
716 chrisfen 862 results are obtained from five separate simulations of 1024 particle
717     systems, and the melting sequences were started from different ice
718     $I_h$ crystals constructed as described previously. Each NPT
719 chrisfen 861 simulation was equilibrated for 100 ps before a 200 ps data collection
720 chrisfen 862 run at each temperature step, and the final configuration from the
721     previous temperature simulation was used as a starting point. All NVE
722     simulations had the same thermalization, equilibration, and data
723 gezelter 921 collection times as stated previously.
724 chrisfen 743
725 chrisfen 862 \begin{figure}
726     \begin{center}
727     \epsfxsize=6in
728     \epsfbox{ssdeDense.epsi}
729 gezelter 1029 \caption{Comparison of densities calculated with {\sc ssd/e} to {\sc ssd1} without a
730 gezelter 921 reaction field, TIP3P [Ref. \citen{Jorgensen98b}], TIP5P
731     [Ref. \citen{Jorgensen00}], SPC/E [Ref. \citen{Clancy94}] and
732     experiment [Ref. \citen{CRC80}]. The window shows a expansion around
733     300 K with error bars included to clarify this region of
734 gezelter 1029 interest. Note that both {\sc ssd1} and {\sc ssd/e} show good agreement with
735 chrisfen 856 experiment when the long-range correction is neglected.}
736 chrisfen 743 \label{ssdedense}
737 chrisfen 862 \end{center}
738 chrisfen 743 \end{figure}
739    
740 gezelter 1029 Fig. \ref{ssdedense} shows the density profile for the {\sc ssd/e} model
741     in comparison to {\sc ssd1} without a reaction field, other common water
742 chrisfen 862 models, and experimental results. The calculated densities for both
743 gezelter 1029 {\sc ssd/e} and {\sc ssd1} have increased significantly over the original {\sc ssd}
744 gezelter 921 model (see fig. \ref{dense1}) and are in better agreement with the
745 gezelter 1029 experimental values. At 298 K, the densities of {\sc ssd/e} and {\sc ssd1} without
746 chrisfen 862 a long-range correction are 0.996$\pm$0.001 g/cm$^3$ and
747     0.999$\pm$0.001 g/cm$^3$ respectively. These both compare well with
748     the experimental value of 0.997 g/cm$^3$, and they are considerably
749 gezelter 1029 better than the {\sc ssd} value of 0.967$\pm$0.003 g/cm$^3$. The changes to
750 chrisfen 862 the dipole moment and sticky switching functions have improved the
751     structuring of the liquid (as seen in figure \ref{grcompare}, but they
752     have shifted the density maximum to much lower temperatures. This
753     comes about via an increase in the liquid disorder through the
754     weakening of the sticky potential and strengthening of the dipolar
755 gezelter 1029 character. However, this increasing disorder in the {\sc ssd/e} model has
756 gezelter 921 little effect on the melting transition. By monitoring $C_p$
757 gezelter 1029 throughout these simulations, the melting transition for {\sc ssd/e} was
758 gezelter 921 shown to occur at 235 K. The same transition temperature observed
759 gezelter 1029 with {\sc ssd} and {\sc ssd1}.
760 chrisfen 743
761 chrisfen 862 \begin{figure}
762     \begin{center}
763     \epsfxsize=6in
764     \epsfbox{ssdrfDense.epsi}
765 gezelter 1029 \caption{Comparison of densities calculated with {\sc ssd/rf} to {\sc ssd1} with a
766 gezelter 921 reaction field, TIP3P [Ref. \citen{Jorgensen98b}], TIP5P
767     [Ref. \citen{Jorgensen00}], SPC/E [Ref. \citen{Clancy94}], and
768     experiment [Ref. \citen{CRC80}]. The inset shows the necessity of
769     reparameterization when utilizing a reaction field long-ranged
770 gezelter 1029 correction - {\sc ssd/rf} provides significantly more accurate densities
771     than {\sc ssd1} when performing room temperature simulations.}
772 chrisfen 743 \label{ssdrfdense}
773 chrisfen 862 \end{center}
774 chrisfen 743 \end{figure}
775    
776 chrisfen 862 Including the reaction field long-range correction in the simulations
777 gezelter 921 results in a more interesting comparison. A density profile including
778 gezelter 1029 {\sc ssd/rf} and {\sc ssd1} with an active reaction field is shown in figure
779 chrisfen 862 \ref{ssdrfdense}. As observed in the simulations without a reaction
780 gezelter 1029 field, the densities of {\sc ssd/rf} and {\sc ssd1} show a dramatic increase over
781     normal {\sc ssd} (see figure \ref{dense1}). At 298 K, {\sc ssd/rf} has a density
782 chrisfen 862 of 0.997$\pm$0.001 g/cm$^3$, directly in line with experiment and
783 gezelter 1029 considerably better than the original {\sc ssd} value of 0.941$\pm$0.001
784     g/cm$^3$ and the {\sc ssd1} value of 0.972$\pm$0.002 g/cm$^3$. These results
785 gezelter 921 further emphasize the importance of reparameterization in order to
786     model the density properly under different simulation conditions.
