| 1 | chrisfen | 861 | %\documentclass[prb,aps,times,twocolumn,tabularx]{revtex4} | 
| 2 | gezelter | 1036 | %\documentclass[preprint,aps,endfloat]{revtex4} | 
| 3 |  |  | \documentclass[11pt]{article} | 
| 4 |  |  | \usepackage{endfloat} | 
| 5 | chrisfen | 743 | \usepackage{amsmath} | 
| 6 | chrisfen | 862 | \usepackage{epsf} | 
| 7 |  |  | \usepackage{berkeley} | 
| 8 |  |  | \usepackage{setspace} | 
| 9 |  |  | \usepackage{tabularx} | 
| 10 | chrisfen | 743 | \usepackage{graphicx} | 
| 11 | gezelter | 1036 | \usepackage[ref]{overcite} | 
| 12 |  |  | \pagestyle{plain} | 
| 13 |  |  | \pagenumbering{arabic} | 
| 14 |  |  | \oddsidemargin 0.0cm \evensidemargin 0.0cm | 
| 15 |  |  | \topmargin -21pt \headsep 10pt | 
| 16 |  |  | \textheight 9.0in \textwidth 6.5in | 
| 17 |  |  | \brokenpenalty=10000 | 
| 18 |  |  |  | 
| 19 | chrisfen | 1033 | %\renewcommand\citemid{\ } % no comma in optional reference note | 
| 20 | chrisfen | 743 |  | 
| 21 |  |  | \begin{document} | 
| 22 |  |  |  | 
| 23 | gezelter | 921 | \title{On the structural and transport properties of the soft sticky | 
| 24 | chrisfen | 1030 | dipole (SSD) and related single point water models} | 
| 25 | chrisfen | 743 |  | 
| 26 | gezelter | 1036 | \author{Christopher J. Fennell and J. Daniel | 
| 27 |  |  | Gezelter\footnote{Corresponding author. \ Electronic mail: | 
| 28 |  |  | gezelter@nd.edu} \\ Department of Chemistry and Biochemistry\\ University of Notre Dame\\ | 
| 29 | chrisfen | 743 | Notre Dame, Indiana 46556} | 
| 30 |  |  |  | 
| 31 |  |  | \date{\today} | 
| 32 |  |  |  | 
| 33 | gezelter | 1036 | \maketitle | 
| 34 |  |  | \doublespacing | 
| 35 | chrisfen | 862 |  | 
| 36 | chrisfen | 743 | \begin{abstract} | 
| 37 | gezelter | 921 | The density maximum and temperature dependence of the self-diffusion | 
| 38 | chrisfen | 1030 | constant were investigated for the soft sticky dipole (SSD) water | 
| 39 | chrisfen | 1033 | model and two related reparameterizations of this single-point model. | 
| 40 | gezelter | 921 | A combination of microcanonical and isobaric-isothermal molecular | 
| 41 |  |  | dynamics simulations were used to calculate these properties, both | 
| 42 |  |  | with and without the use of reaction field to handle long-range | 
| 43 |  |  | electrostatics.  The isobaric-isothermal (NPT) simulations of the | 
| 44 |  |  | melting of both ice-$I_h$ and ice-$I_c$ showed a density maximum near | 
| 45 | chrisfen | 1033 | 260~K.  In most cases, the use of the reaction field resulted in | 
| 46 | gezelter | 921 | calculated densities which were were significantly lower than | 
| 47 |  |  | experimental densities.  Analysis of self-diffusion constants shows | 
| 48 | chrisfen | 1030 | that the original SSD model captures the transport properties of | 
| 49 | chrisfen | 1040 | experimental water very well in both the normal and supercooled | 
| 50 | chrisfen | 1033 | liquid regimes.  We also present our reparameterized versions of SSD | 
| 51 | gezelter | 921 | for use both with the reaction field or without any long-range | 
| 52 | chrisfen | 1030 | electrostatic corrections.  These are called the SSD/RF and SSD/E | 
| 53 | gezelter | 921 | models respectively.  These modified models were shown to maintain or | 
| 54 |  |  | improve upon the experimental agreement with the structural and | 
| 55 | chrisfen | 1030 | transport properties that can be obtained with either the original SSD | 
| 56 |  |  | or the density corrected version of the original model (SSD1). | 
| 57 | gezelter | 921 | Additionally, a novel low-density ice structure is presented | 
| 58 | chrisfen | 1030 | which appears to be the most stable ice structure for the entire SSD | 
| 59 | gezelter | 921 | family. | 
| 60 | chrisfen | 743 | \end{abstract} | 
| 61 |  |  |  | 
| 62 | chrisfen | 862 | \newpage | 
| 63 | chrisfen | 743 |  | 
| 64 |  |  | %\narrowtext | 
| 65 |  |  |  | 
| 66 |  |  | %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% | 
| 67 |  |  | %                              BODY OF TEXT | 
| 68 |  |  | %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% | 
| 69 |  |  |  | 
| 70 |  |  | \section{Introduction} | 
| 71 |  |  |  | 
| 72 | chrisfen | 862 | One of the most important tasks in the simulation of biochemical | 
| 73 | gezelter | 921 | systems is the proper depiction of the aqueous environment of the | 
| 74 |  |  | molecules of interest.  In some cases (such as in the simulation of | 
| 75 |  |  | phospholipid bilayers), the majority of the calculations that are | 
| 76 |  |  | performed involve interactions with or between solvent molecules. | 
| 77 |  |  | Thus, the properties one may observe in biochemical simulations are | 
| 78 |  |  | going to be highly dependent on the physical properties of the water | 
| 79 |  |  | model that is chosen. | 
| 80 | chrisfen | 743 |  | 
| 81 | gezelter | 921 | There is an especially delicate balance between computational | 
| 82 |  |  | efficiency and the ability of the water model to accurately predict | 
| 83 |  |  | the properties of bulk | 
| 84 |  |  | water.\cite{Jorgensen83,Berendsen87,Jorgensen00} For example, the | 
| 85 |  |  | TIP5P model improves on the structural and transport properties of | 
| 86 |  |  | water relative to the previous TIP models, yet this comes at a greater | 
| 87 |  |  | than 50\% increase in computational | 
| 88 |  |  | cost.\cite{Jorgensen01,Jorgensen00} | 
| 89 |  |  |  | 
| 90 |  |  | One recently developed model that largely succeeds in retaining the | 
| 91 |  |  | accuracy of bulk properties while greatly reducing the computational | 
| 92 | chrisfen | 1030 | cost is the Soft Sticky Dipole (SSD) water | 
| 93 |  |  | model.\cite{Ichiye96,Ichiye96b,Ichiye99,Ichiye03} The SSD model | 
| 94 |  |  | was developed by Ichiye \emph{et al.} as a modified form of the | 
| 95 | gezelter | 921 | hard-sphere water model proposed by Bratko, Blum, and | 
| 96 | chrisfen | 1030 | Luzar.\cite{Bratko85,Bratko95} SSD is a {\it single point} model | 
| 97 | chrisfen | 1033 | which has an interaction site that is both a point dipole and a | 
| 98 | gezelter | 921 | 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 | chrisfen | 1030 | the proper hydrogen bond orientation in the first solvation shell. | 
| 103 | gezelter | 921 |  | 
| 104 | chrisfen | 1030 | The interaction between two SSD water molecules \emph{i} and \emph{j} | 
| 105 | gezelter | 921 | is given by the potential | 
| 106 | chrisfen | 743 | \begin{equation} | 
| 107 |  |  | u_{ij} = u_{ij}^{LJ} (r_{ij})\ + u_{ij}^{dp} | 
| 108 | gezelter | 921 | ({\bf r}_{ij},{\bf \Omega}_i,{\bf \Omega}_j)\ + | 
| 109 | chrisfen | 743 | u_{ij}^{sp} | 
| 110 | gezelter | 921 | ({\bf r}_{ij},{\bf \Omega}_i,{\bf \Omega}_j), | 
| 111 | chrisfen | 743 | \end{equation} | 
| 112 | gezelter | 921 | where the ${\bf r}_{ij}$ is the position vector between molecules | 
| 113 |  |  | \emph{i} and \emph{j} with magnitude $r_{ij}$, and | 
| 114 |  |  | ${\bf \Omega}_i$ and ${\bf \Omega}_j$ represent the orientations of | 
| 115 |  |  | the two molecules. The Lennard-Jones and dipole interactions are given | 
| 116 |  |  | by the following familiar forms: | 
| 117 | chrisfen | 743 | \begin{equation} | 
| 118 | gezelter | 921 | u_{ij}^{LJ}(r_{ij}) = 4\epsilon | 
| 119 |  |  | \left[\left(\frac{\sigma}{r_{ij}}\right)^{12}-\left(\frac{\sigma}{r_{ij}}\right)^{6}\right] | 
| 120 |  |  | \ , | 
| 121 | chrisfen | 743 | \end{equation} | 
| 122 | gezelter | 921 | and | 
| 123 | chrisfen | 743 | \begin{equation} | 
| 124 | gezelter | 921 | u_{ij}^{dp} = \frac{|\mu_i||\mu_j|}{4 \pi \epsilon_0 r_{ij}^3} \left( | 
| 125 |  |  | \hat{\bf u}_i \cdot \hat{\bf u}_j - 3(\hat{\bf u}_i\cdot\hat{\bf | 
| 126 |  |  | r}_{ij})(\hat{\bf u}_j\cdot\hat{\bf r}_{ij}) \right)\ , | 
| 127 | chrisfen | 743 | \end{equation} | 
| 128 | gezelter | 921 | where $\hat{\bf u}_i$ and $\hat{\bf u}_j$ are the unit vectors along | 
| 129 |  |  | the dipoles of molecules $i$ and $j$ respectively. $|\mu_i|$ and | 
| 130 |  |  | $|\mu_j|$ are the strengths of the dipole moments, and $\hat{\bf | 
| 131 |  |  | r}_{ij}$ is the unit vector pointing from molecule $j$ to molecule | 
| 132 |  |  | $i$. | 
| 133 |  |  |  | 
| 134 |  |  | The sticky potential is somewhat less familiar: | 
| 135 | chrisfen | 743 | \begin{equation} | 
| 136 |  |  | u_{ij}^{sp} | 
| 137 | gezelter | 921 | ({\bf r}_{ij},{\bf \Omega}_i,{\bf \Omega}_j) = | 
| 138 |  |  | \frac{\nu_0}{2}[s(r_{ij})w({\bf r}_{ij},{\bf \Omega}_i,{\bf \Omega}_j) | 
| 139 |  |  | + s^\prime(r_{ij})w^\prime({\bf r}_{ij},{\bf \Omega}_i,{\bf | 
| 140 |  |  | \Omega}_j)]\ . | 
| 141 | chrisfen | 1017 | \label{stickyfunction} | 
| 142 | chrisfen | 743 | \end{equation} | 
| 143 | gezelter | 921 | Here, $\nu_0$ is a strength parameter for the sticky potential, and | 
| 144 |  |  | $s$ and $s^\prime$ are cubic switching functions which turn off the | 
| 145 |  |  | sticky interaction beyond the first solvation shell. The $w$ function | 
| 146 |  |  | can be thought of as an attractive potential with tetrahedral | 
| 147 |  |  | geometry: | 
| 148 | chrisfen | 743 | \begin{equation} | 
| 149 | gezelter | 921 | w({\bf r}_{ij},{\bf \Omega}_i,{\bf \Omega}_j)=\sin\theta_{ij}\sin2\theta_{ij}\cos2\phi_{ij}, | 
| 150 | chrisfen | 743 | \end{equation} | 
| 151 | gezelter | 921 | while the $w^\prime$ function counters the normal aligned and | 
| 152 |  |  | anti-aligned structures favored by point dipoles: | 
| 153 | chrisfen | 743 | \begin{equation} | 
| 154 | chrisfen | 1017 | w^\prime({\bf r}_{ij},{\bf \Omega}_i,{\bf \Omega}_j) = (\cos\theta_{ij}-0.6)^2(\cos\theta_{ij}+0.8)^2-w^\circ, | 
| 155 | chrisfen | 743 | \end{equation} | 
| 156 | gezelter | 921 | It should be noted that $w$ is proportional to the sum of the $Y_3^2$ | 
| 157 |  |  | and $Y_3^{-2}$ spherical harmonics (a linear combination which | 
