ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/ssdePaper/nptSSD.tex
(Generate patch)

Comparing trunk/ssdePaper/nptSSD.tex (file contents):
Revision 743 by chrisfen, Wed Sep 3 21:36:09 2003 UTC vs.
Revision 1029 by gezelter, Thu Feb 5 20:47:50 2004 UTC

# Line 1 | Line 1
1 < \documentclass[prb,aps,times,twocolumn,tabularx]{revtex4}
2 < %\documentclass[prb,aps,times,tabularx,preprint]{revtex4}
1 > %\documentclass[prb,aps,times,twocolumn,tabularx]{revtex4}
2 > \documentclass[11pt]{article}
3 > \usepackage{endfloat}
4   \usepackage{amsmath}
5 + \usepackage{epsf}
6 + \usepackage{berkeley}
7 + \usepackage{setspace}
8 + \usepackage{tabularx}
9   \usepackage{graphicx}
10 <
6 < %\usepackage{endfloat}
10 > \usepackage[ref]{overcite}
11   %\usepackage{berkeley}
8 %\usepackage{epsf}
9 %\usepackage[ref]{overcite}
10 %\usepackage{setspace}
11 %\usepackage{tabularx}
12 %\usepackage{graphicx}
12   %\usepackage{curves}
13 < %\usepackage{amsmath}
14 < %\pagestyle{plain}
15 < %\pagenumbering{arabic}
16 < %\oddsidemargin 0.0cm \evensidemargin 0.0cm
17 < %\topmargin -21pt \headsep 10pt
18 < %\textheight 9.0in \textwidth 6.5in
19 < %\brokenpenalty=10000
20 < %\renewcommand{\baselinestretch}{1.2}
22 < %\renewcommand\citemid{\ } % no comma in optional reference note
13 > \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  
22   \begin{document}
23  
24 < \title{On the temperature dependent structural and transport properties of the soft sticky dipole (SSD) and related single point water models}
24 > \title{On the structural and transport properties of the soft sticky
25 > dipole ({\sc ssd}) and related single point water models}
26  
27 < \author{Christopher J. Fennell and J. Daniel Gezelter{\thefootnote}
28 < \footnote[1]{Corresponding author. \ Electronic mail: gezelter@nd.edu}}
30 <
31 < \address{Department of Chemistry and Biochemistry\\ University of Notre Dame\\
27 > \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   Notre Dame, Indiana 46556}
30  
31   \date{\today}
32  
33 + \maketitle
34 +
35   \begin{abstract}
36 < NVE and NPT molecular dynamics simulations were performed in order to
37 < investigate the density maximum and temperature dependent transport
38 < for the SSD water model, both with and without the use of reaction
39 < field. The constant pressure simulations of the melting of both $I_h$
40 < and $I_c$ ice showed a density maximum near 260 K. In most cases, the
41 < calculated densities were significantly lower than the densities
42 < calculated in simulations of other water models. Analysis of particle
43 < diffusion showed SSD to capture the transport properties of
44 < experimental very well in both the normal and super-cooled liquid
45 < regimes. In order to correct the density behavior, SSD was
46 < reparameterized for use both with and without a long-range interaction
47 < correction, SSD/RF and SSD/E respectively. In addition to correcting
48 < the abnormally low densities, these new versions were show to maintain
49 < or improve upon the transport and structural features of the original
50 < water model.
36 > The density maximum and temperature dependence of the self-diffusion
37 > constant were investigated for the soft sticky dipole ({\sc ssd}) water
38 > 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 > that the original {\sc ssd} model captures the transport properties of
48 > experimental water very well in both the normal and super-cooled
49 > liquid regimes.  We also present our re-parameterized versions of {\sc ssd}
50 > for use both with the reaction field or without any long-range
51 > electrostatic corrections.  These are called the {\sc ssd/rf} and {\sc ssd/e}
52 > models respectively.  These modified models were shown to maintain or
53 > improve upon the experimental agreement with the structural and
54 > 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 > Additionally, a novel low-density ice structure is presented
57 > which appears to be the most stable ice structure for the entire {\sc ssd}
58 > family.
59   \end{abstract}
60  
61 < \maketitle
61 > \newpage
62  
63   %\narrowtext
64  
# Line 62 | Line 69 | One of the most important tasks in simulations of bioc
69  
70   \section{Introduction}
71  
72 < One of the most important tasks in simulations of biochemical systems
73 < is the proper depiction of water and water solvation. In fact, the
74 < bulk of the calculations performed in solvated simulations are of
75 < interactions with or between solvent molecules. Thus, the outcomes of
76 < these types of simulations are highly dependent on the physical
77 < properties of water, both as individual molecules and in
78 < groups/bulk. Due to the fact that explicit solvent accounts for a
79 < massive portion of the calculations, it necessary to simplify the
73 < solvent to some extent in order to complete simulations in a
74 < reasonable amount of time. In the case of simulating water in
75 < bio-molecular studies, the balance between accurate properties and
76 < computational efficiency is especially delicate, and it has resulted
77 < in a variety of different water
78 < models.\cite{Jorgensen83,Berendsen87,Jorgensen00} Many of these models
79 < get specific properties correct or better than their predecessors, but
80 < this is often at a cost of some other properties or of computer
81 < time. As an example, compare TIP3P or TIP4P to TIP5P. TIP5P succeeds
82 < in improving the structural and transport properties over its
83 < predecessors, yet this comes at a greater than 50\% increase in
84 < computational cost.\cite{Jorgensen01,Jorgensen00} One recently
85 < developed model that succeeds in both retaining accuracy of system
86 < properties and simplifying calculations to increase computational
87 < efficiency is the Soft Sticky Dipole water model.\cite{Ichiye96}
72 > One of the most important tasks in the simulation of biochemical
73 > 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  
81 < The Soft Sticky Dipole (SSD)\ water model was developed by Ichiye
82 < \emph{et al.} as a modified form of the hard-sphere water model
83 < proposed by Bratko, Blum, and Luzar.\cite{Bratko85,Bratko95} SSD
84 < consists of a single point dipole with a Lennard-Jones core and a
85 < sticky potential that directs the particles to assume the proper
86 < hydrogen bond orientation in the first solvation shell. Thus, the
87 < interaction between two SSD water molecules \emph{i} and \emph{j} is
88 < given by the potential
81 > 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 > cost is the Soft Sticky Dipole ({\sc ssd}) water
93 > model.\cite{Ichiye96,Ichiye96b,Ichiye99,Ichiye03} The {\sc ssd} model was
94 > developed by Ichiye \emph{et al.} as a modified form of the
95 > hard-sphere water model proposed by Bratko, Blum, and
96 > Luzar.\cite{Bratko85,Bratko95} {\sc ssd} is a {\it single point} model which
97 > 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 > The interaction between two {\sc ssd} water molecules \emph{i} and \emph{j}
106 > is given by the potential
107   \begin{equation}
108   u_{ij} = u_{ij}^{LJ} (r_{ij})\ + u_{ij}^{dp}
109 < (\mathbf{r}_{ij},\boldsymbol{\Omega}_i,\boldsymbol{\Omega}_j)\ +
109 > ({\bf r}_{ij},{\bf \Omega}_i,{\bf \Omega}_j)\ +
110   u_{ij}^{sp}
111 < (\mathbf{r}_{ij},\boldsymbol{\Omega}_i,\boldsymbol{\Omega}_j),
111 > ({\bf r}_{ij},{\bf \Omega}_i,{\bf \Omega}_j),
112   \end{equation}
113 < where the $\mathbf{r}_{ij}$ is the position vector between molecules
114 < \emph{i} and \emph{j} with magnitude equal to the distance $r_ij$, and
115 < $\boldsymbol{\Omega}_i$ and $\boldsymbol{\Omega}_j$ represent the
116 < orientations of the respective molecules. The Lennard-Jones, dipole,
117 < and sticky parts of the potential are giving by the following
108 < equations,
113 > 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   \begin{equation}
119 < u_{ij}^{LJ}(r_{ij}) = 4\epsilon \left[\left(\frac{\sigma}{r_{ij}}\right)^{12}-\left(\frac{\sigma}{r_{ij}}\right)^{6}\right],
119 > 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   \end{equation}
123 + and
124   \begin{equation}
125 < 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}\ ,
125 > 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   \end{equation}
129 + 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   \begin{equation}
116 \begin{split}
137   u_{ij}^{sp}
138 < (\mathbf{r}_{ij},\boldsymbol{\Omega}_i,\boldsymbol{\Omega}_j)
139 < &=
140 < \frac{\nu_0}{2}[s(r_{ij})w(\mathbf{r}_{ij},\boldsymbol{\Omega}_i,\boldsymbol{\Omega}_j)\\
141 < & \quad \ +
142 < s^\prime(r_{ij})w^\prime(\mathbf{r}_{ij},\boldsymbol{\Omega}_i,\boldsymbol{\Omega}_j)]\ ,
123 < \end{split}
138 > ({\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 > \label{stickyfunction}
143   \end{equation}
144 < where $\boldsymbol{\mu}_i$ and $\boldsymbol{\mu}_j$ are the dipole
145 < unit vectors of particles \emph{i} and \emph{j} with magnitude 2.35 D,
146 < $\nu_0$ scales the strength of the overall sticky potential, $s$ and
147 < $s^\prime$ are cubic switching functions. The $w$ and $w^\prime$
148 < functions take the following forms,
144 > 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   \begin{equation}
150 < w(\mathbf{r}_{ij},\boldsymbol{\Omega}_i,\boldsymbol{\Omega}_j)=\sin\theta_{ij}\sin2\theta_{ij}\cos2\phi_{ij},
150 > w({\bf r}_{ij},{\bf \Omega}_i,{\bf \Omega}_j)=\sin\theta_{ij}\sin2\theta_{ij}\cos2\phi_{ij},
151   \end{equation}
152 + while the $w^\prime$ function counters the normal aligned and
153 + anti-aligned structures favored by point dipoles:
154   \begin{equation}
155 < 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,
155 > 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   \end{equation}
157 < where $w^0 = 0.07715$. The $w$ function is the tetrahedral attractive
158 < term that promotes hydrogen bonding orientations within the first
159 < solvation shell, and $w^\prime$ is a dipolar repulsion term that
160 < repels unrealistic dipolar arrangements within the first solvation
161 < shell. A more detailed description of the functional parts and
162 < variables in this potential can be found in other
163 < articles.\cite{Ichiye96,Ichiye99}
157 > 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 > can be found in the original {\sc ssd}
163 > articles.\cite{Ichiye96,Ichiye96b,Ichiye99,Ichiye03}
164  
165 < Being that this is a one-site point dipole model, the actual force
166 < calculations are simplified significantly. In the original Monte Carlo
167 < simulations using this model, Ichiye \emph{et al.} reported a
168 < calculation speed up of up to an order of magnitude over other
169 < comparable models while maintaining the structural behavior of
170 < water.\cite{Ichiye96} In the original molecular dynamics studies of
171 < SSD, it was shown that it actually improves upon the prediction of
172 < water's dynamical properties 3 and 4-point models.\cite{Ichiye99} This
173 < attractive combination of speed and accurate depiction of solvent
174 < properties makes SSD a model of interest for the simulation of large
175 < scale biological systems, such as membrane phase behavior, a specific
176 < interest within our group.