787     Again, these changes have only a minor effect on the melting point,
788 gezelter 1029 which observed at 245 K for {\sc ssd/rf}, is identical to {\sc ssd} and only 5 K
789     lower than {\sc ssd1} with a reaction field. Additionally, the difference in
790     density maxima is not as extreme, with {\sc ssd/rf} showing a density
791 gezelter 921 maximum at 255 K, fairly close to the density maxima of 260 K and 265
792 gezelter 1029 K, shown by {\sc ssd} and {\sc ssd1} respectively.
793 chrisfen 743
794 chrisfen 862 \begin{figure}
795     \begin{center}
796     \epsfxsize=6in
797     \epsfbox{ssdeDiffuse.epsi}
798 gezelter 1029 \caption{The diffusion constants calculated from {\sc ssd/e} and {\sc ssd1} (both
799     without a reaction field) along with experimental results
800     [Refs. \citen{Gillen72} and \citen{Holz00}]. The NVE calculations were
801     performed at the average densities observed in the 1 atm NPT
802     simulations for the respective models. {\sc ssd/e} is slightly more mobile
803     than experiment at all of the temperatures, but it is closer to
804     experiment at biologically relevant temperatures than {\sc ssd1} without a
805     long-range correction.}
806 chrisfen 861 \label{ssdediffuse}
807 chrisfen 862 \end{center}
808 chrisfen 861 \end{figure}
809    
810 gezelter 1029 The reparameterization of the {\sc ssd} water model, both for use with and
811 chrisfen 743 without an applied long-range correction, brought the densities up to
812     what is expected for simulating liquid water. In addition to improving
813 gezelter 1029 the densities, it is important that the diffusive behavior of {\sc ssd} be
814     maintained or improved. Figure \ref{ssdediffuse} compares the
815     temperature dependence of the diffusion constant of {\sc ssd/e} to {\sc ssd1}
816 chrisfen 1027 without an active reaction field at the densities calculated from
817     their respective NPT simulations at 1 atm. The diffusion constant for
818 gezelter 1029 {\sc ssd/e} is consistently higher than experiment, while {\sc ssd1} remains lower
819 chrisfen 1027 than experiment until relatively high temperatures (around 360
820     K). Both models follow the shape of the experimental curve well below
821     300 K but tend to diffuse too rapidly at higher temperatures, as seen
822 gezelter 1029 in {\sc ssd1}'s crossing above 360 K. This increasing diffusion relative to
823 chrisfen 1027 the experimental values is caused by the rapidly decreasing system
824 gezelter 1029 density with increasing temperature. Both {\sc ssd1} and {\sc ssd/e} show this
825 chrisfen 1027 deviation in particle mobility, but this trend has different
826 gezelter 1029 implications on the diffusive behavior of the models. While {\sc ssd1}
827 chrisfen 1027 shows more experimentally accurate diffusive behavior in the high
828 gezelter 1029 temperature regimes, {\sc ssd/e} shows more accurate behavior in the
829 chrisfen 1027 supercooled and biologically relevant temperature ranges. Thus, the
830     changes made to improve the liquid structure may have had an adverse
831     affect on the density maximum, but they improve the transport behavior
832 gezelter 1029 of {\sc ssd/e} relative to {\sc ssd1} under the most commonly simulated
833 chrisfen 1027 conditions.