| 158 |  |  | enhances the tetrahedral geometry for hydrogen bonded structures), | 
| 159 |  |  | while $w^\prime$ is a purely empirical function.  A more detailed | 
| 160 |  |  | description of the functional parts and variables in this potential | 
| 161 | chrisfen | 1030 | can be found in the original SSD | 
| 162 | gezelter | 921 | articles.\cite{Ichiye96,Ichiye96b,Ichiye99,Ichiye03} | 
| 163 | chrisfen | 743 |  | 
| 164 | chrisfen | 1030 | Since SSD is a single-point {\it dipolar} model, the force | 
| 165 | gezelter | 921 | calculations are simplified significantly relative to the standard | 
| 166 |  |  | {\it charged} multi-point models. In the original Monte Carlo | 
| 167 | chrisfen | 1033 | simulations using this model, Liu and Ichiye reported that using SSD | 
| 168 |  |  | decreased computer time by a factor of 6-7 compared to other | 
| 169 | gezelter | 921 | models.\cite{Ichiye96} What is most impressive is that this savings | 
| 170 |  |  | did not come at the expense of accurate depiction of the liquid state | 
| 171 | chrisfen | 1033 | properties.  Indeed, SSD maintains reasonable agreement with the Soper | 
| 172 |  |  | data for the structural features of liquid | 
| 173 | gezelter | 921 | water.\cite{Soper86,Ichiye96} Additionally, the dynamical properties | 
| 174 | chrisfen | 1030 | exhibited by SSD agree with experiment better than those of more | 
| 175 | gezelter | 921 | computationally expensive models (like TIP3P and | 
| 176 |  |  | SPC/E).\cite{Ichiye99} The combination of speed and accurate depiction | 
| 177 | chrisfen | 1030 | of solvent properties makes SSD a very attractive model for the | 
| 178 | gezelter | 921 | simulation of large scale biochemical simulations. | 
| 179 | chrisfen | 743 |  | 
| 180 | chrisfen | 1030 | One feature of the SSD model is that it was parameterized for | 
| 181 |  |  | use with the Ewald sum to handle long-range interactions.  This would | 
| 182 |  |  | normally be the best way of handling long-range interactions in | 
| 183 |  |  | systems that contain other point charges.  However, our group has | 
| 184 |  |  | recently become interested in systems with point dipoles as mimics for | 
| 185 |  |  | neutral, but polarized regions on molecules (e.g. the zwitterionic | 
| 186 |  |  | head group regions of phospholipids).  If the system of interest does | 
| 187 |  |  | not contain point charges, the Ewald sum and even particle-mesh Ewald | 
| 188 |  |  | become computational bottlenecks.  Their respective ideal | 
| 189 |  |  | $N^\frac{3}{2}$ and $N\log N$ calculation scaling orders for $N$ | 
| 190 |  |  | particles can become prohibitive when $N$ becomes | 
| 191 |  |  | large.\cite{Darden99} In applying this water model in these types of | 
| 192 |  |  | systems, it would be useful to know its properties and behavior under | 
| 193 |  |  | the more computationally efficient reaction field (RF) technique, or | 
| 194 |  |  | even with a simple cutoff. This study addresses these issues by | 
| 195 |  |  | looking at the structural and transport behavior of SSD over a | 
| 196 |  |  | variety of temperatures with the purpose of utilizing the RF | 
| 197 |  |  | correction technique.  We then suggest modifications to the parameters | 
| 198 |  |  | that result in more realistic bulk phase behavior.  It should be noted | 
| 199 |  |  | that in a recent publication, some of the original investigators of | 
| 200 |  |  | the SSD water model have suggested adjustments to the SSD | 
| 201 |  |  | water model to address abnormal density behavior (also observed here), | 
| 202 |  |  | calling the corrected model SSD1.\cite{Ichiye03} In what | 
| 203 |  |  | follows, we compare our reparamaterization of SSD with both the | 
| 204 |  |  | original SSD and SSD1 models with the goal of improving | 
| 205 |  |  | the bulk phase behavior of an SSD-derived model in simulations | 
| 206 | chrisfen | 1033 | utilizing the reaction field. | 
| 207 | chrisfen | 757 |  | 
| 208 | chrisfen | 743 | \section{Methods} | 
| 209 |  |  |  | 
| 210 | gezelter | 921 | Long-range dipole-dipole interactions were accounted for in this study | 
| 211 | chrisfen | 1033 | by using either the reaction field technique or by resorting to a | 
| 212 |  |  | simple cubic switching function at a cutoff radius.  One of the early | 
| 213 |  |  | applications of a reaction field was actually in Monte Carlo | 
| 214 |  |  | simulations of liquid water.\cite{Barker73} Under this method, the | 
| 215 |  |  | magnitude of the reaction field acting on dipole $i$ is | 
| 216 | chrisfen | 743 | \begin{equation} | 
| 217 |  |  | \mathcal{E}_{i} = \frac{2(\varepsilon_{s} - 1)}{2\varepsilon_{s} + 1} | 
| 218 | gezelter | 1029 | \frac{1}{r_{c}^{3}} \sum_{j\in{\mathcal{R}}} {\bf \mu}_{j} s(r_{ij}), | 
| 219 | chrisfen | 743 | \label{rfequation} | 
| 220 |  |  | \end{equation} | 
| 221 |  |  | where $\mathcal{R}$ is the cavity defined by the cutoff radius | 
| 222 |  |  | ($r_{c}$), $\varepsilon_{s}$ is the dielectric constant imposed on the | 
| 223 | gezelter | 921 | system (80 in the case of liquid water), ${\bf \mu}_{j}$ is the dipole | 
| 224 | gezelter | 1029 | moment vector of particle $j$, and $s(r_{ij})$ is a cubic switching | 
| 225 | chrisfen | 743 | function.\cite{AllenTildesley} The reaction field contribution to the | 
| 226 | gezelter | 921 | total energy by particle $i$ is given by $-\frac{1}{2}{\bf | 
| 227 |  |  | \mu}_{i}\cdot\mathcal{E}_{i}$ and the torque on dipole $i$ by ${\bf | 
| 228 |  |  | \mu}_{i}\times\mathcal{E}_{i}$.\cite{AllenTildesley}  Use of the reaction | 
| 229 | chrisfen | 1033 | field is known to alter the bulk orientational properties of simulated | 
| 230 |  |  | water, and there is particular sensitivity of these properties on | 
| 231 |  |  | changes in the length of the cutoff radius.\cite{Berendsen98} This | 
| 232 |  |  | variable behavior makes reaction field a less attractive method than | 
| 233 |  |  | the Ewald sum.  However, for very large systems, the computational | 
| 234 |  |  | benefit of reaction field is dramatic. | 
| 235 | gezelter | 921 |  | 
| 236 |  |  | We have also performed a companion set of simulations {\it without} a | 
| 237 |  |  | surrounding dielectric (i.e. using a simple cubic switching function | 
| 238 | chrisfen | 1022 | at the cutoff radius), and as a result we have two reparamaterizations | 
| 239 | chrisfen | 1030 | of SSD which could be used either with or without the reaction | 
| 240 |  |  | field turned on. | 
| 241 | chrisfen | 777 |  | 
| 242 | gezelter | 1029 | Simulations to obtain the preferred densities of the models were | 
| 243 |  |  | performed in the isobaric-isothermal (NPT) ensemble, while all | 
| 244 |  |  | dynamical properties were obtained from microcanonical (NVE) | 
| 245 |  |  | simulations done at densities matching the NPT density for a | 
| 246 |  |  | particular target temperature.  The constant pressure simulations were | 
| 247 |  |  | implemented using an integral thermostat and barostat as outlined by | 
| 248 |  |  | Hoover.\cite{Hoover85,Hoover86} All molecules were treated as | 
| 249 |  |  | non-linear rigid bodies. Vibrational constraints are not necessary in | 
| 250 | chrisfen | 1030 | simulations of SSD, because there are no explicit hydrogen | 
| 251 |  |  | atoms, and thus no molecular vibrational modes need to be considered. | 
| 252 | chrisfen | 743 |  | 
| 253 |  |  | Integration of the equations of motion was carried out using the | 
| 254 | chrisfen | 1027 | symplectic splitting method proposed by Dullweber, Leimkuhler, and | 
| 255 | chrisfen | 1030 | McLachlan ({\sc dlm}).\cite{Dullweber1997} Our reason for selecting | 
| 256 |  |  | this integrator centers on poor energy conservation of rigid body | 
| 257 |  |  | dynamics using traditional quaternion | 
| 258 |  |  | integration.\cite{Evans77,Evans77b} In typical microcanonical ensemble | 
| 259 |  |  | simulations, the energy drift when using quaternions was substantially | 
| 260 |  |  | greater than when using the {\sc dlm} method (fig. \ref{timestep}). | 
| 261 |  |  | This steady drift in the total energy has also been observed by Kol | 
| 262 |  |  | {\it et al.}\cite{Laird97} | 
| 263 | chrisfen | 743 |  | 
| 264 |  |  | The key difference in the integration method proposed by Dullweber | 
| 265 |  |  | \emph{et al.} is that the entire rotation matrix is propagated from | 
| 266 | gezelter | 921 | one time step to the next.  The additional memory required by the | 
| 267 |  |  | algorithm is inconsequential on modern computers, and translating the | 
| 268 |  |  | rotation matrix into quaternions for storage purposes makes trajectory | 
| 269 |  |  | data quite compact. | 
| 270 | chrisfen | 743 |  | 
| 271 | gezelter | 1029 | The {\sc dlm} method allows for Verlet style integration of both | 
| 272 | chrisfen | 1027 | translational and orientational motion of rigid bodies. In this | 
| 273 | gezelter | 921 | integration method, the orientational propagation involves a sequence | 
| 274 |  |  | of matrix evaluations to update the rotation | 
| 275 |  |  | matrix.\cite{Dullweber1997} These matrix rotations are more costly | 
| 276 |  |  | than the simpler arithmetic quaternion propagation. With the same time | 
| 277 | chrisfen | 1030 | step, a 1000 SSD particle simulation shows an average 7\% | 
| 278 |  |  | increase in computation time using the {\sc dlm} method in place of | 
| 279 |  |  | quaternions. The additional expense per step is justified when one | 
| 280 |  |  | considers the ability to use time steps that are nearly twice as large | 
| 281 |  |  | under {\sc dlm} than would be usable under quaternion dynamics.  The | 
| 282 |  |  | energy conservation of the two methods using a number of different | 
| 283 |  |  | time steps is illustrated in figure | 
| 284 | gezelter | 921 | \ref{timestep}. | 