165 > Since {\sc ssd} is a single-point {\it dipolar} model, the force
166 > 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 > {\sc ssd} decreased computer time by a factor of 6-7 compared to other
170 > 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 > properties.  Indeed, {\sc ssd} maintains reasonable agreement with the Soper
173 > data for the structural features of liquid
174 > water.\cite{Soper86,Ichiye96} Additionally, the dynamical properties
175 > exhibited by {\sc ssd} agree with experiment better than those of more
176 > computationally expensive models (like TIP3P and
177 > SPC/E).\cite{Ichiye99} The combination of speed and accurate depiction
178 > of solvent properties makes {\sc ssd} a very attractive model for the
179 > simulation of large scale biochemical simulations.
180  
181 < Up to this point, a detailed look at the model's structure and ion
182 < solvation abilities has been performed.\cite{Ichiye96} In addition, a
183 < thorough investigation of the dynamic properties of SSD was performed
184 < by Chandra and Ichiye focusing on translational and orientational
185 < properties at 298 K.\cite{Ichiye99} This study focuses on determining
186 < the density maximum for SSD utilizing both microcanonical and
187 < isobaric-isothermal ensemble molecular dynamics, while using the
188 < reaction field method for handling long-ranged dipolar interactions. A
189 < reaction field method has been previously implemented in Monte Carlo
190 < simulations by Liu and Ichiye in order to study the static dielectric
191 < constant for the model.\cite{Ichiye96b} This paper will expand the
192 < scope of these original simulations to look on how the reaction field
193 < affects the physical and dynamic properties of SSD systems.
181 > One feature of the {\sc ssd} model is that it was parameterized for use with
182 > 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 > transport behavior of {\sc ssd} over a variety of temperatures with the
197 > 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 > 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 > behavior (also observed here), calling the corrected model
203 > {\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 > model in simulations utilizing the Reaction Field.
207  
208   \section{Methods}
209  
210 < As stated previously, in this study the long-range dipole-dipole
211 < interactions were accounted for using the reaction field method. The
212 < magnitude of the reaction field acting on dipole \emph{i} is given by
210 > 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 > 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   \begin{equation}
217   \mathcal{E}_{i} = \frac{2(\varepsilon_{s} - 1)}{2\varepsilon_{s} + 1}
218 < \frac{1}{r_{c}^{3}} \sum_{j\in{\mathcal{R}}} \boldsymbol{\mu}_{j} f(r_{ij})\  ,
218 > \frac{1}{r_{c}^{3}} \sum_{j\in{\mathcal{R}}} {\bf \mu}_{j} s(r_{ij}),
219   \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 < system (80 in this case), $\boldsymbol{\mu}_{j}$ is the dipole moment
224 < vector of particle \emph{j}, and $f(r_{ij})$ is a cubic switching
223 > system (80 in the case of liquid water), ${\bf \mu}_{j}$ is the dipole
224 > moment vector of particle $j$, and $s(r_{ij})$ is a cubic switching
225   function.\cite{AllenTildesley} The reaction field contribution to the
226 < total energy by particle \emph{i} is given by
227 < $-\frac{1}{2}\boldsymbol{\mu}_{i}\cdot\mathcal{E}_{i}$ and the torque
228 < on dipole \emph{i} by
229 < $\boldsymbol{\mu}_{i}\times\mathcal{E}_{i}$.\cite{AllenTildesley} Use
230 < of reaction field is known to alter the orientational dynamic
231 < properties, such as the dielectric relaxation time, based on changes
232 < in the length of the cutoff radius.\cite{Berendsen98} This variable
233 < behavior makes reaction field a less attractive method than other
234 < methods, like the Ewald summation; however, for the simulation of
195 < large-scale system, the computational cost benefit of reaction field
196 < is dramatic. To address some of the dynamical property alterations due
197 < to the use of reaction field, simulations were also performed without
198 < a surrounding dielectric and suggestions are proposed on how to make
199 < SSD more compatible with a reaction field.
226 > 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 < Simulations were performed in both the isobaric-isothermal and
237 < microcanonical ensembles. The constant pressure simulations were
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 > at the cutoff radius), and as a result we have two reparamaterizations
239 > of {\sc ssd} which could be used either with or without the reaction field
240 > turned on.
241 >
242 > 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} For the constant pressure
249 < simulations, the \emph{Q} parameter for the was set to 5.0 amu
250 < \(\cdot\)\AA\(^{2}\), and the relaxation time (\(\tau\))\ was set at
251 < 100 ps.
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  
253   Integration of the equations of motion was carried out using the
254 < symplectic splitting method proposed by Dullweber \emph{et
255 < al.}.\cite{Dullweber1997} The reason for this integrator selection
256 < deals with poor energy conservation of rigid body systems using
257 < quaternions. While quaternions work well for orientational motion in
258 < alternate ensembles, the microcanonical ensemble has a constant energy
259 < requirement that is actually quite sensitive to errors in the
260 < equations of motion. The original implementation of this code utilized
261 < quaternions for rotational motion propagation; however, a detailed
218 < investigation showed that they resulted in a steady drift in the total
219 < energy, something that has been observed by others.\cite{Laird97}
254 > symplectic splitting method proposed by Dullweber, Leimkuhler, and
255 > McLachlan ({\sc dlm}).\cite{Dullweber1997} Our reason for selecting this
256 > 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 > using quaternions was substantially greater than when using the {\sc dlm}
260 > method (fig. \ref{timestep}).  This steady drift in the total energy
261 > has also been observed by Kol {\it et al.}\cite{Laird97}
262  
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 < one time step to the next. In the past, this would not have been as
266 < feasible a option, being that the rotation matrix for a single body is
267 < nine elements long as opposed to 3 or 4 elements for Euler angles and
268 < quaternions respectively. System memory has become much less of an
227 < issue in recent times, and this has resulted in substantial benefits
228 < in energy conservation. There is still the issue of an additional 5 or
229 < 6 additional elements for describing the orientation of each particle,
230 < which will increase dump files substantially. Simply translating the
231 < rotation matrix into its component Euler angles or quaternions for
232 < storage purposes relieves this burden.