834 chrisfen 743
835 chrisfen 862 \begin{figure}
836     \begin{center}
837     \epsfxsize=6in
838     \epsfbox{ssdrfDiffuse.epsi}
839 gezelter 1029 \caption{The diffusion constants calculated from {\sc ssd/rf} and {\sc ssd1} (both
840     with an active reaction field) along with experimental results
841     [Refs. \citen{Gillen72} and \citen{Holz00}]. The NVE calculations were
842     performed at the average densities observed in the 1 atm NPT
843     simulations for both of the models. {\sc ssd/rf} simulates the diffusion of
844     water throughout this temperature range very well. The rapidly
845     increasing diffusion constants at high temperatures for both models
846     can be attributed to lower calculated densities than those observed in
847     experiment.}
848 chrisfen 856 \label{ssdrfdiffuse}
849 chrisfen 862 \end{center}
850 chrisfen 743 \end{figure}
851    
852 gezelter 1029 In figure \ref{ssdrfdiffuse}, the diffusion constants for {\sc ssd/rf} are
853     compared to {\sc ssd1} with an active reaction field. Note that {\sc ssd/rf}
854 gezelter 921 tracks the experimental results quantitatively, identical within error
855 chrisfen 1017 throughout most of the temperature range shown and exhibiting only a
856 gezelter 1029 slight increasing trend at higher temperatures. {\sc ssd1} tends to diffuse
857 chrisfen 1017 more slowly at low temperatures and deviates to diffuse too rapidly at
858 gezelter 921 temperatures greater than 330 K. As stated above, this deviation away
859     from the ideal trend is due to a rapid decrease in density at higher
860 gezelter 1029 temperatures. {\sc ssd/rf} does not suffer from this problem as much as {\sc ssd1}
861 gezelter 921 because the calculated densities are closer to the experimental
862     values. These results again emphasize the importance of careful
863     reparameterization when using an altered long-range correction.
864 chrisfen 743
865 chrisfen 1017 \begin{table}
866 gezelter 1029 \begin{minipage}{\linewidth}
867     \renewcommand{\thefootnote}{\thempfootnote}
868 chrisfen 1017 \begin{center}
869 gezelter 1029 \caption{Properties of the single-point water models compared with
870     experimental data at ambient conditions}
871 chrisfen 1017 \begin{tabular}{ l c c c c c }
872     \hline \\[-3mm]
873 gezelter 1029 \ \ \ \ \ \ & \ \ \ {\sc ssd1} \ \ \ & \ {\sc ssd/e} \ \ \ & \ {\sc ssd1} (RF) \ \
874     \ & \ {\sc ssd/rf} \ \ \ & \ Expt. \\
875 chrisfen 1017 \hline \\[-3mm]
876     \ \ \ $\rho$ (g/cm$^3$) & 0.999 $\pm$0.001 & 0.996 $\pm$0.001 & 0.972 $\pm$0.002 & 0.997 $\pm$0.001 & 0.997 \\
877     \ \ \ $C_p$ (cal/mol K) & 28.80 $\pm$0.11 & 25.45 $\pm$0.09 & 28.28 $\pm$0.06 & 23.83 $\pm$0.16 & 17.98 \\
878 gezelter 1029 \ \ \ $D$ ($10^{-5}$ cm$^2$/s) & 1.78 $\pm$0.07 & 2.51 $\pm$0.18 &
879     2.00 $\pm$0.17 & 2.32 $\pm$0.06 & 2.299\cite{Mills73} \\
880     \ \ \ Coordination Number ($n_C$) & 3.9 & 4.3 & 3.8 & 4.4 &
881     4.7\footnote{Calculated by integrating $g_{\text{OO}}(r)$ in
882     Ref. \citen{Head-Gordon00_1}} \\
883     \ \ \ H-bonds per particle ($n_H$) & 3.7 & 3.6 & 3.7 & 3.7 &
884     3.5\footnote{Calculated by integrating $g_{\text{OH}}(r)$ in
885     Ref. \citen{Soper86}} \\
886     \ \ \ $\tau_1$ (ps) & 10.9 $\pm$0.6 & 7.3 $\pm$0.4 & 7.5 $\pm$0.7 &
887     7.2 $\pm$0.4 & 5.7\footnote{Calculated for 298 K from data in Ref. \citen{Eisenberg69}} \\
888     \ \ \ $\tau_2$ (ps) & 4.7 $\pm$0.4 & 3.1 $\pm$0.2 & 3.5 $\pm$0.3 & 3.2
889     $\pm$0.2 & 2.3\footnote{Calculated for 298 K from data in