| 285 | chrisfen | 743 |  | 
| 286 | gezelter | 1036 | \begin{figure} | 
| 287 |  |  | \begin{center} | 
| 288 |  |  | \epsfxsize=6in | 
| 289 |  |  | \epsfbox{timeStep.epsi} | 
| 290 |  |  | \caption{Energy conservation using both quaternion-based integration and the | 
| 291 |  |  | {\sc dlm} method with increasing time step. The larger time step plots | 
| 292 |  |  | are shifted from the true energy baseline (that of $\Delta t$ = | 
| 293 |  |  | 0.1~fs) for clarity.} | 
| 294 |  |  | \label{timestep} | 
| 295 |  |  | \end{center} | 
| 296 |  |  | \end{figure} | 
| 297 | chrisfen | 743 |  | 
| 298 |  |  | In figure \ref{timestep}, the resulting energy drift at various time | 
| 299 | chrisfen | 1030 | steps for both the {\sc dlm} and quaternion integration schemes is | 
| 300 |  |  | compared.  All of the 1000 SSD particle simulations started with | 
| 301 |  |  | the same configuration, and the only difference was the method used to | 
| 302 | chrisfen | 1033 | handle orientational motion. At time steps of 0.1 and 0.5~fs, both | 
| 303 | chrisfen | 1030 | methods for propagating the orientational degrees of freedom conserve | 
| 304 |  |  | energy fairly well, with the quaternion method showing a slight energy | 
| 305 | chrisfen | 1033 | drift over time in the 0.5~fs time step simulation. At time steps of 1 | 
| 306 |  |  | and 2~fs, the energy conservation benefits of the {\sc dlm} method are | 
| 307 | chrisfen | 1030 | clearly demonstrated. Thus, while maintaining the same degree of | 
| 308 |  |  | energy conservation, one can take considerably longer time steps, | 
| 309 |  |  | leading to an overall reduction in computation time. | 
| 310 | chrisfen | 743 |  | 
| 311 | chrisfen | 1030 | Energy drift in the simulations using {\sc dlm} integration was | 
| 312 | chrisfen | 1033 | unnoticeable for time steps up to 3~fs. A slight energy drift on the | 
| 313 |  |  | order of 0.012~kcal/mol per nanosecond was observed at a time step of | 
| 314 |  |  | 4~fs, and as expected, this drift increases dramatically with | 
| 315 | chrisfen | 1030 | increasing time step. To insure accuracy in our microcanonical | 
| 316 | chrisfen | 1033 | simulations, time steps were set at 2~fs and kept at this value for | 
| 317 | chrisfen | 1030 | constant pressure simulations as well. | 
| 318 | chrisfen | 743 |  | 
| 319 | gezelter | 921 | Proton-disordered ice crystals in both the $I_h$ and $I_c$ lattices | 
| 320 |  |  | were generated as starting points for all simulations. The $I_h$ | 
| 321 | chrisfen | 1033 | crystals were formed by first arranging the centers of mass of the SSD | 
| 322 |  |  | particles into a ``hexagonal'' ice lattice of 1024 particles. Because | 
| 323 |  |  | of the crystal structure of $I_h$ ice, the simulation box assumed an | 
| 324 |  |  | orthorhombic shape with an edge length ratio of approximately | 
| 325 |  |  | 1.00$\times$1.06$\times$1.23. The particles were then allowed to | 
| 326 |  |  | orient freely about fixed positions with angular momenta randomized at | 
| 327 |  |  | 400~K for varying times. The rotational temperature was then scaled | 
| 328 |  |  | down in stages to slowly cool the crystals to 25~K. The particles were | 
| 329 |  |  | then allowed to translate with fixed orientations at a constant | 
| 330 |  |  | pressure of 1 atm for 50~ps at 25~K. Finally, all constraints were | 
| 331 |  |  | removed and the ice crystals were allowed to equilibrate for 50~ps at | 
| 332 |  |  | 25~K and a constant pressure of 1~atm.  This procedure resulted in | 
| 333 |  |  | structurally stable $I_h$ ice crystals that obey the Bernal-Fowler | 
| 334 | chrisfen | 862 | rules.\cite{Bernal33,Rahman72} This method was also utilized in the | 
| 335 | chrisfen | 743 | making of diamond lattice $I_c$ ice crystals, with each cubic | 
| 336 |  |  | simulation box consisting of either 512 or 1000 particles. Only | 
| 337 |  |  | isotropic volume fluctuations were performed under constant pressure, | 
| 338 |  |  | so the ratio of edge lengths remained constant throughout the | 
| 339 |  |  | simulations. | 
| 340 |  |  |  | 
| 341 |  |  | \section{Results and discussion} | 
| 342 |  |  |  | 
| 343 |  |  | Melting studies were performed on the randomized ice crystals using | 
| 344 | gezelter | 921 | isobaric-isothermal (NPT) dynamics. During melting simulations, the | 
| 345 |  |  | melting transition and the density maximum can both be observed, | 
| 346 |  |  | provided that the density maximum occurs in the liquid and not the | 
| 347 |  |  | supercooled regime. An ensemble average from five separate melting | 
| 348 |  |  | simulations was acquired, each starting from different ice crystals | 
| 349 |  |  | generated as described previously. All simulations were equilibrated | 
| 350 | chrisfen | 1033 | for 100~ps prior to a 200~ps data collection run at each temperature | 
| 351 |  |  | setting. The temperature range of study spanned from 25 to 400~K, with | 
| 352 |  |  | a maximum degree increment of 25~K. For regions of interest along this | 
| 353 |  |  | stepwise progression, the temperature increment was decreased from | 
| 354 |  |  | 25~K to 10 and 5~K.  The above equilibration and production times were | 
| 355 | gezelter | 921 | sufficient in that fluctuations in the volume autocorrelation function | 
| 356 | chrisfen | 1033 | were damped out in all simulations in under 20~ps. | 
| 357 | chrisfen | 743 |  | 
| 358 |  |  | \subsection{Density Behavior} | 
| 359 |  |  |  | 
| 360 | chrisfen | 1030 | Our initial simulations focused on the original SSD water model, | 
| 361 |  |  | and an average density versus temperature plot is shown in figure | 
| 362 | gezelter | 921 | \ref{dense1}. Note that the density maximum when using a reaction | 
| 363 | chrisfen | 1033 | field appears between 255 and 265~K.  There were smaller fluctuations | 
| 364 |  |  | in the density at 260~K than at either 255 or 265~K, so we report this | 
| 365 | gezelter | 921 | value as the location of the density maximum. Figure \ref{dense1} was | 
| 366 |  |  | constructed using ice $I_h$ crystals for the initial configuration; | 
| 367 |  |  | though not pictured, the simulations starting from ice $I_c$ crystal | 
| 368 |  |  | configurations showed similar results, with a liquid-phase density | 
| 369 | chrisfen | 1033 | maximum in this same region (between 255 and 260~K). | 
| 370 | gezelter | 921 |  | 
| 371 | gezelter | 1036 | \begin{figure} | 
| 372 |  |  | \begin{center} | 
| 373 |  |  | \epsfxsize=6in | 
| 374 |  |  | \epsfbox{denseSSDnew.eps} | 
| 375 |  |  | \caption{ Density versus temperature for TIP4P [Ref. \citen{Jorgensen98b}], | 
| 376 |  |  | TIP3P [Ref. \citen{Jorgensen98b}], SPC/E [Ref. \citen{Clancy94}], SSD | 
| 377 |  |  | without Reaction Field, SSD, and experiment [Ref. \citen{CRC80}]. The | 
| 378 |  |  | arrows indicate the change in densities observed when turning off the | 
| 379 |  |  | reaction field. The the lower than expected densities for the SSD | 
| 380 |  |  | model were what prompted the original reparameterization of SSD1 | 
| 381 |  |  | [Ref. \citen{Ichiye03}].} | 
| 382 |  |  | \label{dense1} | 
| 383 |  |  | \end{center} | 
| 384 |  |  | \end{figure} | 
| 385 | chrisfen | 743 |  | 
| 386 | chrisfen | 1030 | The density maximum for SSD compares quite favorably to other | 
| 387 |  |  | simple water models. Figure \ref{dense1} also shows calculated | 
| 388 |  |  | densities of several other models and experiment obtained from other | 
| 389 | chrisfen | 743 | sources.\cite{Jorgensen98b,Clancy94,CRC80} Of the listed simple water | 
| 390 | chrisfen | 1030 | models, SSD has a temperature closest to the experimentally | 
| 391 |  |  | observed density maximum. Of the {\it charge-based} models in | 
| 392 | gezelter | 921 | Fig. \ref{dense1}, TIP4P has a density maximum behavior most like that | 
| 393 | chrisfen | 1030 | seen in SSD. Though not included in this plot, it is useful to | 
| 394 |  |  | note that TIP5P has a density maximum nearly identical to the | 
| 395 | gezelter | 921 | experimentally measured temperature. | 
| 396 | chrisfen | 743 |  | 
| 397 | gezelter | 921 | It has been observed that liquid state densities in water are | 
| 398 |  |  | dependent on the cutoff radius used both with and without the use of | 
| 399 |  |  | reaction field.\cite{Berendsen98} In order to address the possible | 
| 400 |  |  | effect of cutoff radius, simulations were performed with a dipolar | 
| 401 | chrisfen | 1033 | cutoff radius of 12.0~\AA\ to complement the previous SSD | 
| 402 |  |  | simulations, all performed with a cutoff of 9.0~\AA. All of the | 
| 403 | chrisfen | 1030 | resulting densities overlapped within error and showed no significant | 
| 404 |  |  | trend toward lower or higher densities as a function of cutoff radius, | 
| 405 |  |  | for simulations both with and without reaction field. These results | 
| 406 |  |  | indicate that there is no major benefit in choosing a longer cutoff | 
| 407 |  |  | radius in simulations using SSD. This is advantageous in that | 
| 408 |  |  | the use of a longer cutoff radius results in a significant increase in | 
| 409 |  |  | the time required to obtain a single trajectory. | 
| 410 | chrisfen | 743 |  | 
| 411 | chrisfen | 862 | The key feature to recognize in figure \ref{dense1} is the density | 
| 412 | chrisfen | 1030 | scaling of SSD relative to other common models at any given | 
| 413 |  |  | temperature. SSD assumes a lower density than any of the other | 
| 414 |  |  | listed models at the same pressure, behavior which is especially | 
| 415 | chrisfen | 1033 | apparent at temperatures greater than 300~K. Lower than expected | 
| 416 | chrisfen | 1030 | densities have been observed for other systems using a reaction field | 
| 417 |  |  | for long-range electrostatic interactions, so the most likely reason | 