265 > 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  
270 < The symplectic splitting method allows for Verlet style integration of
271 < both linear and angular motion of rigid bodies. In the integration
272 < method, the orientational propagation involves a sequence of matrix
273 < evaluations to update the rotation matrix.\cite{Dullweber1997} These
274 < matrix rotations end up being more costly computationally than the
275 < simpler arithmetic quaternion propagation. On average, a 1000 SSD
276 < particle simulation shows a 7\% increase in simulation time using the
277 < symplectic step method in place of quaternions. This cost is more than
278 < justified when comparing the energy conservation of the two methods as
279 < illustrated in figure \ref{timestep}.
270 > The {\sc dlm} method allows for Verlet style integration of both
271 > translational and orientational motion of rigid bodies. In this
272 > 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 > 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 > additional expense per step is justified when one considers the
279 > ability to use time steps that are nearly twice as large under {\sc dlm}
280 > 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 > \ref{timestep}.
284  
285   \begin{figure}
286 < \includegraphics[width=61mm, angle=-90]{timeStep.epsi}
287 < \caption{Energy conservation using quaternion based integration versus
288 < the symplectic step method proposed by Dullweber \emph{et al.} with
289 < increasing time step. For each time step, the dotted line is total
290 < energy using the symplectic step integrator, and the solid line comes
291 < from the quaternion integrator. The larger time step plots are shifted
292 < up from the true energy baseline for clarity.}
286 > \begin{center}
287 > \epsfxsize=6in
288 > \epsfbox{timeStep.epsi}
289 > \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   \label{timestep}
294 + \end{center}
295   \end{figure}
296  
297   In figure \ref{timestep}, the resulting energy drift at various time
298 < steps for both the symplectic step and quaternion integration schemes
299 < is compared. All of the 1000 SSD particle simulations started with the
300 < same configuration, and the only difference was the method for
301 < handling rotational motion. At time steps of 0.1 and 0.5 fs, both
302 < methods for propagating particle rotation conserve energy fairly well,
303 < with the quaternion method showing a slight energy drift over time in
304 < the 0.5 fs time step simulation. At time steps of 1 and 2 fs, the
305 < energy conservation benefits of the symplectic step method are clearly
306 < demonstrated.
307 <
308 < Energy drift in these SSD particle simulations was unnoticeable for
268 < 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.
298 > 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 > 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 > fs, the energy conservation benefits of the {\sc dlm} method are clearly
306 > 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  
310 < Ice crystals in both the $I_h$ and $I_c$ lattices were generated as
311 < starting points for all the simulations. The $I_h$ crystals were
312 < formed by first arranging the center of masses of the SSD particles
313 < into a ``hexagonal'' ice lattice of 1024 particles. Because of the
314 < crystal structure of $I_h$ ice, the simulation box assumed a
315 < rectangular shape with a edge length ratio of approximately
310 > Energy drift in the simulations using {\sc dlm} integration was unnoticeable
311 > 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 >
318 > 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 > crystals were formed by first arranging the centers of mass of the {\sc ssd}
321 > 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   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 < down in stages to slowly cool the crystals down to 25 K. The particles
328 < were then allowed translate with fixed orientations at a constant
327 > 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   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 < rules\cite{Bernal33,Rahman72}.  This method was also utilized in the
333 > rules.\cite{Bernal33,Rahman72} This method was also utilized in the
334   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,
# Line 297 | Line 340 | constant pressure and temperature dynamics. This invol
340   \section{Results and discussion}
341  
342   Melting studies were performed on the randomized ice crystals using
343 < constant pressure and temperature dynamics. This involved an initial
344 < randomization of velocities about the starting temperature of 25 K for
345 < varying amounts of time. The systems were all equilibrated for 100 ps
346 < prior to a 200 ps data collection run at each temperature setting,
347 < ranging from 25 to 400 K, with a maximum degree increment of 25 K. For
348 < regions of interest along this stepwise progression, the temperature
349 < increment was decreased from 25 K to 10 and then 5 K. The above
350 < equilibration and production times were sufficient in that the system
351 < volume fluctuations dampened out in all but the very cold simulations
352 < (below 225 K). In order to further improve statistics, five separate
353 < simulation progressions were performed, and the averaged results from
354 < the $I_h$ melting simulations are shown in figure \ref{dense1}.
343 > 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  
313 \begin{figure}
314 \includegraphics[width=65mm, angle=-90]{1hdense.epsi}
315 \caption{Average density of SSD water at increasing temperatures
316 starting from ice $I_h$ lattice.}
317 \label{dense1}
318 \end{figure}
319
357   \subsection{Density Behavior}
321 In the initial average density versus temperature plot, the density
322 maximum clearly appears between 255 and 265 K. The calculated
323 densities within this range were nearly indistinguishable, as can be
324 seen in the zoom of this region of interest, shown in figure
325 \ref{dense1}. The greater certainty of the average value at 260 K makes
326 a good argument for the actual density maximum residing at this
327 midpoint value. Figure \ref{dense1} was constructed using ice $I_h$
328 crystals for the initial configuration; and though not pictured, the
329 simulations starting from ice $I_c$ crystal configurations showed
330 similar results, with a liquid-phase density maximum in this same
331 region (between 255 and 260 K). In addition, the $I_c$ crystals are
332 more fragile than the $I_h$ crystals, leading them to deform into a
333 dense glassy state at lower temperatures. This resulted in an overall
334 low temperature density maximum at 200 K, but they still retained a
335 common liquid state density maximum with the $I_h$ simulations.
358  
359 + Our initial simulations focused on the original {\sc ssd} water model, and
360 + 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   \begin{figure}
371 < \includegraphics[width=65mm,angle=-90]{dense2.eps}
372 < \caption{Density versus temperature for TIP4P\cite{Jorgensen98b},
373 < TIP3P\cite{Jorgensen98b}, SPC/E\cite{Clancy94}, SSD without Reaction
374 < Field, SSD, and Experiment\cite{CRC80}. }
375 < \label{dense2}
371 > \begin{center}
372 > \epsfxsize=6in
373 > \epsfbox{denseSSD.eps}
374 > \caption{Density versus temperature for TIP4P [Ref. \citen{Jorgensen98b}],
375 > TIP3P [Ref. \citen{Jorgensen98b}], SPC/E [Ref. \citen{Clancy94}], {\sc ssd}
376 > without Reaction Field, {\sc ssd}, and experiment [Ref. \citen{CRC80}]. The
377 > arrows indicate the change in densities observed when turning off the
378 > reaction field. The the lower than expected densities for the {\sc ssd}
379 > model were what prompted the original reparameterization of {\sc ssd1}
380 > [Ref. \citen{Ichiye03}].}
381 > \label{dense1}
382 > \end{center}
383   \end{figure}
384  
385 < The density maximum for SSD actually compares quite favorably to other
386 < simple water models. Figure \ref{dense2} shows a plot of these
387 < findings with the density progression of several other models and
348 < experiment obtained from other
385 > The density maximum for {\sc ssd} compares quite favorably to other simple
386 > water models. Figure \ref{dense1} also shows calculated densities of
387 > several other models and experiment obtained from other
388   sources.\cite{Jorgensen98b,Clancy94,CRC80} Of the listed simple water
389 < models, SSD has results closest to the experimentally observed water
390 < density maximum. Of the listed water models, TIP4P has a density
391 < maximum behavior most like that seen in SSD. Though not shown, it is
392 < useful to note that TIP5P has a water density maximum nearly identical
393 < to experiment.
389 > models, {\sc ssd} has a temperature closest to the experimentally observed
390 > density maximum. Of the {\it charge-based} models in
391 > Fig. \ref{dense1}, TIP4P has a density maximum behavior most like that
392 > seen in {\sc ssd}. Though not included in this plot, it is useful
393 > to note that TIP5P has a density maximum nearly identical to the
394 > experimentally measured temperature.
395  
396 < Possibly of more importance is the density scaling of SSD relative to
397 < other common models at any given temperature (Fig. \ref{dense2}). Note
398 < that the SSD model assumes a lower density than any of the other
399 < listed models at the same pressure, behavior which is especially
400 < apparent at temperatures greater than 300 K. Lower than expected
401 < densities have been observed for other systems with the use of a
402 < reaction field for long-range electrostatic interactions, so the most
403 < likely reason for these significantly lower densities in these
404 < simulations is the presence of the reaction field.\cite{Berendsen98}
405 < In order to test the effect of the reaction field on the density of
406 < the systems, the simulations were repeated for the temperature region
407 < of interest without a reaction field present. The results of these
408 < simulations are also displayed in figure \ref{dense2}. Without
369 < reaction field, these densities increase considerably to more
370 < experimentally reasonable values, especially around the freezing point
371 < of liquid water. The shape of the curve is similar to the curve
372 < produced from SSD simulations using reaction field, specifically the
373 < rapidly decreasing densities at higher temperatures; however, a slight
374 < shift in the density maximum location, down to 245 K, is
375 < observed. This is probably a more accurate comparison to the other
376 < listed water models in that no long range corrections were applied in
377 < those simulations.\cite{Clancy94,Jorgensen98b}
396 > 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 > cutoff radius of 12.0 \AA\ to complement the previous {\sc ssd} simulations,
401 > 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 > simulations using {\sc ssd}. This is advantageous in that the use of a
407 > longer cutoff radius results in a significant increase in the time
408 > required to obtain a single trajectory.