890     Ref. \citen{Krynicki66}}
891 chrisfen 1017 \end{tabular}
892     \label{liquidproperties}
893     \end{center}
894 gezelter 1029 \end{minipage}
895 chrisfen 1017 \end{table}
896    
897     Table \ref{liquidproperties} gives a synopsis of the liquid state
898     properties of the water models compared in this study along with the
899     experimental values for liquid water at ambient conditions. The
900 gezelter 1029 coordination number ($n_C$) and number of hydrogen bonds per particle
901     ($n_H$) were calculated by integrating the following relations:
902 chrisfen 1017 \begin{equation}
903 gezelter 1029 n_C = 4\pi\rho_{\text{OO}}\int_{0}^{a}r^2\text{g}_{\text{OO}}(r)dr,
904 chrisfen 1017 \end{equation}
905 chrisfen 1027 \begin{equation}
906 gezelter 1029 n_H = 4\pi\rho_{\text{OH}}\int_{0}^{b}r^2\text{g}_{\text{OH}}(r)dr,
907 chrisfen 1027 \end{equation}
908     where $\rho$ is the number density of the specified pair interactions,
909     $a$ and $b$ are the radial locations of the minima following the first
910 gezelter 1029 peak in g$_\text{OO}(r)$ or g$_\text{OH}(r)$ respectively. The number
911     of hydrogen bonds stays relatively constant across all of the models,
912     but the coordination numbers of {\sc ssd/e} and {\sc ssd/rf} show an improvement
913     over {\sc ssd1}. This improvement is primarily due to extension of the
914     first solvation shell in the new parameter sets. Because $n_H$ and
915     $n_C$ are nearly identical in {\sc ssd1}, it appears that all molecules in
916     the first solvation shell are involved in hydrogen bonds. Since $n_H$
917     and $n_C$ differ in the newly parameterized models, the orientations
918     in the first solvation shell are a bit more ``fluid''. Therefore {\sc ssd1}
919     overstructures the first solvation shell and our reparameterizations
920     have returned this shell to more realistic liquid-like behavior.
921 chrisfen 1017
922 gezelter 1029 The time constants for the orientational autocorrelation functions
923 chrisfen 1017 are also displayed in Table \ref{liquidproperties}. The dipolar
924 gezelter 1029 orientational time correlation functions ($C_{l}$) are described
925 chrisfen 1017 by:
926     \begin{equation}
927 gezelter 1029 C_{l}(t) = \langle P_l[\hat{\mathbf{u}}_j(0)\cdot\hat{\mathbf{u}}_j(t)]\rangle,
928 chrisfen 1017 \end{equation}
929 gezelter 1029 where $P_l$ are Legendre polynomials of order $l$ and
930     $\hat{\mathbf{u}}_j$ is the unit vector pointing along the molecular
931     dipole.\cite{Rahman71} From these correlation functions, the
932     orientational relaxation time of the dipole vector can be calculated
933     from an exponential fit in the long-time regime ($t >
934     \tau_l$).\cite{Rothschild84} Calculation of these time constants were
935     averaged over five detailed NVE simulations performed at the ambient
936     conditions for each of the respective models. It should be noted that
937     the commonly cited value of 1.9 ps for $\tau_2$ was determined from
938     the NMR data in Ref. \citen{Krynicki66} at a temperature near
939     34$^\circ$C.\cite{Rahman71} Because of the strong temperature
940     dependence of $\tau_2$, it is necessary to recalculate it at 298 K to
941     make proper comparisons. The value shown in Table
942 chrisfen 1022 \ref{liquidproperties} was calculated from the same NMR data in the
943 gezelter 1029 fashion described in Ref. \citen{Krynicki66}. Similarly, $\tau_1$ was
944     recomputed for 298 K from the data in Ref. \citen{Eisenberg69}.