| 418 |  |  | for the significantly lower densities seen in these simulations is the | 
| 419 | gezelter | 921 | presence of the reaction field.\cite{Berendsen98,Nezbeda02} In order | 
| 420 |  |  | to test the effect of the reaction field on the density of the | 
| 421 |  |  | systems, the simulations were repeated without a reaction field | 
| 422 |  |  | present. The results of these simulations are also displayed in figure | 
| 423 |  |  | \ref{dense1}. Without the reaction field, the densities increase | 
| 424 |  |  | to more experimentally reasonable values, especially around the | 
| 425 |  |  | freezing point of liquid water. The shape of the curve is similar to | 
| 426 | chrisfen | 1030 | the curve produced from SSD simulations using reaction field, | 
| 427 | gezelter | 921 | specifically the rapidly decreasing densities at higher temperatures; | 
| 428 | chrisfen | 1033 | however, a shift in the density maximum location, down to 245~K, is | 
| 429 | gezelter | 921 | observed. This is a more accurate comparison to the other listed water | 
| 430 |  |  | models, in that no long range corrections were applied in those | 
| 431 |  |  | simulations.\cite{Clancy94,Jorgensen98b} However, even without the | 
| 432 | chrisfen | 1033 | reaction field, the density around 300~K is still significantly lower | 
| 433 | chrisfen | 861 | than experiment and comparable water models. This anomalous behavior | 
| 434 | chrisfen | 1027 | was what lead Tan {\it et al.} to recently reparameterize | 
| 435 | chrisfen | 1030 | SSD.\cite{Ichiye03} Throughout the remainder of the paper our | 
| 436 |  |  | reparamaterizations of SSD will be compared with their newer SSD1 | 
| 437 | gezelter | 1029 | model. | 
| 438 | chrisfen | 861 |  | 
| 439 | chrisfen | 743 | \subsection{Transport Behavior} | 
| 440 |  |  |  | 
| 441 | gezelter | 921 | Accurate dynamical properties of a water model are particularly | 
| 442 |  |  | important when using the model to study permeation or transport across | 
| 443 |  |  | biological membranes.  In order to probe transport in bulk water, | 
| 444 |  |  | constant energy (NVE) simulations were performed at the average | 
| 445 |  |  | density obtained by the NPT simulations at an identical target | 
| 446 |  |  | temperature. Simulations started with randomized velocities and | 
| 447 | chrisfen | 1033 | underwent 50~ps of temperature scaling and 50~ps of constant energy | 
| 448 |  |  | equilibration before a 200~ps data collection run. Diffusion constants | 
| 449 | gezelter | 921 | were calculated via linear fits to the long-time behavior of the | 
| 450 |  |  | mean-square displacement as a function of time. The averaged results | 
| 451 |  |  | from five sets of NVE simulations are displayed in figure | 
| 452 |  |  | \ref{diffuse}, alongside experimental, SPC/E, and TIP5P | 
| 453 | chrisfen | 1022 | results.\cite{Gillen72,Holz00,Clancy94,Jorgensen01} | 
| 454 | gezelter | 921 |  | 
| 455 | gezelter | 1036 | \begin{figure} | 
| 456 |  |  | \begin{center} | 
| 457 |  |  | \epsfxsize=6in | 
| 458 |  |  | \epsfbox{betterDiffuse.epsi} | 
| 459 |  |  | \caption{ Average self-diffusion constant as a function of temperature for | 
| 460 |  |  | SSD, SPC/E [Ref. \citen{Clancy94}], and TIP5P | 
| 461 |  |  | [Ref. \citen{Jorgensen01}] compared with experimental data | 
| 462 |  |  | [Refs. \citen{Gillen72} and \citen{Holz00}]. Of the three water models | 
| 463 |  |  | shown, SSD has the least deviation from the experimental values. The | 
| 464 |  |  | rapidly increasing diffusion constants for TIP5P and SSD correspond to | 
| 465 |  |  | significant decreases in density at the higher temperatures.} | 
| 466 |  |  | \label{diffuse} | 
| 467 |  |  | \end{center} | 
| 468 |  |  | \end{figure} | 
| 469 | chrisfen | 743 |  | 
| 470 |  |  | The observed values for the diffusion constant point out one of the | 
| 471 | chrisfen | 1030 | strengths of the SSD model. Of the three models shown, the SSD model | 
| 472 | gezelter | 921 | has the most accurate depiction of self-diffusion in both the | 
| 473 |  |  | supercooled and liquid regimes.  SPC/E does a respectable job by | 
| 474 | chrisfen | 1033 | reproducing values similar to experiment around 290~K; however, it | 
| 475 | gezelter | 921 | deviates at both higher and lower temperatures, failing to predict the | 
| 476 | chrisfen | 1030 | correct thermal trend. TIP5P and SSD both start off low at colder | 
| 477 | gezelter | 921 | temperatures and tend to diffuse too rapidly at higher temperatures. | 
| 478 |  |  | This behavior at higher temperatures is not particularly surprising | 
| 479 | chrisfen | 1030 | since the densities of both TIP5P and SSD are lower than experimental | 
| 480 | gezelter | 921 | water densities at higher temperatures.  When calculating the | 
| 481 | chrisfen | 1030 | diffusion coefficients for SSD at experimental densities | 
| 482 |  |  | (instead of the densities from the NPT simulations), the resulting | 
| 483 |  |  | values fall more in line with experiment at these temperatures. | 
| 484 | chrisfen | 743 |  | 
| 485 |  |  | \subsection{Structural Changes and Characterization} | 
| 486 | gezelter | 921 |  | 
| 487 | chrisfen | 743 | By starting the simulations from the crystalline state, the melting | 
| 488 | gezelter | 921 | transition and the ice structure can be obtained along with the liquid | 
| 489 | chrisfen | 862 | phase behavior beyond the melting point. The constant pressure heat | 
| 490 |  |  | capacity (C$_\text{p}$) was monitored to locate the melting transition | 
| 491 |  |  | in each of the simulations. In the melting simulations of the 1024 | 
| 492 |  |  | particle ice $I_h$ simulations, a large spike in C$_\text{p}$ occurs | 
| 493 | chrisfen | 1033 | at 245~K, indicating a first order phase transition for the melting of | 
| 494 | chrisfen | 862 | these ice crystals. When the reaction field is turned off, the melting | 
| 495 | chrisfen | 1033 | transition occurs at 235~K.  These melting transitions are | 
| 496 | gezelter | 921 | considerably lower than the experimental value. | 
| 497 | chrisfen | 743 |  | 
| 498 | gezelter | 1036 | \begin{figure} | 
| 499 |  |  | \begin{center} | 
| 500 |  |  | \epsfxsize=6in | 
| 501 |  |  | \epsfbox{fullContours.eps} | 
| 502 |  |  | \caption{ Contour plots of 2D angular pair correlation functions for | 
| 503 |  |  | 512 SSD molecules at 100~K (A \& B) and 300~K (C \& D). Dark areas | 
| 504 |  |  | signify regions of enhanced density while light areas signify | 
| 505 |  |  | depletion relative to the bulk density. White areas have pair | 
| 506 |  |  | correlation values below 0.5 and black areas have values above 1.5.} | 
| 507 |  |  | \label{contour} | 
| 508 |  |  | \end{center} | 
| 509 |  |  | \end{figure} | 
| 510 | chrisfen | 862 |  | 
| 511 | gezelter | 1036 | \begin{figure} | 
| 512 |  |  | \begin{center} | 
| 513 |  |  | \epsfxsize=6in | 
| 514 |  |  | \epsfbox{corrDiag.eps} | 
| 515 |  |  | \caption{ An illustration of angles involved in the correlations observed in Fig. \ref{contour}.} | 
| 516 |  |  | \label{corrAngle} | 
| 517 |  |  | \end{center} | 
| 518 |  |  | \end{figure} | 
| 519 | chrisfen | 743 |  | 
| 520 | gezelter | 921 | Additional analysis of the melting process was performed using | 
| 521 |  |  | two-dimensional structure and dipole angle correlations. Expressions | 
| 522 |  |  | for these correlations are as follows: | 
| 523 | chrisfen | 861 |  | 
| 524 | chrisfen | 862 | \begin{equation} | 
| 525 | gezelter | 921 | g_{\text{AB}}(r,\cos\theta) = \frac{V}{N_\text{A}N_\text{B}}\langle\sum_{i\in\text{A}}\sum_{j\in\text{B}}\delta(\cos\theta-\cos\theta_{ij})\delta(r-\left|{\bf r}_{ij}\right|)\rangle\ , | 
| 526 | chrisfen | 862 | \end{equation} | 
| 527 |  |  | \begin{equation} | 
| 528 |  |  | g_{\text{AB}}(r,\cos\omega) = | 
| 529 | gezelter | 921 | \frac{V}{N_\text{A}N_\text{B}}\langle\sum_{i\in\text{A}}\sum_{j\in\text{B}}\delta(\cos\omega-\cos\omega_{ij})\delta(r-\left|{\bf r}_{ij}\right|)\rangle\ , | 
| 530 | chrisfen | 862 | \end{equation} | 
| 531 | chrisfen | 861 | where $\theta$ and $\omega$ refer to the angles shown in figure | 
| 532 |  |  | \ref{corrAngle}. By binning over both distance and the cosine of the | 
| 533 | gezelter | 921 | desired angle between the two dipoles, the $g(r)$ can be analyzed to | 
| 534 |  |  | determine the common dipole arrangements that constitute the peaks and | 
| 535 |  |  | troughs in the standard one-dimensional $g(r)$ plots. Frames A and B | 
| 536 |  |  | of figure \ref{contour} show results from an ice $I_c$ simulation. The | 
| 537 |  |  | first peak in the $g(r)$ consists primarily of the preferred hydrogen | 
| 538 | chrisfen | 861 | bonding arrangements as dictated by the tetrahedral sticky potential - | 
| 539 | gezelter | 921 | one peak for the hydrogen bond donor and the other for the hydrogen | 
| 540 |  |  | bond acceptor.  Due to the high degree of crystallinity of the sample, | 
| 541 |  |  | the second and third solvation shells show a repeated peak arrangement | 
| 542 | chrisfen | 743 | which decays at distances around the fourth solvation shell, near the | 
| 543 |  |  | imposed cutoff for the Lennard-Jones and dipole-dipole interactions. | 
| 544 | chrisfen | 861 | In the higher temperature simulation shown in frames C and D, these | 
| 545 | gezelter | 921 | long-range features deteriorate rapidly. The first solvation shell | 
| 546 |  |  | still shows the strong effect of the sticky-potential, although it | 
| 547 |  |  | covers a larger area, extending to include a fraction of aligned | 
| 548 |  |  | dipole peaks within the first solvation shell. The latter peaks lose | 
| 549 |  |  | due to thermal motion and as the competing dipole force overcomes the | 