409  
410 < It has been observed that densities are dependent on the cutoff radius
411 < used for a variety of water models in simulations both with and
412 < without the use of reaction field.\cite{Berendsen98} In order to
413 < address the possible affect of cutoff radius, simulations were
414 < performed with a dipolar cutoff radius of 12.0 \AA\ to compliment the
415 < previous SSD simulations, all performed with a cutoff of 9.0 \AA. All
416 < the resulting densities overlapped within error and showed no
417 < significant trend in lower or higher densities as a function of cutoff
418 < radius, both for simulations with and without reaction field. These
419 < results indicate that there is no major benefit in choosing a longer
420 < cutoff radius in simulations using SSD. This is comforting in that the
421 < use of a longer cutoff radius results in a near doubling of the time
422 < required to compute a single trajectory.
410 > The key feature to recognize in figure \ref{dense1} is the density
411 > 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 > 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 > the curve produced from {\sc ssd} simulations using reaction field,
426 > 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 > reaction field, the density around 300 K is still significantly lower
432 > than experiment and comparable water models. This anomalous behavior
433 > was what lead Tan {\it et al.} to recently reparameterize
434 > {\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  
438   \subsection{Transport Behavior}
439 < Of importance in these types of studies are the transport properties
440 < of the particles and how they change when altering the environmental
441 < conditions. In order to probe transport, constant energy simulations
442 < were performed about the average density uncovered by the constant
443 < pressure simulations. Simulations started with randomized velocities
444 < and underwent 50 ps of temperature scaling and 50 ps of constant
445 < energy equilibration before obtaining a 200 ps trajectory. Diffusion
446 < constants were calculated via root-mean square deviation analysis. The
447 < averaged results from 5 sets of these NVE simulations is displayed in
448 < figure \ref{diffuse}, alongside experimental, SPC/E, and TIP5P
449 < results.\cite{Gillen72,Mills73,Clancy94,Jorgensen01}
439 >
440 > 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 > results.\cite{Gillen72,Holz00,Clancy94,Jorgensen01}
453  
454   \begin{figure}
455 < \includegraphics[width=65mm, angle=-90]{betterDiffuse.epsi}
456 < \caption{Average diffusion coefficient over increasing temperature for
457 < SSD, SPC/E\cite{Clancy94}, TIP5P\cite{Jorgensen01}, and Experimental
458 < data from Gillen \emph{et al.}\cite{Gillen72}, and from
459 < Mills\cite{Mills73}.}
455 > \begin{center}
456 > \epsfxsize=6in
457 > \epsfbox{betterDiffuse.epsi}
458 > \caption{Average self-diffusion constant as a function of temperature for
459 > {\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   \label{diffuse}
466 + \end{center}
467   \end{figure}
468  
469   The observed values for the diffusion constant point out one of the
470 < strengths of the SSD model. Of the three experimental models shown,
471 < the SSD model has the most accurate depiction of the diffusion trend
472 < seen in experiment in both the supercooled and normal regimes. SPC/E
473 < does a respectable job by getting similar values as SSD and experiment
474 < around 290 K; however, it deviates at both higher and lower
475 < temperatures, failing to predict the experimental trend. TIP5P and SSD
476 < both start off low at the colder temperatures and tend to diffuse too
477 < rapidly at the higher temperatures. This type of trend at the higher
478 < temperatures is not surprising in that the densities of both TIP5P and
479 < SSD are lower than experimental water at temperatures higher than room
480 < temperature. When calculating the diffusion coefficients for SSD at
481 < experimental densities, the resulting values fall more in line with
482 < experiment at these temperatures, albeit not at standard
429 < pressure. Results under these conditions can be found later in this
430 < paper.
470 > strengths of the {\sc ssd} model. Of the three models shown, the {\sc ssd} model
471 > 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 > correct thermal trend. TIP5P and {\sc ssd} both start off low at colder
476 > temperatures and tend to diffuse too rapidly at higher temperatures.
477 > This behavior at higher temperatures is not particularly surprising
478 > since the densities of both TIP5P and {\sc ssd} are lower than experimental
479 > water densities at higher temperatures.  When calculating the
480 > diffusion coefficients for {\sc ssd} at experimental densities (instead of
481 > the densities from the NPT simulations), the resulting values fall
482 > more in line with experiment at these temperatures.
483  
484   \subsection{Structural Changes and Characterization}
485 +
486   By starting the simulations from the crystalline state, the melting
487 < transition and the ice structure can be studied along with the liquid
488 < phase behavior beyond the melting point. To locate the melting
489 < transition, the constant pressure heat capacity (C$_\text{p}$) was
490 < monitored in each of the simulations. In the melting simulations of
491 < the 1024 particle ice $I_h$ simulations, a large spike in C$_\text{p}$
492 < occurs at 245 K, indicating a first order phase transition for the
493 < melting of these ice crystals. When the reaction field is turned off,
494 < the melting transition occurs at 235 K.  These melting transitions are
495 < considerably lower than the experimental value, but this is not
443 < surprising in that SSD is a simple rigid body model with a fixed
444 < dipole.
487 > transition and the ice structure can be obtained along with the liquid
488 > 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 > considerably lower than the experimental value.
496  
446 \begin{figure}
447 \includegraphics[width=85mm]{fullContours.eps}
448 \caption{Contour plots of 2D angular g($r$)'s for 512 SSD systems at
449 100 K (A \& B) and 300 K (C \& D). Contour colors are inverted for
450 clarity: dark areas signify peaks while light areas signify
451 depressions. White areas have g(\emph{r}) values below 0.5 and black
452 areas have values above 1.5.}
453 \label{contour}
454 \end{figure}
455
456 Additional analyses for understanding the melting phase-transition
457 process were performed via two-dimensional structure and dipole angle
458 correlations. Expressions for the correlations are as follows:
459
497   \begin{figure}
498 < \includegraphics[width=45mm]{corrDiag.eps}
499 < \caption{Two dimensional illustration of the angles involved in the
500 < correlations observed in figure \ref{contour}.}
498 > \begin{center}
499 > \epsfxsize=6in
500 > \epsfbox{corrDiag.eps}
501 > \caption{An illustration of angles involved in the correlations observed in Fig. \ref{contour}.}
502   \label{corrAngle}
503 + \end{center}
504   \end{figure}
505  
506 < \begin{multline}
507 < g_{\text{AB}}(r,\cos\theta) = \\
508 < \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\ ,
509 < \end{multline}
510 < \begin{multline}
511 < g_{\text{AB}}(r,\cos\omega) = \\
512 < \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\ ,
513 < \end{multline}
514 < where $\theta$ and $\omega$ refer to the angles shown in the above
515 < illustration. By binning over both distance and the cosine of the
516 < desired angle between the two dipoles, the g(\emph{r}) can be
517 < dissected to determine the common dipole arrangements that constitute
518 < the peaks and troughs. Frames A and B of figure \ref{contour} show a
519 < relatively crystalline state of an ice $I_c$ simulation. The first
520 < peak of the g(\emph{r}) primarily consists of the preferred hydrogen
521 < bonding arrangements as dictated by the tetrahedral sticky potential,
522 < one peak for the donating and the other for the accepting hydrogen
523 < bonds. Due to the high degree of crystallinity of the sample, the
524 < second and third solvation shells show a repeated peak arrangement
506 > \begin{figure}
507 > \begin{center}
508 > \epsfxsize=6in
509 > \epsfbox{fullContours.eps}
510 > \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 > \label{contour}
516 > \end{center}
517 > \end{figure}
518 >
519 > 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 >
523 > \begin{equation}
524 > 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 > \end{equation}
526 > \begin{equation}
527 > g_{\text{AB}}(r,\cos\omega) =
528 > \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 > \end{equation}
530 > 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 > 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 > bonding arrangements as dictated by the tetrahedral sticky potential -
538 > 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   which decays at distances around the fourth solvation shell, near the
542   imposed cutoff for the Lennard-Jones and dipole-dipole interactions.
543 < In the higher temperature simulation shown in frames C and D, the
544 < repeated peak features are significantly blurred. The first solvation
545 < shell still shows the strong effect of the sticky-potential, although
546 < it covers a larger area, extending to include a fraction of aligned
543 > In the higher temperature simulation shown in frames C and D, these
544 > 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 < definition as thermal motion and the competing dipole force overcomes
549 < the sticky potential's tight tetrahedral structuring of the fluid.
548 > due to thermal motion and as the competing dipole force overcomes the
549 > sticky potential's tight tetrahedral structuring of the crystal.