945     Again, {\sc ssd/e} and {\sc ssd/rf} show improved behavior over {\sc ssd1}, both with
946 chrisfen 1027 and without an active reaction field. Turning on the reaction field
947 gezelter 1029 leads to much improved time constants for {\sc ssd1}; however, these results
948     also include a corresponding decrease in system density.
949     Orientational relaxation times published in the original {\sc ssd} dynamics
950     papers are smaller than the values observed here, and this difference
951     can be attributed to the use of the Ewald sum.\cite{Ichiye99}
952 chrisfen 1017
953 chrisfen 743 \subsection{Additional Observations}
954    
955     \begin{figure}
956 chrisfen 862 \begin{center}
957     \epsfxsize=6in
958 chrisfen 1027 \epsfbox{icei_bw.eps}
959 gezelter 1029 \caption{The most stable crystal structure assumed by the {\sc ssd} family
960     of water models. We refer to this structure as Ice-{\it i} to
961     indicate its origins in computer simulation. This image was taken of
962     the (001) face of the crystal.}
963 chrisfen 743 \label{weirdice}
964 chrisfen 862 \end{center}
965 chrisfen 743 \end{figure}
966    
967 gezelter 921 While performing a series of melting simulations on an early iteration
968 gezelter 1029 of {\sc ssd/e} not discussed in this paper, we observed recrystallization
969 gezelter 921 into a novel structure not previously known for water. After melting
970     at 235 K, two of five systems underwent crystallization events near
971     245 K. The two systems remained crystalline up to 320 and 330 K,
972     respectively. The crystal exhibits an expanded zeolite-like structure
973     that does not correspond to any known form of ice. This appears to be
974     an artifact of the point dipolar models, so to distinguish it from the
975     experimentally observed forms of ice, we have denoted the structure
976 gezelter 1029 Ice-$\sqrt{\smash[b]{-\text{I}}}$ (Ice-{\it i}). A large enough
977 gezelter 921 portion of the sample crystallized that we have been able to obtain a
978 gezelter 1029 near ideal crystal structure of Ice-{\it i}. Figure \ref{weirdice}
979 gezelter 921 shows the repeating crystal structure of a typical crystal at 5
980     K. Each water molecule is hydrogen bonded to four others; however, the
981     hydrogen bonds are bent rather than perfectly straight. This results
982     in a skewed tetrahedral geometry about the central molecule. In
983     figure \ref{isosurface}, it is apparent that these flexed hydrogen
984     bonds are allowed due to the conical shape of the attractive regions,
985     with the greatest attraction along the direct hydrogen bond
986 chrisfen 863 configuration. Though not ideal, these flexed hydrogen bonds are
987 gezelter 921 favorable enough to stabilize an entire crystal generated around them.
988 chrisfen 743
989 gezelter 1029 Initial simulations indicated that Ice-{\it i} is the preferred ice
990     structure for at least the {\sc ssd/e} model. To verify this, a
991     comparison was made between near ideal crystals of ice~$I_h$,
992     ice~$I_c$, and Ice-{\it i} at constant pressure with {\sc ssd/e}, {\sc
993     ssd/rf}, and {\sc ssd1}. Near-ideal versions of the three types of
994     crystals were cooled to 1 K, and enthalpies of formation of each were
995     compared using all three water models. Enthalpies were estimated from
996     the isobaric-isothermal simulations using $H=U+P_{\text ext}V$ where
997     $P_{\text ext}$ is the applied pressure. A constant value of
998     -60.158 kcal / mol has been added to place our zero for the
999     enthalpies of formation for these systems at the traditional state
1000     (elemental forms at standard temperature and pressure). With every
1001     model in the {\sc ssd} family, Ice-{\it i} had the lowest calculated
1002     enthalpy of formation.