| 550 |  |  | sticky potential's tight tetrahedral structuring of the crystal. | 
| 551 | chrisfen | 743 |  | 
| 552 |  |  | This complex interplay between dipole and sticky interactions was | 
| 553 |  |  | remarked upon as a possible reason for the split second peak in the | 
| 554 | gezelter | 1029 | oxygen-oxygen pair correlation function, | 
| 555 |  |  | $g_\mathrm{OO}(r)$.\cite{Ichiye96} At low temperatures, the second | 
| 556 |  |  | solvation shell peak appears to have two distinct components that | 
| 557 |  |  | blend together to form one observable peak. At higher temperatures, | 
| 558 | chrisfen | 1033 | this split character alters to show the leading 4~\AA\ peak dominated | 
| 559 | gezelter | 1029 | by equatorial anti-parallel dipole orientations. There is also a | 
| 560 |  |  | tightly bunched group of axially arranged dipoles that most likely | 
| 561 |  |  | consist of the smaller fraction of aligned dipole pairs. The trailing | 
| 562 | chrisfen | 1033 | component of the split peak at 5~\AA\ is dominated by aligned dipoles | 
| 563 | gezelter | 1029 | that assume hydrogen bond arrangements similar to those seen in the | 
| 564 |  |  | first solvation shell. This evidence indicates that the dipole pair | 
| 565 |  |  | interaction begins to dominate outside of the range of the dipolar | 
| 566 |  |  | repulsion term.  The energetically favorable dipole arrangements | 
| 567 |  |  | populate the region immediately outside this repulsion region (around | 
| 568 | chrisfen | 1033 | 4~\AA), while arrangements that seek to satisfy both the sticky and | 
| 569 | gezelter | 1029 | dipole forces locate themselves just beyond this initial buildup | 
| 570 | chrisfen | 1033 | (around 5~\AA). | 
| 571 | chrisfen | 743 |  | 
| 572 |  |  | From these findings, the split second peak is primarily the product of | 
| 573 | chrisfen | 861 | the dipolar repulsion term of the sticky potential. In fact, the inner | 
| 574 |  |  | peak can be pushed out and merged with the outer split peak just by | 
| 575 | gezelter | 921 | extending the switching function ($s^\prime(r_{ij})$) from its normal | 
| 576 | chrisfen | 1033 | 4.0~\AA\ cutoff to values of 4.5 or even 5~\AA. This type of | 
| 577 | chrisfen | 861 | correction is not recommended for improving the liquid structure, | 
| 578 | chrisfen | 862 | since the second solvation shell would still be shifted too far | 
| 579 | chrisfen | 861 | out. In addition, this would have an even more detrimental effect on | 
| 580 |  |  | the system densities, leading to a liquid with a more open structure | 
| 581 | chrisfen | 1030 | and a density considerably lower than the already low SSD | 
| 582 |  |  | density.  A better correction would be to include the | 
| 583 |  |  | quadrupole-quadrupole interactions for the water particles outside of | 
| 584 |  |  | the first solvation shell, but this would remove the simplicity and | 
| 585 |  |  | speed advantage of SSD. | 
| 586 | chrisfen | 743 |  | 
| 587 | chrisfen | 1030 | \subsection{Adjusted Potentials: SSD/RF and SSD/E} | 
| 588 | gezelter | 921 |  | 
| 589 | chrisfen | 1030 | The propensity of SSD to adopt lower than expected densities under | 
| 590 | chrisfen | 743 | varying conditions is troubling, especially at higher temperatures. In | 
| 591 | chrisfen | 861 | order to correct this model for use with a reaction field, it is | 
| 592 |  |  | necessary to adjust the force field parameters for the primary | 
| 593 |  |  | intermolecular interactions. In undergoing a reparameterization, it is | 
| 594 |  |  | important not to focus on just one property and neglect the other | 
| 595 |  |  | important properties. In this case, it would be ideal to correct the | 
| 596 | gezelter | 921 | densities while maintaining the accurate transport behavior. | 
| 597 | chrisfen | 743 |  | 
| 598 | chrisfen | 1017 | The parameters available for tuning include the $\sigma$ and | 
| 599 |  |  | $\epsilon$ Lennard-Jones parameters, the dipole strength ($\mu$), the | 
| 600 | gezelter | 1029 | strength of the sticky potential ($\nu_0$), and the cutoff distances | 
| 601 |  |  | for the sticky attractive and dipole repulsive cubic switching | 
| 602 |  |  | function cutoffs ($r_l$, $r_u$ and $r_l^\prime$, $r_u^\prime$ | 
| 603 |  |  | respectively). The results of the reparameterizations are shown in | 
| 604 |  |  | table \ref{params}. We are calling these reparameterizations the Soft | 
| 605 | chrisfen | 1030 | Sticky Dipole / Reaction Field (SSD/RF - for use with a reaction | 
| 606 |  |  | field) and Soft Sticky Dipole Extended (SSD/E - an attempt to improve | 
| 607 | gezelter | 1029 | the liquid structure in simulations without a long-range correction). | 
| 608 | chrisfen | 743 |  | 
| 609 |  |  | \begin{table} | 
| 610 | chrisfen | 862 | \begin{center} | 
| 611 | gezelter | 1036 | \caption{ Parameters for the original and adjusted models} | 
| 612 | chrisfen | 856 | \begin{tabular}{ l  c  c  c  c } | 
| 613 | chrisfen | 743 | \hline \\[-3mm] | 
| 614 | gezelter | 1036 | \ \ \ Parameters\ \ \  & \ \ \ SSD [Ref. \citen{Ichiye96}] \ \ \ | 
| 615 |  |  | & \ SSD1 [Ref. \citen{Ichiye03}]\ \  & \ SSD/E\ \  & \ \ SSD/RF \\ | 
| 616 | chrisfen | 743 | \hline \\[-3mm] | 
| 617 | chrisfen | 856 | \ \ \ $\sigma$ (\AA)  & 3.051 & 3.016 & 3.035 & 3.019\\ | 
| 618 |  |  | \ \ \ $\epsilon$ (kcal/mol) & 0.152 & 0.152 & 0.152 & 0.152\\ | 
| 619 |  |  | \ \ \ $\mu$ (D) & 2.35 & 2.35 & 2.42 & 2.48\\ | 
| 620 |  |  | \ \ \ $\nu_0$ (kcal/mol) & 3.7284 & 3.6613 & 3.90 & 3.90\\ | 
| 621 | chrisfen | 1017 | \ \ \ $\omega^\circ$ & 0.07715 & 0.07715 & 0.07715 & 0.07715\\ | 
| 622 | chrisfen | 856 | \ \ \ $r_l$ (\AA) & 2.75 & 2.75 & 2.40 & 2.40\\ | 
| 623 |  |  | \ \ \ $r_u$ (\AA) & 3.35 & 3.35 & 3.80 & 3.80\\ | 
| 624 |  |  | \ \ \ $r_l^\prime$ (\AA) & 2.75 & 2.75 & 2.75 & 2.75\\ | 
| 625 |  |  | \ \ \ $r_u^\prime$ (\AA) & 4.00 & 4.00 & 3.35 & 3.35\\ | 
| 626 | chrisfen | 743 | \end{tabular} | 
| 627 |  |  | \label{params} | 
| 628 | chrisfen | 862 | \end{center} | 
| 629 | chrisfen | 743 | \end{table} | 
| 630 |  |  |  | 
| 631 | gezelter | 1036 | \begin{figure} | 
| 632 |  |  | \begin{center} | 
| 633 |  |  | \epsfxsize=5in | 
| 634 |  |  | \epsfbox{GofRCompare.epsi} | 
| 635 |  |  | \caption{ Plots comparing experiment [Ref. \citen{Head-Gordon00_1}] with | 
| 636 |  |  | SSD/E and SSD1 without reaction field (top), as well as | 
| 637 |  |  | SSD/RF and SSD1 with reaction field turned on | 
| 638 |  |  | (bottom). The insets show the respective first peaks in detail. Note | 
| 639 |  |  | how the changes in parameters have lowered and broadened the first | 
| 640 |  |  | peak of SSD/E and SSD/RF.} | 
| 641 |  |  | \label{grcompare} | 
| 642 |  |  | \end{center} | 
| 643 |  |  | \end{figure} | 
| 644 | chrisfen | 743 |  | 
| 645 | gezelter | 1036 | \begin{figure} | 
| 646 |  |  | \begin{center} | 
| 647 |  |  | \epsfxsize=6in | 
| 648 |  |  | \epsfbox{dualsticky_bw.eps} | 
| 649 |  |  | \caption{ Positive and negative isosurfaces of the sticky potential for | 
| 650 |  |  | SSD1 (left) and SSD/E \& SSD/RF (right). Light areas | 
| 651 |  |  | correspond to the tetrahedral attractive component, and darker areas | 
| 652 |  |  | correspond to the dipolar repulsive component.} | 
| 653 |  |  | \label{isosurface} | 
| 654 |  |  | \end{center} | 
| 655 |  |  | \end{figure} | 
| 656 | chrisfen | 743 |  | 
| 657 | chrisfen | 1030 | In the original paper detailing the development of SSD, Liu and Ichiye | 
| 658 | gezelter | 921 | placed particular emphasis on an accurate description of the first | 
| 659 |  |  | solvation shell. This resulted in a somewhat tall and narrow first | 
| 660 |  |  | peak in $g(r)$ that integrated to give similar coordination numbers to | 
| 661 | chrisfen | 862 | the experimental data obtained by Soper and | 
| 662 |  |  | Phillips.\cite{Ichiye96,Soper86} New experimental x-ray scattering | 
| 663 |  |  | data from the Head-Gordon lab indicates a slightly lower and shifted | 
| 664 | chrisfen | 1030 | first peak in the g$_\mathrm{OO}(r)$, so our adjustments to SSD were | 
| 665 | gezelter | 1029 | made after taking into consideration the new experimental | 
| 666 | chrisfen | 862 | findings.\cite{Head-Gordon00_1} Figure \ref{grcompare} shows the | 
| 667 | gezelter | 921 | relocation of the first peak of the oxygen-oxygen $g(r)$ by comparing | 
| 668 | chrisfen | 1030 | the revised SSD model (SSD1), SSD/E, and SSD/RF to the new | 
| 669 | chrisfen | 862 | experimental results. Both modified water models have shorter peaks | 
| 670 | gezelter | 921 | that match more closely to the experimental peak (as seen in the | 
| 671 |  |  | insets of figure \ref{grcompare}).  This structural alteration was | 
| 672 | chrisfen | 862 | accomplished by the combined reduction in the Lennard-Jones $\sigma$ | 
| 673 | gezelter | 921 | variable and adjustment of the sticky potential strength and cutoffs. | 
| 674 |  |  | As can be seen in table \ref{params}, the cutoffs for the tetrahedral | 
| 675 |  |  | attractive and dipolar repulsive terms were nearly swapped with each | 
| 676 |  |  | other.  Isosurfaces of the original and modified sticky potentials are | 
| 677 |  |  | shown in figure \ref{isosurface}. In these isosurfaces, it is easy to | 
| 678 |  |  | see how altering the cutoffs changes the repulsive and attractive | 
| 679 |  |  | character of the particles. With a reduced repulsive surface (darker | 
| 680 |  |  | region), the particles can move closer to one another, increasing the | 
| 681 | chrisfen | 1030 | density for the overall system.  This change in interaction cutoff | 
| 682 |  |  | also results in a more gradual orientational motion by allowing the | 
| 683 | gezelter | 921 | particles to maintain preferred dipolar arrangements before they begin | 
| 684 |  |  | to feel the pull of the tetrahedral restructuring. As the particles | 