550  
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 < oxygen-oxygen g(\emph{r}).\cite{Ichiye96} At low temperatures, the
554 < second solvation shell peak appears to have two distinct parts that
553 > 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, and there is tightly
559 < bunched group of axially arranged dipoles that most likely consist of
560 < the smaller fraction aligned dipole pairs. The trailing part of the
561 < split peak at 5 \AA\ is dominated by aligned dipoles that range
562 < primarily within the axial to the chief hydrogen bond arrangements
563 < similar to those seen in the first solvation shell. This evidence
564 < indicates that the dipole pair interaction begins to dominate outside
565 < of the range of the dipolar repulsion term, with the primary
566 < energetically favorable dipole arrangements populating the region
567 < immediately outside of it's range (around 4 \AA), and arrangements
568 < that seek to ideally satisfy both the sticky and dipole forces locate
569 < themselves just beyond this region (around 5 \AA).
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  
571   From these findings, the split second peak is primarily the product of
572 < the dipolar repulsion term of the sticky potential. In fact, the
573 < leading of the two peaks can be pushed out and merged with the outer
574 < split peak just by extending the switching function cutoff
575 < ($s^\prime(r_{ij})$) from its normal 4.0 \AA\ to values of 4.5 or even
576 < 5 \AA. This type of correction is not recommended for improving the
577 < liquid structure, because the second solvation shell will still be
578 < shifted too far out. In addition, this would have an even more
579 < detrimental effect on the system densities, leading to a liquid with a
580 < more open structure and a density considerably lower than the normal
581 < SSD behavior shown previously. A better correction would be to include
582 < the quadrupole-quadrupole interactions for the water particles outside
583 < of the first solvation shell, but this reduces the simplicity and
584 < speed advantage of SSD, so it is not the most desirable path to take.
572 > 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 > 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 > correction is not recommended for improving the liquid structure,
577 > since the second solvation shell would still be shifted too far
578 > 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 > and a density considerably lower than the already low {\sc ssd} density.  A
581 > 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 > {\sc ssd}.
585  
586 < \subsection{Adjusted Potentials: SSD/E and SSD/RF}
587 < The propensity of SSD to adopt lower than expected densities under
586 > \subsection{Adjusted Potentials: {\sc ssd/rf} and {\sc ssd/e}}
587 >
588 > The propensity of {\sc ssd} to adopt lower than expected densities under
589   varying conditions is troubling, especially at higher temperatures. In
590 < order to correct this behavior, it's necessary to adjust the force
591 < field parameters for the primary intermolecular interactions. In
592 < undergoing a reparameterization, it is important not to focus on just
593 < one property and neglect the other important properties. In this case,
594 < it would be ideal to correct the densities while maintaining the
595 < accurate transport properties.
590 > 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 > densities while maintaining the accurate transport behavior.
596  
597 < The possible parameters for tuning include the $\sigma$ and $\epsilon$
598 < Lennard-Jones parameters, the dipole strength ($\mu$), and the sticky
599 < attractive and dipole repulsive terms with their respective
600 < cutoffs. To alter the attractive and repulsive terms of the sticky
601 < potential independently, it is necessary to separate the terms as
602 < follows:
603 < \begin{equation}
604 < \begin{split}
605 < u_{ij}^{sp}
606 < (\mathbf{r}_{ij},\boldsymbol{\Omega}_i,\boldsymbol{\Omega}_j) &=
550 < \frac{\nu_0}{2}[s(r_{ij})w(\mathbf{r}_{ij},\boldsymbol{\Omega}_i,\boldsymbol{\Omega}_j)]\\
551 < & \quad \ + \frac{\nu_0^\prime}{2}
552 < [s^\prime(r_{ij})w^\prime(\mathbf{r}_{ij},\boldsymbol{\Omega}_i,\boldsymbol{\Omega}_j)],
553 < \end{split}
554 < \end{equation}
597 > The parameters available for tuning include the $\sigma$ and
598 > $\epsilon$ Lennard-Jones parameters, the dipole strength ($\mu$), the
599 > 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  
556 where $\nu_0$ scales the strength of the tetrahedral attraction and
557 $\nu_0^\prime$ acts in an identical fashion on the dipole repulsion
558 term. For purposes of the reparameterization, the separation was
559 performed, but the final parameters were adjusted so that it is
560 unnecessary to separate the terms when implementing the adjusted water
561 potentials. The results of the reparameterizations are shown in table
562 \ref{params}. Note that both the tetrahedral attractive and dipolar
563 repulsive don't share the same lower cutoff ($r_l$) in the newly
564 parameterized potentials - soft sticky dipole enhanced (SSD/E) and
565 soft sticky dipole reaction field (SSD/RF).
566
608   \begin{table}
609 + \begin{center}
610   \caption{Parameters for the original and adjusted models}
611 < \begin{tabular}{ l  c  c  c }
611 > \begin{tabular}{ l  c  c  c  c }
612   \hline \\[-3mm]
613 < \ Parameters & \ \ \  SSD$^\dagger$\ \ \ \  & \ SSD/E\ \  & \ SSD/RF\ \ \\
613 > \ \ \ Parameters\ \ \  & \ \ \ {\sc ssd} [Ref. \citen{Ichiye96}] \ \ \
614 > & \ {\sc ssd1} [Ref. \citen{Ichiye03}]\ \  & \ {\sc ssd/e}\ \  & \ {\sc ssd/rf} \\
615   \hline \\[-3mm]
616 < \ \ \ $\sigma$ (\AA)  & 3.051 & 3.035 & 3.019\\
617 < \ \ \ $\epsilon$ (kcal/mol)\ \ & 0.152 & 0.152 & 0.152\\
618 < \ \ \ $\mu$ (D) & 2.35 & 2.418 & 2.480\\
619 < \ \ \ $\nu_0$ (kcal/mol)\ \ & 3.7284 & 3.90 & 3.90\\
620 < \ \ \ $r_l$ (\AA) & 2.75 & 2.40 & 2.40\\
621 < \ \ \ $r_u$ (\AA) & 3.35 & 3.80 & 3.80\\
622 < \ \ \ $\nu_0^\prime$ (kcal/mol)\ \ & 3.7284 & 3.90 & 3.90\\
623 < \ \ \ $r_l^\prime$ (\AA) & 2.75 & 2.75 & 2.75\\
624 < \ \ \ $r_u^\prime$ (\AA) & 4.00 & 3.35 & 3.35\\
582 < \\[-2mm]$^\dagger$ ref. \onlinecite{Ichiye96}
616 > \ \ \ $\sigma$ (\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 > \ \ \ $\omega^\circ$ & 0.07715 & 0.07715 & 0.07715 & 0.07715\\
621 > \ \ \ $r_l$ (\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   \end{tabular}
626   \label{params}
627 + \end{center}
628   \end{table}
629  
630 < \begin{figure}
631 < \includegraphics[width=85mm]{gofrCompare.epsi}
632 < \caption{Plots comparing experiment\cite{Head-Gordon00_1} with SSD/E
633 < and SSD without reaction field (top), as well as SSD/RF and SSD with
630 > \begin{figure}
631 > \begin{center}
632 > \epsfxsize=5in
633 > \epsfbox{GofRCompare.epsi}
634 > \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   reaction field turned on (bottom). The insets show the respective
637 < first peaks in detail. Solid Line - experiment, dashed line - SSD/E
638 < and SSD/RF, and dotted line - SSD (with and without reaction field).}
637 > first peaks in detail. Note how the changes in parameters have lowered
638 > and broadened the first peak of {\sc ssd/e} and {\sc ssd/rf}.}
639   \label{grcompare}
640 + \end{center}
641   \end{figure}
642  
643 < \begin{figure}
644 < \includegraphics[width=85mm]{dualsticky.ps}
645 < \caption{Isosurfaces of the sticky potential for SSD (left) and SSD/E \&
646 < SSD/RF (right). Light areas correspond to the tetrahedral attractive
647 < part, and the darker areas correspond to the dipolar repulsive part.}
643 > \begin{figure}
644 > \begin{center}
645 > \epsfxsize=6in
646 > \epsfbox{dualsticky_bw.eps}
647 > \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   \label{isosurface}
652 + \end{center}
653   \end{figure}
654  
655 < In the paper detailing the development of SSD, Liu and Ichiye placed
656 < particular emphasis on an accurate description of the first solvation
657 < shell. This resulted in a somewhat tall and sharp first peak that
658 < integrated to give similar coordination numbers to the experimental
659 < data obtained by Soper and Phillips.\cite{Ichiye96,Soper86} New
660 < experimental x-ray scattering data from the Head-Gordon lab indicates
661 < a slightly lower and shifted first peak in the g$_\mathrm{OO}(r)$, so
662 < adjustments to SSD were made while taking into consideration the new
663 < experimental findings.\cite{Head-Gordon00_1} Figure \ref{grcompare}
664 < shows the relocation of the first peak of the oxygen-oxygen
665 < g(\emph{r}) by comparing the original SSD (with and without reaction
666 < field), SSD-E, and SSD-RF to the new experimental results. Both the
667 < modified water models have shorter peaks that are brought in more
668 < closely to the experimental peak (as seen in the insets of figure
669 < \ref{grcompare}). This structural alteration was accomplished by a
670 < reduction in the Lennard-Jones $\sigma$ variable as well as adjustment
671 < of the sticky potential strength and cutoffs. The cutoffs for the
672 < tetrahedral attractive and dipolar repulsive terms were nearly swapped
673 < with each other. Isosurfaces of the original and modified sticky
674 < potentials are shown in figure \cite{isosurface}. In these
675 < isosurfaces, it is easy to see how altering the cutoffs changes the
676 < repulsive and attractive character of the particles. With a reduced
677 < repulsive surface (the darker region), the particles can move closer
678 < to one another, increasing the density for the overall system. This
679 < change in interaction cutoff also results in a more gradual
680 < orientational motion by allowing the particles to maintain preferred
681 < dipolar arrangements before they begin to feel the pull of the
682 < tetrahedral restructuring. Upon moving closer together, the dipolar
683 < repulsion term becomes active and excludes the unphysical
684 < arrangements. This compares with the original SSD's excluding dipolar
685 < before the particles feel the pull of the ``hydrogen bonds''. Aside
686 < from improving the shape of the first peak in the g(\emph{r}), this
687 < improves the densities considerably by allowing the persistence of
688 < full dipolar character below the previous 4.0 \AA\ cutoff.