1003 chrisfen 743
1004 gezelter 921 \begin{table}
1005     \begin{center}
1006 gezelter 1029 \caption{Enthalpies of Formation (in kcal / mol) of the three crystal
1007     structures (at 1 K) exhibited by the {\sc ssd} family of water models}
1008 gezelter 921 \begin{tabular}{ l c c c }
1009     \hline \\[-3mm]
1010     \ \ \ Water Model \ \ \ & \ \ \ Ice-$I_h$ \ \ \ & \ Ice-$I_c$\ \ & \
1011     Ice-{\it i} \\
1012     \hline \\[-3mm]
1013 gezelter 1029 \ \ \ {\sc ssd/e} & -12.286 & -12.292 & -13.590 \\
1014     \ \ \ {\sc ssd/rf} & -12.935 & -12.917 & -14.022 \\
1015     \ \ \ {\sc ssd1} & -12.496 & -12.411 & -13.417 \\
1016     \ \ \ {\sc ssd1} (RF) & -12.504 & -12.411 & -13.134 \\
1017 gezelter 921 \end{tabular}
1018     \label{iceenthalpy}
1019     \end{center}
1020     \end{table}
1021 chrisfen 743
1022 gezelter 921 In addition to these energetic comparisons, melting simulations were
1023 gezelter 1029 performed with ice-{\it i} as the initial configuration using {\sc ssd/e},
1024     {\sc ssd/rf}, and {\sc ssd1} both with and without a reaction field. The melting
1025     transitions for both {\sc ssd/e} and {\sc ssd1} without reaction field occurred at
1026     temperature in excess of 375~K. {\sc ssd/rf} and {\sc ssd1} with a reaction field
1027 gezelter 921 showed more reasonable melting transitions near 325~K. These melting
1028 gezelter 1029 point observations clearly show that all of the {\sc ssd}-derived models
1029 gezelter 921 prefer the ice-{\it i} structure.
1030 chrisfen 743
1031     \section{Conclusions}
1032    
1033 gezelter 921 The density maximum and temperature dependence of the self-diffusion
1034 gezelter 1029 constant were studied for the {\sc ssd} water model, both with and without
1035 gezelter 921 the use of reaction field, via a series of NPT and NVE
1036     simulations. The constant pressure simulations showed a density
1037     maximum near 260 K. In most cases, the calculated densities were
1038     significantly lower than the densities obtained from other water
1039 gezelter 1029 models (and experiment). Analysis of self-diffusion showed {\sc ssd} to
1040 gezelter 921 capture the transport properties of water well in both the liquid and
1041 chrisfen 1027 supercooled liquid regimes.
1042 gezelter 921
1043 gezelter 1029 In order to correct the density behavior, the original {\sc ssd} model was
1044     reparameterized for use both with and without a reaction field ({\sc ssd/rf}
1045     and {\sc ssd/e}), and comparisons were made with {\sc ssd1}, Ichiye's density
1046     corrected version of {\sc ssd}. Both models improve the liquid structure,
1047 gezelter 921 densities, and diffusive properties under their respective simulation
1048     conditions, indicating the necessity of reparameterization when
1049     changing the method of calculating long-range electrostatic
1050     interactions. In general, however, these simple water models are
1051     excellent choices for representing explicit water in large scale
1052     simulations of biochemical systems.
1053    
1054     The existence of a novel low-density ice structure that is preferred
1055 gezelter 1029 by the {\sc ssd} family of water models is somewhat troubling, since liquid
1056 gezelter 921 simulations on this family of water models at room temperature are
1057 chrisfen 1027 effectively simulations of supercooled or metastable liquids. One
1058     way to destabilize this unphysical ice structure would be to make the
1059 gezelter 921 range of angles preferred by the attractive part of the sticky
1060     potential much narrower. This would require extensive
1061     reparameterization to maintain the same level of agreement with the
1062     experiments.
1063    
1064 gezelter 1029 Additionally, our initial calculations show that the Ice-{\it i}
1065 gezelter 921 structure may also be a preferred crystal structure for at least one
1066     other popular multi-point water model (TIP3P), and that much of the
1067     simulation work being done using this popular model could also be at
1068     risk for crystallization into this unphysical structure. A future
1069     publication will detail the relative stability of the known ice
1070     structures for a wide range of popular water models.
1071    
1072 chrisfen 743 \section{Acknowledgments}
1073 chrisfen 777 Support for this project was provided by the National Science
1074     Foundation under grant CHE-0134881. Computation time was provided by
1075     the Notre Dame Bunch-of-Boxes (B.o.B) computer cluster under NSF grant
1076 gezelter 921 DMR-0079647.
1077 chrisfen 743
1078 chrisfen 862 \newpage
1079    
1080 chrisfen 743 \bibliographystyle{jcp}
1081     \bibliography{nptSSD}
1082    
1083     %\pagebreak
1084    
1085     \end{document}