| 685 |  |  | move closer together, the dipolar repulsion term becomes active and | 
| 686 |  |  | excludes unphysical nearest-neighbor arrangements. This compares with | 
| 687 | chrisfen | 1030 | how SSD and SSD1 exclude preferred dipole alignments before the | 
| 688 | gezelter | 921 | particles feel the pull of the ``hydrogen bonds''. Aside from | 
| 689 |  |  | improving the shape of the first peak in the g(\emph{r}), this | 
| 690 |  |  | modification improves the densities considerably by allowing the | 
| 691 | chrisfen | 1033 | persistence of full dipolar character below the previous 4.0~\AA\ | 
| 692 | gezelter | 921 | cutoff. | 
| 693 | chrisfen | 743 |  | 
| 694 | gezelter | 921 | While adjusting the location and shape of the first peak of $g(r)$ | 
| 695 |  |  | improves the densities, these changes alone are insufficient to bring | 
| 696 |  |  | the system densities up to the values observed experimentally.  To | 
| 697 |  |  | further increase the densities, the dipole moments were increased in | 
| 698 | chrisfen | 1033 | both of our adjusted models. Since SSD is a dipole based model, the | 
| 699 |  |  | structure and transport are very sensitive to changes in the dipole | 
| 700 |  |  | moment. The original SSD simply used the dipole moment calculated from | 
| 701 |  |  | the TIP3P water model, which at 2.35~D is significantly greater than | 
| 702 |  |  | the experimental gas phase value of 1.84~D. The larger dipole moment | 
| 703 |  |  | is a more realistic value and improves the dielectric properties of | 
| 704 |  |  | the fluid. Both theoretical and experimental measurements indicate a | 
| 705 |  |  | liquid phase dipole moment ranging from 2.4~D to values as high as | 
| 706 |  |  | 3.11~D, providing a substantial range of reasonable values for a | 
| 707 |  |  | dipole moment.\cite{Sprik91,Kusalik02,Badyal00,Barriol64} Moderately | 
| 708 |  |  | increasing the dipole moments to 2.42 and 2.48~D for SSD/E and SSD/RF, | 
| 709 |  |  | respectively, leads to significant changes in the density and | 
| 710 |  |  | transport of the water models. | 
| 711 | chrisfen | 743 |  | 
| 712 | chrisfen | 861 | In order to demonstrate the benefits of these reparameterizations, a | 
| 713 | chrisfen | 743 | series of NPT and NVE simulations were performed to probe the density | 
| 714 |  |  | and transport properties of the adapted models and compare the results | 
| 715 | chrisfen | 1030 | to the original SSD model. This comparison involved full NPT melting | 
| 716 |  |  | sequences for both SSD/E and SSD/RF, as well as NVE transport | 
| 717 | chrisfen | 861 | calculations at the calculated self-consistent densities. Again, the | 
| 718 | chrisfen | 862 | results are obtained from five separate simulations of 1024 particle | 
| 719 |  |  | systems, and the melting sequences were started from different ice | 
| 720 |  |  | $I_h$ crystals constructed as described previously. Each NPT | 
| 721 | chrisfen | 1033 | simulation was equilibrated for 100~ps before a 200~ps data collection | 
| 722 | chrisfen | 862 | run at each temperature step, and the final configuration from the | 
| 723 |  |  | previous temperature simulation was used as a starting point. All NVE | 
| 724 |  |  | simulations had the same thermalization, equilibration, and data | 
| 725 | gezelter | 921 | collection times as stated previously. | 
| 726 | chrisfen | 743 |  | 
| 727 | gezelter | 1036 | \begin{figure} | 
| 728 |  |  | \begin{center} | 
| 729 |  |  | \epsfxsize=6in | 
| 730 |  |  | \epsfbox{ssdeDense.epsi} | 
| 731 |  |  | \caption{ Comparison of densities calculated with SSD/E to | 
| 732 |  |  | SSD1 without a reaction field, TIP3P [Ref. \citen{Jorgensen98b}], | 
| 733 |  |  | TIP5P [Ref. \citen{Jorgensen00}], SPC/E [Ref. \citen{Clancy94}] and | 
| 734 |  |  | experiment [Ref. \citen{CRC80}]. The window shows a expansion around | 
| 735 |  |  | 300 K with error bars included to clarify this region of | 
| 736 |  |  | interest. Note that both SSD1 and SSD/E show good agreement with | 
| 737 |  |  | experiment when the long-range correction is neglected.} | 
| 738 |  |  | \label{ssdedense} | 
| 739 |  |  | \end{center} | 
| 740 |  |  | \end{figure} | 
| 741 | chrisfen | 743 |  | 
| 742 | chrisfen | 1040 | Figure \ref{ssdedense} shows the density profile for the SSD/E | 
| 743 | chrisfen | 1030 | model in comparison to SSD1 without a reaction field, other | 
| 744 |  |  | common water models, and experimental results. The calculated | 
| 745 |  |  | densities for both SSD/E and SSD1 have increased | 
| 746 |  |  | significantly over the original SSD model (see | 
| 747 |  |  | fig. \ref{dense1}) and are in better agreement with the experimental | 
| 748 |  |  | values. At 298 K, the densities of SSD/E and SSD1 without | 
| 749 | chrisfen | 862 | a long-range correction are 0.996$\pm$0.001 g/cm$^3$ and | 
| 750 |  |  | 0.999$\pm$0.001 g/cm$^3$ respectively.  These both compare well with | 
| 751 |  |  | the experimental value of 0.997 g/cm$^3$, and they are considerably | 
| 752 | chrisfen | 1030 | better than the SSD value of 0.967$\pm$0.003 g/cm$^3$. The | 
| 753 |  |  | changes to the dipole moment and sticky switching functions have | 
| 754 |  |  | improved the structuring of the liquid (as seen in figure | 
| 755 | chrisfen | 1040 | \ref{grcompare}), but they have shifted the density maximum to much | 
| 756 | chrisfen | 1030 | lower temperatures. This comes about via an increase in the liquid | 
| 757 |  |  | disorder through the weakening of the sticky potential and | 
| 758 |  |  | strengthening of the dipolar character. However, this increasing | 
| 759 |  |  | disorder in the SSD/E model has little effect on the melting | 
| 760 |  |  | transition. By monitoring $C_p$ throughout these simulations, the | 
| 761 | chrisfen | 1033 | melting transition for SSD/E was shown to occur at 235~K.  The | 
| 762 | chrisfen | 1030 | same transition temperature observed with SSD and SSD1. | 
| 763 | chrisfen | 743 |  | 
| 764 | gezelter | 1036 | \begin{figure} | 
| 765 |  |  | \begin{center} | 
| 766 |  |  | \epsfxsize=6in | 
| 767 |  |  | \epsfbox{ssdrfDense.epsi} | 
| 768 |  |  | \caption{ Comparison of densities calculated with SSD/RF to | 
| 769 |  |  | SSD1 with a reaction field, TIP3P [Ref. \citen{Jorgensen98b}], | 
| 770 |  |  | TIP5P [Ref. \citen{Jorgensen00}], SPC/E [Ref. \citen{Clancy94}], and | 
| 771 |  |  | experiment [Ref. \citen{CRC80}]. The inset shows the necessity of | 
| 772 |  |  | reparameterization when utilizing a reaction field long-ranged | 
| 773 |  |  | correction - SSD/RF provides significantly more accurate | 
| 774 |  |  | densities than SSD1 when performing room temperature | 
| 775 |  |  | simulations.} | 
| 776 |  |  | \label{ssdrfdense} | 
| 777 |  |  | \end{center} | 
| 778 |  |  | \end{figure} | 
| 779 | chrisfen | 743 |  | 
| 780 | chrisfen | 862 | Including the reaction field long-range correction in the simulations | 
| 781 | gezelter | 921 | results in a more interesting comparison.  A density profile including | 
| 782 | chrisfen | 1030 | SSD/RF and SSD1 with an active reaction field is shown in figure | 
| 783 | chrisfen | 862 | \ref{ssdrfdense}.  As observed in the simulations without a reaction | 
| 784 | chrisfen | 1030 | field, the densities of SSD/RF and SSD1 show a dramatic increase over | 
| 785 |  |  | normal SSD (see figure \ref{dense1}). At 298 K, SSD/RF has a density | 
| 786 | chrisfen | 862 | of 0.997$\pm$0.001 g/cm$^3$, directly in line with experiment and | 
| 787 | chrisfen | 1030 | considerably better than the original SSD value of 0.941$\pm$0.001 | 
| 788 |  |  | g/cm$^3$ and the SSD1 value of 0.972$\pm$0.002 g/cm$^3$. These results | 
| 789 | gezelter | 921 | further emphasize the importance of reparameterization in order to | 
| 790 |  |  | model the density properly under different simulation conditions. | 
| 791 |  |  | Again, these changes have only a minor effect on the melting point, | 
| 792 | chrisfen | 1033 | which observed at 245~K for SSD/RF, is identical to SSD and only 5~K | 
| 793 | chrisfen | 1030 | lower than SSD1 with a reaction field. Additionally, the difference in | 
| 794 |  |  | density maxima is not as extreme, with SSD/RF showing a density | 
| 795 | chrisfen | 1033 | maximum at 255~K, fairly close to the density maxima of 260~K and | 
| 796 |  |  | 265~K, shown by SSD and SSD1 respectively. | 
| 797 | chrisfen | 743 |  | 
| 798 | gezelter | 1036 | \begin{figure} | 
| 799 |  |  | \begin{center} | 
| 800 |  |  | \epsfxsize=6in | 
| 801 |  |  | \epsfbox{ssdeDiffuse.epsi} | 
| 802 |  |  | \caption{ The diffusion constants calculated from SSD/E and | 
| 803 |  |  | SSD1 (both without a reaction field) along with experimental results | 
| 804 |  |  | [Refs. \citen{Gillen72} and \citen{Holz00}]. The NVE calculations were | 
| 805 |  |  | performed at the average densities observed in the 1 atm NPT | 
| 806 |  |  | simulations for the respective models. SSD/E is slightly more mobile | 
| 807 |  |  | than experiment at all of the temperatures, but it is closer to | 
| 808 |  |  | experiment at biologically relevant temperatures than SSD1 without a | 
| 809 |  |  | long-range correction.} | 
| 810 |  |  | \label{ssdediffuse} | 
| 811 |  |  | \end{center} | 
| 812 |  |  | \end{figure} | 
| 813 | chrisfen | 861 |  | 
| 814 | chrisfen | 1030 | The reparameterization of the SSD water model, both for use with and | 
| 815 | chrisfen | 743 | without an applied long-range correction, brought the densities up to | 
| 816 |  |  | what is expected for simulating liquid water. In addition to improving | 
| 817 | chrisfen | 1030 | the densities, it is important that the diffusive behavior of SSD be | 
| 818 | gezelter | 1029 | maintained or improved. Figure \ref{ssdediffuse} compares the | 
| 819 | chrisfen | 1030 | temperature dependence of the diffusion constant of SSD/E to SSD1 | 