655 > In the original paper detailing the development of {\sc ssd}, Liu and Ichiye
656 > 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 > 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 > 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 > findings.\cite{Head-Gordon00_1} Figure \ref{grcompare} shows the
665 > relocation of the first peak of the oxygen-oxygen $g(r)$ by comparing
666 > the revised {\sc ssd} model ({\sc ssd1}), {\sc ssd/e}, and {\sc ssd/rf} to the new
667 > experimental results. Both modified water models have shorter peaks
668 > that match more closely to the experimental peak (as seen in the
669 > insets of figure \ref{grcompare}).  This structural alteration was
670 > accomplished by the combined reduction in the Lennard-Jones $\sigma$
671 > 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 > how {\sc ssd} and {\sc ssd1} exclude preferred dipole alignments before the
686 > 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  
692 < While adjusting the location and shape of the first peak of
693 < g(\emph{r}) improves the densities to some degree, these changes alone
694 < are insufficient to bring the system densities up to the values
695 < observed experimentally. To finish bringing up the densities, the
696 < dipole moments were increased in both the adjusted models. Being a
697 < dipole based model, the structure and transport are very sensitive to
698 < changes in the dipole moment. The original SSD simply used the dipole
699 < moment calculated from the TIP3P water model, which at 2.35 D is
700 < significantly greater than the experimental gas phase value of 1.84
701 < D. The larger dipole moment is a more realistic value and improve the
702 < dielectric properties of the fluid. Both theoretical and experimental
703 < measurements indicate a liquid phase dipole moment ranging from 2.4 D
704 < to values as high as 3.11 D, so there is quite a range available for
705 < adjusting the dipole
706 < moment.\cite{Sprik91,Kusalik02,Badyal00,Barriol64} The increasing of
707 < the dipole moments to 2.418 and 2.48 D for SSD/E and SSD/RF
708 < respectively is moderate in the range of the experimental values;
657 < however, it leads to significant changes in the density and transport
658 < of the water models.
692 > 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 > both of our adjusted models. Since {\sc ssd} is a dipole based model, the
697 > structure and transport are very sensitive to changes in the dipole
698 > moment. The original {\sc ssd} simply used the dipole moment calculated from
699 > 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 > increasing the dipole moments to 2.42 and 2.48 D for {\sc ssd/e} and {\sc ssd/rf},
707 > respectively, leads to significant changes in the density and
708 > transport of the water models.
709  
710 < In order to demonstrate the benefits of this reparameterization, a
710 > In order to demonstrate the benefits of these reparameterizations, a
711   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 < to the original SSD model. This comparison involved full NPT melting
714 < sequences for both SSD/E and SSD/RF, as well as NVE transport
715 < calculations at both self-consistent and experimental
716 < densities. Again, the results come from five separate simulations of
717 < 1024 particle systems, and the melting sequences were started from
718 < different ice $I_h$ crystals constructed as stated previously. Like
719 < before, all of the NPT simulations were equilibrated for 100 ps before
720 < a 200 ps data collection run, and they used the previous temperature's
721 < final configuration as a starting point. All of the NVE simulations
722 < had the same thermalization, equilibration, and data collection times
723 < stated earlier in this paper.
713 > 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 > calculations at the calculated self-consistent densities. Again, the
716 > 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 > simulation was equilibrated for 100 ps before a 200 ps data collection
720 > 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 > collection times as stated previously.
724  
725 < \begin{figure}
726 < \includegraphics[width=85mm]{ssdecompare.epsi}
727 < \caption{Comparison of densities calculated with SSD/E to SSD without a
728 < reaction field, TIP4P\cite{Jorgensen98b}, TIP3P\cite{Jorgensen98b},
729 < SPC/E\cite{Clancy94}, and Experiment\cite{CRC80}. The upper plot
730 < includes error bars, and the calculated results from the other
731 < references were removed for clarity.}
725 > \begin{figure}
726 > \begin{center}
727 > \epsfxsize=6in
728 > \epsfbox{ssdeDense.epsi}
729 > \caption{Comparison of densities calculated with {\sc ssd/e} to {\sc ssd1} without a
730 > 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 > interest. Note that both {\sc ssd1} and {\sc ssd/e} show good agreement with
735 > experiment when the long-range correction is neglected.}
736   \label{ssdedense}
737 + \end{center}
738   \end{figure}
739  
740 < Figure \ref{ssdedense} shows the density profile for the SSD/E water
741 < model in comparison to the original SSD without a reaction field,
742 < experiment, and the other common water models considered
743 < previously. The calculated densities have increased significantly over
744 < the original SSD model and match the experimental value just below 298
745 < K. At 298 K, the density of SSD/E is 0.995$\pm$0.001 g/cm$^3$, which
746 < compares well with the experimental value of 0.997 g/cm$^3$ and is
747 < considerably better than the SSD value of 0.967$\pm$0.003
748 < g/cm$^3$. The increased dipole moment in SSD/E has helped to flatten
749 < out the curve at higher temperatures, only the improvement is marginal
750 < at best. This steep drop in densities is due to the dipolar rather
751 < than charge based interactions which decay more rapidly at longer
752 < distances.
753 <
754 < By monitoring C$\text{p}$ throughout these simulations, the melting
755 < transition for SSD/E was observed at 230 K, about 5 degrees lower than
756 < SSD. The resulting density maximum is located at 240 K, again about 5
757 < degrees lower than the SSD value of 245 K. Though there is a decrease
758 < in both of these values, the corrected densities near room temperature
759 < justify the modifications taken.
740 > 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 > models, and experimental results. The calculated densities for both
743 > {\sc ssd/e} and {\sc ssd1} have increased significantly over the original {\sc ssd}
744 > model (see fig. \ref{dense1}) and are in better agreement with the
745 > experimental values. At 298 K, the densities of {\sc ssd/e} and {\sc ssd1} without
746 > 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 > better than the {\sc ssd} value of 0.967$\pm$0.003 g/cm$^3$. The changes to
750 > 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 > character. However, this increasing disorder in the {\sc ssd/e} model has
756 > little effect on the melting transition. By monitoring $C_p$
757 > throughout these simulations, the melting transition for {\sc ssd/e} was
758 > shown to occur at 235 K.  The same transition temperature observed
759 > with {\sc ssd} and {\sc ssd1}.
760  
761 < \begin{figure}
762 < \includegraphics[width=85mm]{ssdrfcompare.epsi}
763 < \caption{Comparison of densities calculated with SSD/RF to SSD with a
764 < reaction field, TIP4P\cite{Jorgensen98b}, TIP3P\cite{Jorgensen98b},
765 < SPC/E\cite{Clancy94}, and Experiment\cite{CRC80}. The upper plot
766 < includes error bars, and the calculated results from the other
767 < references were removed for clarity.}
761 > \begin{figure}
762 > \begin{center}
763 > \epsfxsize=6in
764 > \epsfbox{ssdrfDense.epsi}
765 > \caption{Comparison of densities calculated with {\sc ssd/rf} to {\sc ssd1} with a
766 > 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 > correction - {\sc ssd/rf} provides significantly more accurate densities
771 > than {\sc ssd1} when performing room temperature simulations.}
772   \label{ssdrfdense}
773 + \end{center}
774   \end{figure}
775  
776 < Figure \ref{ssdrfdense} shows a density comparison between SSD/RF and
777 < SSD with an active reaction field. Like in the simulations of SSD/E,
778 < the densities show a dramatic increase over normal SSD. At 298 K,
779 < SSD/RF has a density of 0.997$\pm$0.001 g/cm$^3$, right in line with
780 < experiment and considerably better than the SSD value of
781 < 0.941$\pm$0.001 g/cm$^3$. The melting point is observed at 240 K,
782 < which is 5 degrees lower than SSD with a reaction field, and the
783 < density maximum at 255 K, again 5 degrees lower than SSD. The density
784 < at higher temperature still drops off more rapidly than the charge
785 < based models but is in better agreement than SSD/E.
776 > Including the reaction field long-range correction in the simulations
777 > results in a more interesting comparison.  A density profile including
778 > {\sc ssd/rf} and {\sc ssd1} with an active reaction field is shown in figure
779 > \ref{ssdrfdense}.  As observed in the simulations without a reaction
780 > field, the densities of {\sc ssd/rf} and {\sc ssd1} show a dramatic increase over
781 > normal {\sc ssd} (see figure \ref{dense1}). At 298 K, {\sc ssd/rf} has a density
782 > of 0.997$\pm$0.001 g/cm$^3$, directly in line with experiment and
783 > considerably better than the original {\sc ssd} value of 0.941$\pm$0.001
784 > g/cm$^3$ and the {\sc ssd1} value of 0.972$\pm$0.002 g/cm$^3$. These results
785 > 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 > 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 > maximum at 255 K, fairly close to the density maxima of 260 K and 265
792 > K, shown by {\sc ssd} and {\sc ssd1} respectively.
793  
794 < The reparameterization of the SSD water model, both for use with and
795 < without an applied long-range correction, brought the densities up to
796 < what is expected for simulating liquid water. In addition to improving
797 < the densities, it is important that particle transport be maintained
798 < or improved. Figure \ref{ssdediffuse} compares the temperature
799 < dependence of the diffusion constant of SSD/E to SSD without an active
800 < reaction field, both at the densities calculated at 1 atm and at the
801 < experimentally calculated densities for super-cooled and liquid
802 < water. In the upper plot, the diffusion constant for SSD/E is
803 < consistently a little faster than experiment, while SSD starts off
804 < slower than experiment and crosses to merge with SSD/E at high
805 < temperatures. Both models follow the experimental trend well, but
739 < diffuse too rapidly at higher temperatures. This abnormally fast
740 < diffusion is caused by the decreased system density. Since the
741 < densities of SSD/E don't deviate as much from experiment as those of
742 < SSD, it follows the experimental trend more closely. This observation
743 < is backed up by looking at the lower plot. The diffusion constants for
744 < SSD/E track with the experimental values while SSD deviates on the low
745 < side of the trend with increasing temperature. This is again a product
746 < of SSD/E having densities closer to experiment, and not deviating to
747 < lower densities with increasing temperature as rapidly.
748 <
749 < \begin{figure}
750 < \includegraphics[width=85mm]{ssdediffuse.epsi}
751 < \caption{Plots of the diffusion constants calculated from SSD/E and SSD,
752 < both without a reaction field along with experimental results from
753 < Gillen \emph{et al.}\cite{Gillen72} and Mills\cite{Mills73}. The
754 < upper plot is at densities calculated from the NPT simulations at a
755 < pressure of 1 atm, while the lower plot is at the experimentally
756 < calculated densities.}
794 > \begin{figure}
795 > \begin{center}
796 > \epsfxsize=6in
797 > \epsfbox{ssdeDiffuse.epsi}
798 > \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   \label{ssdediffuse}
807 + \end{center}
808   \end{figure}
809  
810 < \begin{figure}
811 < \includegraphics[width=85mm]{ssdrfdiffuse.epsi}
812 < \caption{Plots of the diffusion constants calculated from SSD/RF and SSD,
813 < both with an active reaction field along with experimental results
814 < from Gillen \emph{et al.}\cite{Gillen72} and Mills\cite{Mills73}. The
815 < upper plot is at densities calculated from the NPT simulations at a
816 < pressure of 1 atm, while the lower plot is at the experimentally
817 < calculated densities.}
810 > The reparameterization of the {\sc ssd} water model, both for use with and
811 > without an applied long-range correction, brought the densities up to
812 > what is expected for simulating liquid water. In addition to improving
813 > 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 > without an active reaction field at the densities calculated from
817 > their respective NPT simulations at 1 atm. The diffusion constant for
818 > {\sc ssd/e} is consistently higher than experiment, while {\sc ssd1} remains lower
819 > 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 > in {\sc ssd1}'s crossing above 360 K.  This increasing diffusion relative to
823 > the experimental values is caused by the rapidly decreasing system
824 > density with increasing temperature.  Both {\sc ssd1} and {\sc ssd/e} show this
825 > deviation in particle mobility, but this trend has different
826 > implications on the diffusive behavior of the models.  While {\sc ssd1}
827 > shows more experimentally accurate diffusive behavior in the high
828 > temperature regimes, {\sc ssd/e} shows more accurate behavior in the
829 > 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 > of {\sc ssd/e} relative to {\sc ssd1} under the most commonly simulated
833 > conditions.
834 >
835 > \begin{figure}
836 > \begin{center}
837 > \epsfxsize=6in
838 > \epsfbox{ssdrfDiffuse.epsi}
839 > \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   \label{ssdrfdiffuse}
849 + \end{center}
850   \end{figure}
851  
852 < In figure \ref{ssdrfdiffuse}, the diffusion constants for SSD/RF are
853 < compared with SSD with an active reaction field. In the upper plot,
854 < SSD/RF tracks with the experimental results incredibly well, identical
855 < within error throughout the temperature range and only showing a
856 < slight increasing trend at higher temperatures. SSD also tracks
857 < experiment well, only it tends to diffuse a little more slowly at low
858 < temperatures and deviates to diffuse too rapidly at high
859 < temperatures. As was stated in the SSD/E comparisons, this deviation
860 < away from the ideal trend is due to a rapid decrease in density at
861 < higher temperatures. SSD/RF doesn't suffer from this problem as much
862 < as SSD, because the calculated densities are more true to
863 < experiment. This is again emphasized in the lower plot, where SSD/RF
783 < tracks the experimental diffusion exactly while SSD's diffusion
784 < constants are slightly too low due to its need for a lower density at
785 < the specified temperature.
852 > 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 > tracks the experimental results quantitatively, identical within error
855 > throughout most of the temperature range shown and exhibiting only a
856 > slight increasing trend at higher temperatures. {\sc ssd1} tends to diffuse
857 > more slowly at low temperatures and deviates to diffuse too rapidly at
858 > 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 > temperatures. {\sc ssd/rf} does not suffer from this problem as much as {\sc ssd1}
861 > 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  
865 < \subsection{Additional Observations}
865 > \begin{table}
866 > \begin{minipage}{\linewidth}
867 > \renewcommand{\thefootnote}{\thempfootnote}
868 > \begin{center}
869 > \caption{Properties of the single-point water models compared with
870 > experimental data at ambient conditions}
871 > \begin{tabular}{ l  c  c  c  c  c }
872 > \hline \\[-3mm]
873 > \ \ \ \ \ \  & \ \ \ {\sc ssd1} \ \ \ & \ {\sc ssd/e} \ \ \ & \ {\sc ssd1} (RF) \ \
874 > \ & \ {\sc ssd/rf} \ \ \ & \ Expt. \\
875 > \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 > \ \ \ $D$ ($10^{-5}$ 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$0.2 & 2.3\footnote{Calculated for 298 K from data in
890 > Ref. \citen{Krynicki66}}
891 > \end{tabular}
892 > \label{liquidproperties}
893 > \end{center}
894 > \end{minipage}
895 > \end{table}
896  
897 < While performing the melting sequences of SSD/E, some interesting
898 < observations were made. After melting at 230 K, two of the systems
899 < underwent crystallization events near 245 K. As the heating process
900 < continued, the two systems remained crystalline until finally melting
901 < between 320 and 330 K. These simulations were excluded from the data
902 < set shown in figure \ref{ssdedense} and replaced with two additional
903 < melting sequences that did not undergo this anomalous phase
904 < transition, while this crystallization event was investigated
905 < separately.