| 820 | chrisfen | 1027 | without an active reaction field at the densities calculated from | 
| 821 |  |  | their respective NPT simulations at 1 atm. The diffusion constant for | 
| 822 | chrisfen | 1030 | SSD/E is consistently higher than experiment, while SSD1 remains lower | 
| 823 | chrisfen | 1027 | than experiment until relatively high temperatures (around 360 | 
| 824 |  |  | K). Both models follow the shape of the experimental curve well below | 
| 825 | chrisfen | 1033 | 300~K but tend to diffuse too rapidly at higher temperatures, as seen | 
| 826 |  |  | in SSD1's crossing above 360~K.  This increasing diffusion relative to | 
| 827 | chrisfen | 1027 | the experimental values is caused by the rapidly decreasing system | 
| 828 | chrisfen | 1030 | density with increasing temperature.  Both SSD1 and SSD/E show this | 
| 829 | chrisfen | 1027 | deviation in particle mobility, but this trend has different | 
| 830 | chrisfen | 1030 | implications on the diffusive behavior of the models.  While SSD1 | 
| 831 | chrisfen | 1027 | shows more experimentally accurate diffusive behavior in the high | 
| 832 | chrisfen | 1030 | temperature regimes, SSD/E shows more accurate behavior in the | 
| 833 | chrisfen | 1027 | supercooled and biologically relevant temperature ranges.  Thus, the | 
| 834 |  |  | changes made to improve the liquid structure may have had an adverse | 
| 835 |  |  | affect on the density maximum, but they improve the transport behavior | 
| 836 | chrisfen | 1030 | of SSD/E relative to SSD1 under the most commonly simulated | 
| 837 | chrisfen | 1027 | conditions. | 
| 838 | chrisfen | 743 |  | 
| 839 | gezelter | 1036 | \begin{figure} | 
| 840 |  |  | \begin{center} | 
| 841 |  |  | \epsfxsize=6in | 
| 842 |  |  | \epsfbox{ssdrfDiffuse.epsi} | 
| 843 |  |  | \caption{ The diffusion constants calculated from SSD/RF and | 
| 844 |  |  | SSD1 (both with an active reaction field) along with | 
| 845 |  |  | experimental results [Refs. \citen{Gillen72} and \citen{Holz00}]. The | 
| 846 |  |  | NVE calculations were performed at the average densities observed in | 
| 847 |  |  | the 1 atm NPT simulations for both of the models. SSD/RF | 
| 848 |  |  | simulates the diffusion of water throughout this temperature range | 
| 849 |  |  | very well. The rapidly increasing diffusion constants at high | 
| 850 |  |  | temperatures for both models can be attributed to lower calculated | 
| 851 |  |  | densities than those observed in experiment.} | 
| 852 |  |  | \label{ssdrfdiffuse} | 
| 853 |  |  | \end{center} | 
| 854 |  |  | \end{figure} | 
| 855 | chrisfen | 743 |  | 
| 856 | chrisfen | 1030 | In figure \ref{ssdrfdiffuse}, the diffusion constants for SSD/RF are | 
| 857 |  |  | compared to SSD1 with an active reaction field. Note that SSD/RF | 
| 858 | gezelter | 921 | tracks the experimental results quantitatively, identical within error | 
| 859 | chrisfen | 1017 | throughout most of the temperature range shown and exhibiting only a | 
| 860 | chrisfen | 1030 | slight increasing trend at higher temperatures. SSD1 tends to diffuse | 
| 861 | chrisfen | 1017 | more slowly at low temperatures and deviates to diffuse too rapidly at | 
| 862 | chrisfen | 1033 | temperatures greater than 330~K.  As stated above, this deviation away | 
| 863 | gezelter | 921 | from the ideal trend is due to a rapid decrease in density at higher | 
| 864 | chrisfen | 1030 | temperatures. SSD/RF does not suffer from this problem as much as SSD1 | 
| 865 | gezelter | 921 | because the calculated densities are closer to the experimental | 
| 866 |  |  | values. These results again emphasize the importance of careful | 
| 867 |  |  | reparameterization when using an altered long-range correction. | 
| 868 | chrisfen | 743 |  | 
| 869 | chrisfen | 1017 | \begin{table} | 
| 870 | gezelter | 1029 | \begin{minipage}{\linewidth} | 
| 871 |  |  | \renewcommand{\thefootnote}{\thempfootnote} | 
| 872 | chrisfen | 1017 | \begin{center} | 
| 873 | gezelter | 1036 | \caption{ Properties of the single-point water models compared with | 
| 874 | chrisfen | 1033 | experimental data at ambient conditions. Deviations of the of the | 
| 875 |  |  | averages are given in parentheses.} | 
| 876 | chrisfen | 1017 | \begin{tabular}{ l  c  c  c  c  c } | 
| 877 |  |  | \hline \\[-3mm] | 
| 878 | chrisfen | 1033 | \ \ \ \ \ \  & \ \ \ SSD1 \ \ \ & \ \ SSD/E \ \ \ & \ \ SSD1 (RF) \ \ | 
| 879 |  |  | \ & \ \ SSD/RF \ \ \ & \ \ Expt. \\ | 
| 880 | chrisfen | 1017 | \hline \\[-3mm] | 
| 881 | chrisfen | 1033 | \ \ $\rho$ (g/cm$^3$) & 0.999(0.001) & 0.996(0.001) & 0.972(0.002) & 0.997(0.001) & 0.997 \\ | 
| 882 |  |  | \ \ $C_p$ (cal/mol K) & 28.80(0.11) & 25.45(0.09) & 28.28(0.06) & 23.83(0.16) & 17.98 \\ | 
| 883 |  |  | \ \ $D$ ($10^{-5}$ cm$^2$/s) & 1.78(0.7) & 2.51(0.18) & 2.00(0.17) & 2.32(0.06) & 2.299\cite{Mills73} \\ | 
| 884 |  |  | \ \ Coordination Number ($n_C$) & 3.9 & 4.3 & 3.8 & 4.4 & | 
| 885 | gezelter | 1029 | 4.7\footnote{Calculated by integrating $g_{\text{OO}}(r)$ in | 
| 886 | gezelter | 1036 | Ref. \citen{Head-Gordon00_1}} \\ | 
| 887 | chrisfen | 1033 | \ \ H-bonds per particle ($n_H$) & 3.7 & 3.6 & 3.7 & 3.7 & | 
| 888 | gezelter | 1029 | 3.5\footnote{Calculated by integrating $g_{\text{OH}}(r)$ in | 
| 889 | gezelter | 1036 | Ref. \citen{Soper86}}  \\ | 
| 890 |  |  | \ \ $\tau_1$ (ps) & 10.9(0.6) & 7.3(0.4) & 7.5(0.7) & 7.2(0.4) & 5.7\footnote{Calculated for 298 K from data in Ref. \citen{Eisenberg69}} \\ | 
| 891 | chrisfen | 1033 | \ \ $\tau_2$ (ps) & 4.7(0.4) & 3.1(0.2) & 3.5(0.3) & 3.2(0.2) & 2.3\footnote{Calculated for 298 K from data in | 
| 892 | gezelter | 1036 | Ref. \citen{Krynicki66}} | 
| 893 | chrisfen | 1017 | \end{tabular} | 
| 894 |  |  | \label{liquidproperties} | 
| 895 |  |  | \end{center} | 
| 896 | gezelter | 1029 | \end{minipage} | 
| 897 | chrisfen | 1017 | \end{table} | 
| 898 |  |  |  | 
| 899 |  |  | Table \ref{liquidproperties} gives a synopsis of the liquid state | 
| 900 |  |  | properties of the water models compared in this study along with the | 
| 901 |  |  | experimental values for liquid water at ambient conditions. The | 
| 902 | gezelter | 1029 | coordination number ($n_C$) and number of hydrogen bonds per particle | 
| 903 |  |  | ($n_H$) were calculated by integrating the following relations: | 
| 904 | chrisfen | 1017 | \begin{equation} | 
| 905 | gezelter | 1029 | n_C = 4\pi\rho_{\text{OO}}\int_{0}^{a}r^2\text{g}_{\text{OO}}(r)dr, | 
| 906 | chrisfen | 1017 | \end{equation} | 
| 907 | chrisfen | 1027 | \begin{equation} | 
| 908 | gezelter | 1029 | n_H = 4\pi\rho_{\text{OH}}\int_{0}^{b}r^2\text{g}_{\text{OH}}(r)dr, | 
| 909 | chrisfen | 1027 | \end{equation} | 
| 910 |  |  | where $\rho$ is the number density of the specified pair interactions, | 
| 911 |  |  | $a$ and $b$ are the radial locations of the minima following the first | 
| 912 | gezelter | 1029 | peak in g$_\text{OO}(r)$ or g$_\text{OH}(r)$ respectively. The number | 
| 913 |  |  | of hydrogen bonds stays relatively constant across all of the models, | 
| 914 | chrisfen | 1030 | but the coordination numbers of SSD/E and SSD/RF show an | 
| 915 |  |  | improvement over SSD1.  This improvement is primarily due to | 
| 916 |  |  | extension of the first solvation shell in the new parameter sets. | 
| 917 |  |  | Because $n_H$ and $n_C$ are nearly identical in SSD1, it appears | 
| 918 |  |  | that all molecules in the first solvation shell are involved in | 
| 919 |  |  | hydrogen bonds.  Since $n_H$ and $n_C$ differ in the newly | 
| 920 |  |  | parameterized models, the orientations in the first solvation shell | 
| 921 |  |  | are a bit more ``fluid''.  Therefore SSD1 overstructures the | 
| 922 |  |  | first solvation shell and our reparameterizations have returned this | 
| 923 |  |  | shell to more realistic liquid-like behavior. | 
| 924 | chrisfen | 1017 |  | 
| 925 | gezelter | 1029 | The time constants for the orientational autocorrelation functions | 
| 926 | chrisfen | 1017 | are also displayed in Table \ref{liquidproperties}. The dipolar | 
| 927 | gezelter | 1029 | orientational time correlation functions ($C_{l}$) are described | 
| 928 | chrisfen | 1017 | by: | 
| 929 |  |  | \begin{equation} | 
| 930 | gezelter | 1029 | C_{l}(t) = \langle P_l[\hat{\mathbf{u}}_j(0)\cdot\hat{\mathbf{u}}_j(t)]\rangle, | 
| 931 | chrisfen | 1017 | \end{equation} | 
| 932 | gezelter | 1029 | where $P_l$ are Legendre polynomials of order $l$ and | 
| 933 |  |  | $\hat{\mathbf{u}}_j$ is the unit vector pointing along the molecular | 
| 934 |  |  | dipole.\cite{Rahman71} From these correlation functions, the | 
| 935 |  |  | orientational relaxation time of the dipole vector can be calculated | 
| 936 |  |  | from an exponential fit in the long-time regime ($t > | 
| 937 |  |  | \tau_l$).\cite{Rothschild84} Calculation of these time constants were | 
| 938 |  |  | averaged over five detailed NVE simulations performed at the ambient | 
| 939 |  |  | conditions for each of the respective models. It should be noted that | 
| 940 |  |  | the commonly cited value of 1.9 ps for $\tau_2$ was determined from | 
| 941 | gezelter | 1036 | the NMR data in Ref. \citen{Krynicki66} at a temperature near | 
| 942 | gezelter | 1029 | 34$^\circ$C.\cite{Rahman71} Because of the strong temperature | 
| 943 | chrisfen | 1033 | dependence of $\tau_2$, it is necessary to recalculate it at 298~K to | 
| 944 | gezelter | 1029 | make proper comparisons. The value shown in Table | 
| 945 | chrisfen | 1022 | \ref{liquidproperties} was calculated from the same NMR data in the | 
| 946 | gezelter | 1036 | fashion described in Ref. \citen{Krynicki66}. Similarly, $\tau_1$ was | 
| 947 |  |  | recomputed for 298~K from the data in Ref. \citen{Eisenberg69}. | 