897 > Table \ref{liquidproperties} gives a synopsis of the liquid state
898 > properties of the water models compared in this study along with the
899 > experimental values for liquid water at ambient conditions. The
900 > coordination number ($n_C$) and number of hydrogen bonds per particle
901 > ($n_H$) were calculated by integrating the following relations:
902 > \begin{equation}
903 > n_C = 4\pi\rho_{\text{OO}}\int_{0}^{a}r^2\text{g}_{\text{OO}}(r)dr,
904 > \end{equation}
905 > \begin{equation}
906 > n_H = 4\pi\rho_{\text{OH}}\int_{0}^{b}r^2\text{g}_{\text{OH}}(r)dr,
907 > \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 > 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  
922 < \begin{figure}
923 < \includegraphics[width=85mm]{povIce.ps}
924 < \caption{Crystal structure of an ice 0 lattice shown from the (001) face.}
925 < \label{weirdice}
926 < \end{figure}
922 > The time constants for the orientational autocorrelation functions
923 > are also displayed in Table \ref{liquidproperties}. The dipolar
924 > orientational time correlation functions ($C_{l}$) are described
925 > by:
926 > \begin{equation}
927 > C_{l}(t) = \langle P_l[\hat{\mathbf{u}}_j(0)\cdot\hat{\mathbf{u}}_j(t)]\rangle,
928 > \end{equation}
929 > 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 > \ref{liquidproperties} was calculated from the same NMR data in the
943 > 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 > and without an active reaction field. Turning on the reaction field
947 > 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  
953 < The final configurations of these two melting sequences shows an
806 < expanded zeolite-like crystal structure that does not correspond to
807 < any known form of ice. For convenience and to help distinguish it from
808 < the experimentally observed forms of ice, this crystal structure will
809 < henceforth be referred to as ice-zero (ice 0). The crystallinity was
810 < extensive enough than a near ideal crystal structure could be
811 < obtained. Figure \ref{weirdice} shows the repeating crystal structure
812 < of a typical crystal at 5 K. The unit cell contains eight molecules,
813 < and figure \ref{unitcell} shows a unit cell built from the water
814 < particle center of masses that can be used to construct a repeating
815 < lattice, similar to figure \ref{weirdice}. Each molecule is hydrogen
816 < bonded to four other water molecules; however, the hydrogen bonds are
817 < flexed rather than perfectly straight. This results in a skewed
818 < tetrahedral geometry about the central molecule. Looking back at
819 < figure \ref{isosurface}, it is easy to see how these flexed hydrogen
820 < bonds are allowed in that the attractive regions are conical in shape,
821 < with the greatest attraction in the central region. Though not ideal,
822 < these flexed hydrogen bonds are favorable enough to stabilize an
823 < entire crystal generated around them. In fact, the imperfect ice 0
824 < crystals were so stable that they melted at greater than room
825 < temperature.
953 > \subsection{Additional Observations}
954  
955   \begin{figure}
956 < \includegraphics[width=65mm]{ice0cell.eps}
957 < \caption{Simple unit cell for constructing ice 0. In this cell, $c$ is
958 < equal to $0.4714\times a$, and a typical value for $a$ is 8.25 \AA.}
959 < \label{unitcell}
956 > \begin{center}
957 > \epsfxsize=6in
958 > \epsfbox{icei_bw.eps}
959 > \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 > \label{weirdice}
964 > \end{center}
965   \end{figure}
966  
967 < The initial simulations indicated that ice 0 is the preferred ice
968 < structure for at least SSD/E. To verify this, a comparison was made
969 < between near ideal crystals of ice $I_h$, ice $I_c$, and ice 0 at
970 < constant pressure with SSD/E, SSD/RF, and SSD. Near ideal versions of
971 < the three types of crystals were cooled to ~1 K, and the potential
972 < energies of each were compared using all three water models. With
973 < every water model, ice 0 turned out to have the lowest potential
974 < energy: 5\% lower than $I_h$ with SSD, 6.5\% lower with SSD/E, and
975 < 7.5\% lower with SSD/RF. In all three of these water models, ice $I_c$
976 < was observed to be ~2\% less stable than ice $I_h$. In addition to
977 < having the lowest potential energy, ice 0 was the most expanded of the
978 < three ice crystals, ~5\% less dense than ice $I_h$ with all of the
979 < water models. In all three water models, ice $I_c$ was observed to be
980 < ~2\% more dense than ice $I_h$.
967 > While performing a series of melting simulations on an early iteration
968 > of {\sc ssd/e} not discussed in this paper, we observed recrystallization
969 > 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 > Ice-$\sqrt{\smash[b]{-\text{I}}}$ (Ice-{\it i}).  A large enough
977 > portion of the sample crystallized that we have been able to obtain a
978 > near ideal crystal structure of Ice-{\it i}. Figure \ref{weirdice}
979 > 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 > configuration. Though not ideal, these flexed hydrogen bonds are
987 > favorable enough to stabilize an entire crystal generated around them.
988  
989 < In addition to the low temperature comparisons, melting sequences were
990 < performed with ice 0 as the initial configuration using SSD/E, SSD/RF,
991 < and SSD both with and without a reaction field. The melting
992 < transitions for both SSD/E and SSD without a reaction field occurred
993 < at temperature in excess of 375 K. SSD/RF and SSD with a reaction
994 < field had more reasonable melting transitions, down near 325 K. These
995 < melting point observations emphasize how preferred this crystal
996 < structure is over the most common types of ice when using these single
997 < point water models.
989 > 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  
1004 < Recognizing that the above tests show ice 0 to be both the most stable
1005 < and lowest density crystal structure for these single point water
1006 < models, it is interesting to speculate on the favorability of this
1007 < crystal structure with the different charge based models. As a quick
1008 < test, these 3 crystal types were converted from SSD type particles to
1009 < TIP3P waters and read into CHARMM.\cite{Karplus83} Identical energy
1010 < minimizations were performed on all of these crystals to compare the
1011 < system energies. Again, ice 0 was observed to have the lowest total
1012 < system energy. The total energy of ice 0 was ~2\% lower than ice
1013 < $I_h$, which was in turn ~3\% lower than ice $I_c$. From these initial
1014 < results, we would not be surprised if results from the other common
1015 < water models show ice 0 to be the lowest energy crystal structure. A
1016 < continuation on work studing ice 0 with multipoint water models will
1017 < be published in a coming article.
1004 > \begin{table}
1005 > \begin{center}
1006 > \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 > \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 > \ \ \ {\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 > \end{tabular}
1018 > \label{iceenthalpy}
1019 > \end{center}
1020 > \end{table}
1021  
1022 + In addition to these energetic comparisons, melting simulations were
1023 + 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 + showed more reasonable melting transitions near 325~K.  These melting
1028 + point observations clearly show that all of the {\sc ssd}-derived models
1029 + prefer the ice-{\it i} structure.
1030 +
1031   \section{Conclusions}
1032 < The density maximum and temperature dependent transport for the SSD
1033 < water model, both with and without the use of reaction field, were
1034 < studied via a series of NPT and NVE simulations. The constant pressure
1035 < simulations of the melting of both $I_h$ and $I_c$ ice showed a
1036 < density maximum near 260 K. In most cases, the calculated densities
1037 < were significantly lower than the densities calculated in simulations
1038 < of other water models. Analysis of particle diffusion showed SSD to
1039 < capture the transport properties of experimental very well in both the
1040 < normal and super-cooled liquid regimes. In order to correct the
1041 < density behavior, SSD was reparameterized for use both with and
1042 < without a long-range interaction correction, SSD/RF and SSD/E
1043 < respectively. In addition to correcting the abnormally low densities,
1044 < these new versions were show to maintain or improve upon the transport
1045 < and structural features of the original water model, all while
1046 < maintaining the fast performance of the SSD water model. This work
1047 < shows these simple water models, and in particular SSD/E and SSD/RF,
1048 < to be excellent choices to represent explicit water in future
1032 >
1033 > The density maximum and temperature dependence of the self-diffusion
1034 > constant were studied for the {\sc ssd} water model, both with and without
1035 > 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 > models (and experiment). Analysis of self-diffusion showed {\sc ssd} to
1040 > capture the transport properties of water well in both the liquid and
1041 > supercooled liquid regimes.
1042 >
1043 > 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 > 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 + by the {\sc ssd} family of water models is somewhat troubling, since liquid
1056 + simulations on this family of water models at room temperature are
1057 + effectively simulations of supercooled or metastable liquids.  One
1058 + way to destabilize this unphysical ice structure would be to make the
1059 + 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 + Additionally, our initial calculations show that the Ice-{\it i}
1065 + 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   \section{Acknowledgments}
1073 < The authors would like to thank the National Science Foundation for
1074 < funding under grant CHE-0134881. Computation time was provided by the
1075 < Notre Dame Bunch-of-Boxes (B.o.B) computer cluster under NSF grant DMR
1076 < 00 79647.
1073 > 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 > DMR-0079647.
1077  
1078 < \bibliographystyle{jcp}
1078 > \newpage
1079  
1080 + \bibliographystyle{jcp}
1081   \bibliography{nptSSD}
1082  
1083   %\pagebreak

Diff Legend

Removed lines
+ Added lines
< Changed lines
> Changed lines