| 948 | chrisfen | 1030 | Again, SSD/E and SSD/RF show improved behavior over SSD1, both with | 
| 949 | chrisfen | 1027 | and without an active reaction field. Turning on the reaction field | 
| 950 | chrisfen | 1030 | leads to much improved time constants for SSD1; however, these results | 
| 951 | gezelter | 1029 | also include a corresponding decrease in system density. | 
| 952 | chrisfen | 1030 | Orientational relaxation times published in the original SSD dynamics | 
| 953 | gezelter | 1029 | papers are smaller than the values observed here, and this difference | 
| 954 |  |  | can be attributed to the use of the Ewald sum.\cite{Ichiye99} | 
| 955 | chrisfen | 1017 |  | 
| 956 | chrisfen | 743 | \subsection{Additional Observations} | 
| 957 |  |  |  | 
| 958 | gezelter | 1036 | \begin{figure} | 
| 959 |  |  | \begin{center} | 
| 960 |  |  | \epsfxsize=6in | 
| 961 |  |  | \epsfbox{icei_bw.eps} | 
| 962 |  |  | \caption{ The most stable crystal structure assumed by the SSD family | 
| 963 |  |  | of water models.  We refer to this structure as Ice-{\it i} to | 
| 964 |  |  | indicate its origins in computer simulation.  This image was taken of | 
| 965 |  |  | the (001) face of the crystal.} | 
| 966 |  |  | \label{weirdice} | 
| 967 |  |  | \end{center} | 
| 968 |  |  | \end{figure} | 
| 969 | chrisfen | 743 |  | 
| 970 | gezelter | 921 | While performing a series of melting simulations on an early iteration | 
| 971 | chrisfen | 1030 | of SSD/E not discussed in this paper, we observed | 
| 972 |  |  | recrystallization into a novel structure not previously known for | 
| 973 | chrisfen | 1033 | water.  After melting at 235~K, two of five systems underwent | 
| 974 |  |  | crystallization events near 245~K.  The two systems remained | 
| 975 |  |  | crystalline up to 320 and 330~K, respectively.  The crystal exhibits | 
| 976 | chrisfen | 1030 | an expanded zeolite-like structure that does not correspond to any | 
| 977 |  |  | known form of ice.  This appears to be an artifact of the point | 
| 978 |  |  | dipolar models, so to distinguish it from the experimentally observed | 
| 979 |  |  | forms of ice, we have denoted the structure | 
| 980 | gezelter | 1029 | Ice-$\sqrt{\smash[b]{-\text{I}}}$ (Ice-{\it i}).  A large enough | 
| 981 | gezelter | 921 | portion of the sample crystallized that we have been able to obtain a | 
| 982 | gezelter | 1029 | near ideal crystal structure of Ice-{\it i}. Figure \ref{weirdice} | 
| 983 | gezelter | 921 | shows the repeating crystal structure of a typical crystal at 5 | 
| 984 |  |  | K. Each water molecule is hydrogen bonded to four others; however, the | 
| 985 |  |  | hydrogen bonds are bent rather than perfectly straight. This results | 
| 986 |  |  | in a skewed tetrahedral geometry about the central molecule.  In | 
| 987 |  |  | figure \ref{isosurface}, it is apparent that these flexed hydrogen | 
| 988 |  |  | bonds are allowed due to the conical shape of the attractive regions, | 
| 989 |  |  | with the greatest attraction along the direct hydrogen bond | 
| 990 | chrisfen | 863 | configuration. Though not ideal, these flexed hydrogen bonds are | 
| 991 | gezelter | 921 | favorable enough to stabilize an entire crystal generated around them. | 
| 992 | chrisfen | 743 |  | 
| 993 | gezelter | 1029 | Initial simulations indicated that Ice-{\it i} is the preferred ice | 
| 994 | chrisfen | 1030 | structure for at least the SSD/E model. To verify this, a comparison | 
| 995 |  |  | was made between near ideal crystals of ice~$I_h$, ice~$I_c$, and | 
| 996 |  |  | Ice-{\it i} at constant pressure with SSD/E, SSD/RF, and | 
| 997 |  |  | SSD1. Near-ideal versions of the three types of crystals were cooled | 
| 998 |  |  | to 1 K, and enthalpies of formation of each were compared using all | 
| 999 |  |  | three water models.  Enthalpies were estimated from the | 
| 1000 |  |  | isobaric-isothermal simulations using $H=U+P_{\text ext}V$ where | 
| 1001 |  |  | $P_{\text ext}$ is the applied pressure.  A constant value of -60.158 | 
| 1002 |  |  | kcal / mol has been added to place our zero for the enthalpies of | 
| 1003 |  |  | formation for these systems at the traditional state (elemental forms | 
| 1004 |  |  | at standard temperature and pressure).  With every model in the SSD | 
| 1005 |  |  | family, Ice-{\it i} had the lowest calculated enthalpy of formation. | 
| 1006 | chrisfen | 743 |  | 
| 1007 | gezelter | 921 | \begin{table} | 
| 1008 |  |  | \begin{center} | 
| 1009 | gezelter | 1036 | \caption{ Enthalpies of Formation (in kcal / mol) of the three crystal | 
| 1010 | chrisfen | 1030 | structures (at 1 K) exhibited by the SSD family of water models} | 
| 1011 | gezelter | 921 | \begin{tabular}{ l  c  c  c  } | 
| 1012 |  |  | \hline \\[-3mm] | 
| 1013 | chrisfen | 1033 | \ \ \ Water Model \ \ \  & \ \ \ Ice-$I_h$ \ \ \ & \ \ \ Ice-$I_c$ \ \ \  & | 
| 1014 |  |  | \ \ \ \ Ice-{\it i} \\ | 
| 1015 | gezelter | 921 | \hline \\[-3mm] | 
| 1016 | chrisfen | 1030 | \ \ \ SSD/E & -72.444 & -72.450 & -73.748 \\ | 
| 1017 |  |  | \ \ \ SSD/RF & -73.093 & -73.075 & -74.180 \\ | 
| 1018 |  |  | \ \ \ SSD1 & -72.654 & -72.569 & -73.575 \\ | 
| 1019 |  |  | \ \ \ SSD1 (RF) & -72.662 & -72.569 & -73.292 \\ | 
| 1020 | gezelter | 921 | \end{tabular} | 
| 1021 |  |  | \label{iceenthalpy} | 
| 1022 |  |  | \end{center} | 
| 1023 |  |  | \end{table} | 
| 1024 | chrisfen | 743 |  | 
| 1025 | gezelter | 921 | In addition to these energetic comparisons, melting simulations were | 
| 1026 | chrisfen | 1033 | performed with Ice-{\it i} as the initial configuration using SSD/E, | 
| 1027 | chrisfen | 1030 | SSD/RF, and SSD1 both with and without a reaction field. The melting | 
| 1028 |  |  | transitions for both SSD/E and SSD1 without reaction field occurred at | 
| 1029 |  |  | temperature in excess of 375~K.  SSD/RF and SSD1 with a reaction field | 
| 1030 | gezelter | 921 | showed more reasonable melting transitions near 325~K.  These melting | 
| 1031 | chrisfen | 1030 | point observations clearly show that all of the SSD-derived models | 
| 1032 | gezelter | 921 | prefer the ice-{\it i} structure. | 
| 1033 | chrisfen | 743 |  | 
| 1034 |  |  | \section{Conclusions} | 
| 1035 |  |  |  | 
| 1036 | gezelter | 921 | The density maximum and temperature dependence of the self-diffusion | 
| 1037 | chrisfen | 1030 | constant were studied for the SSD water model, both with and | 
| 1038 |  |  | without the use of reaction field, via a series of NPT and NVE | 
| 1039 | gezelter | 921 | simulations. The constant pressure simulations showed a density | 
| 1040 |  |  | maximum near 260 K. In most cases, the calculated densities were | 
| 1041 |  |  | significantly lower than the densities obtained from other water | 
| 1042 | chrisfen | 1030 | models (and experiment). Analysis of self-diffusion showed SSD | 
| 1043 |  |  | to capture the transport properties of water well in both the liquid | 
| 1044 |  |  | and supercooled liquid regimes. | 
| 1045 | gezelter | 921 |  | 
| 1046 | chrisfen | 1030 | In order to correct the density behavior, the original SSD model was | 
| 1047 |  |  | reparameterized for use both with and without a reaction field (SSD/RF | 
| 1048 |  |  | and SSD/E), and comparisons were made with SSD1, Ichiye's density | 
| 1049 |  |  | corrected version of SSD. Both models improve the liquid structure, | 
| 1050 | gezelter | 921 | densities, and diffusive properties under their respective simulation | 
| 1051 |  |  | conditions, indicating the necessity of reparameterization when | 
| 1052 |  |  | changing the method of calculating long-range electrostatic | 
| 1053 |  |  | interactions.  In general, however, these simple water models are | 
| 1054 |  |  | excellent choices for representing explicit water in large scale | 
| 1055 |  |  | simulations of biochemical systems. | 
| 1056 |  |  |  | 
| 1057 |  |  | The existence of a novel low-density ice structure that is preferred | 
| 1058 | chrisfen | 1030 | by the SSD family of water models is somewhat troubling, since | 
| 1059 |  |  | liquid simulations on this family of water models at room temperature | 
| 1060 |  |  | are effectively simulations of supercooled or metastable liquids.  One | 
| 1061 | chrisfen | 1027 | way to destabilize this unphysical ice structure would be to make the | 
| 1062 | gezelter | 921 | range of angles preferred by the attractive part of the sticky | 
| 1063 |  |  | potential much narrower.  This would require extensive | 
| 1064 |  |  | reparameterization to maintain the same level of agreement with the | 
| 1065 |  |  | experiments. | 
| 1066 |  |  |  | 
| 1067 | gezelter | 1029 | Additionally, our initial calculations show that the Ice-{\it i} | 
| 1068 | gezelter | 921 | structure may also be a preferred crystal structure for at least one | 
| 1069 |  |  | other popular multi-point water model (TIP3P), and that much of the | 
| 1070 |  |  | simulation work being done using this popular model could also be at | 
| 1071 |  |  | risk for crystallization into this unphysical structure.  A future | 
| 1072 |  |  | publication will detail the relative stability of the known ice | 
| 1073 |  |  | structures for a wide range of popular water models. | 
| 1074 |  |  |  | 
| 1075 | chrisfen | 743 | \section{Acknowledgments} | 
| 1076 | chrisfen | 777 | Support for this project was provided by the National Science | 
| 1077 |  |  | Foundation under grant CHE-0134881. Computation time was provided by | 
| 1078 |  |  | the Notre Dame Bunch-of-Boxes (B.o.B) computer cluster under NSF grant | 
| 1079 | gezelter | 921 | DMR-0079647. | 
| 1080 | chrisfen | 743 |  | 
| 1081 | chrisfen | 862 | \newpage | 
| 1082 |  |  |  | 
| 1083 | chrisfen | 743 | \bibliographystyle{jcp} | 
| 1084 | chrisfen | 1033 | \bibliography{nptSSD} | 
| 1085 | chrisfen | 743 |  | 
| 1086 |  |  |  | 
| 1087 |  |  | \end{document} |