| 1 | chrisfen | 2575 | %\documentclass[prb,aps,twocolumn,tabularx]{revtex4} | 
| 2 | gezelter | 2617 | %\documentclass[aps,prb,preprint]{revtex4} | 
| 3 | chrisfen | 2741 | \documentclass[11pt]{article} | 
| 4 |  |  | \usepackage{endfloat} | 
| 5 | chrisfen | 2636 | \usepackage{amsmath,bm} | 
| 6 | chrisfen | 2594 | \usepackage{amssymb} | 
| 7 | chrisfen | 2575 | \usepackage{epsf} | 
| 8 |  |  | \usepackage{times} | 
| 9 | gezelter | 2617 | \usepackage{mathptmx} | 
| 10 | chrisfen | 2575 | \usepackage{setspace} | 
| 11 |  |  | \usepackage{tabularx} | 
| 12 |  |  | \usepackage{graphicx} | 
| 13 | chrisfen | 2595 | \usepackage{booktabs} | 
| 14 | chrisfen | 2605 | \usepackage{bibentry} | 
| 15 |  |  | \usepackage{mathrsfs} | 
| 16 | chrisfen | 2575 | \usepackage[ref]{overcite} | 
| 17 |  |  | \pagestyle{plain} | 
| 18 |  |  | \pagenumbering{arabic} | 
| 19 |  |  | \oddsidemargin 0.0cm \evensidemargin 0.0cm | 
| 20 |  |  | \topmargin -21pt \headsep 10pt | 
| 21 |  |  | \textheight 9.0in \textwidth 6.5in | 
| 22 |  |  | \brokenpenalty=10000 | 
| 23 | chrisfen | 2742 | %\renewcommand{\baselinestretch}{1.2} | 
| 24 |  |  | \renewcommand{\baselinestretch}{2} | 
| 25 | chrisfen | 2575 | \renewcommand\citemid{\ } % no comma in optional reference note | 
| 26 | chrisfen | 2742 | \AtBeginDelayedFloats{\renewcommand{\baselinestretch}{2}} %doublespace captions | 
| 27 |  |  | \let\Caption\caption | 
| 28 |  |  | \renewcommand\caption[1]{% | 
| 29 |  |  | \Caption[#1]{}% | 
| 30 |  |  | } | 
| 31 | chrisfen | 2575 |  | 
| 32 | chrisfen | 2742 |  | 
| 33 | chrisfen | 2575 | \begin{document} | 
| 34 |  |  |  | 
| 35 | chrisfen | 2667 | \title{Is the Ewald summation still necessary? \\ | 
| 36 | chrisfen | 2742 | Pairwise alternatives to the accepted standard for | 
| 37 | chrisfen | 2741 | long-range electrostatics in molecular simulations} | 
| 38 | chrisfen | 2575 |  | 
| 39 | gezelter | 2617 | \author{Christopher J. Fennell and J. Daniel Gezelter\footnote{Corresponding author. \ Electronic mail: | 
| 40 |  |  | gezelter@nd.edu} \\ | 
| 41 | chrisfen | 2575 | Department of Chemistry and Biochemistry\\ | 
| 42 |  |  | University of Notre Dame\\ | 
| 43 |  |  | Notre Dame, Indiana 46556} | 
| 44 |  |  |  | 
| 45 |  |  | \date{\today} | 
| 46 |  |  |  | 
| 47 |  |  | \maketitle | 
| 48 | chrisfen | 2742 | %\doublespacing | 
| 49 | gezelter | 2617 |  | 
| 50 | chrisfen | 2575 | \begin{abstract} | 
| 51 | gezelter | 2656 | We investigate pairwise electrostatic interaction methods and show | 
| 52 |  |  | that there are viable and computationally efficient $(\mathscr{O}(N))$ | 
| 53 |  |  | alternatives to the Ewald summation for typical modern molecular | 
| 54 |  |  | simulations.  These methods are extended from the damped and | 
| 55 | chrisfen | 2667 | cutoff-neutralized Coulombic sum originally proposed by | 
| 56 |  |  | [D. Wolf, P. Keblinski, S.~R. Phillpot, and J. Eggebrecht, {\it J. Chem. Phys.} {\bf 110}, 8255 (1999)] One of these, the damped shifted force method, shows | 
| 57 | gezelter | 2656 | a remarkable ability to reproduce the energetic and dynamic | 
| 58 |  |  | characteristics exhibited by simulations employing lattice summation | 
| 59 |  |  | techniques.  Comparisons were performed with this and other pairwise | 
| 60 | chrisfen | 2667 | methods against the smooth particle mesh Ewald ({\sc spme}) summation | 
| 61 |  |  | to see how well they reproduce the energetics and dynamics of a | 
| 62 |  |  | variety of simulation types. | 
| 63 | chrisfen | 2575 | \end{abstract} | 
| 64 |  |  |  | 
| 65 | gezelter | 2617 | \newpage | 
| 66 |  |  |  | 
| 67 | chrisfen | 2575 | %\narrowtext | 
| 68 |  |  |  | 
| 69 | chrisfen | 2620 | %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% | 
| 70 | chrisfen | 2575 | %                              BODY OF TEXT | 
| 71 | chrisfen | 2620 | %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% | 
| 72 | chrisfen | 2575 |  | 
| 73 |  |  | \section{Introduction} | 
| 74 |  |  |  | 
| 75 | gezelter | 2617 | In molecular simulations, proper accumulation of the electrostatic | 
| 76 | gezelter | 2643 | interactions is essential and is one of the most | 
| 77 |  |  | computationally-demanding tasks.  The common molecular mechanics force | 
| 78 |  |  | fields represent atomic sites with full or partial charges protected | 
| 79 |  |  | by Lennard-Jones (short range) interactions.  This means that nearly | 
| 80 |  |  | every pair interaction involves a calculation of charge-charge forces. | 
| 81 |  |  | Coupled with relatively long-ranged $r^{-1}$ decay, the monopole | 
| 82 |  |  | interactions quickly become the most expensive part of molecular | 
| 83 |  |  | simulations.  Historically, the electrostatic pair interaction would | 
| 84 |  |  | not have decayed appreciably within the typical box lengths that could | 
| 85 |  |  | be feasibly simulated.  In the larger systems that are more typical of | 
| 86 |  |  | modern simulations, large cutoffs should be used to incorporate | 
| 87 |  |  | electrostatics correctly. | 
| 88 | chrisfen | 2604 |  | 
| 89 | gezelter | 2643 | There have been many efforts to address the proper and practical | 
| 90 |  |  | handling of electrostatic interactions, and these have resulted in a | 
| 91 |  |  | variety of techniques.\cite{Roux99,Sagui99,Tobias01} These are | 
| 92 |  |  | typically classified as implicit methods (i.e., continuum dielectrics, | 
| 93 |  |  | static dipolar fields),\cite{Born20,Grossfield00} explicit methods | 
| 94 |  |  | (i.e., Ewald summations, interaction shifting or | 
| 95 | chrisfen | 2640 | truncation),\cite{Ewald21,Steinbach94} or a mixture of the two (i.e., | 
| 96 | chrisfen | 2639 | reaction field type methods, fast multipole | 
| 97 |  |  | methods).\cite{Onsager36,Rokhlin85} The explicit or mixed methods are | 
| 98 | gezelter | 2643 | often preferred because they physically incorporate solvent molecules | 
| 99 |  |  | in the system of interest, but these methods are sometimes difficult | 
| 100 |  |  | to utilize because of their high computational cost.\cite{Roux99} In | 
| 101 |  |  | addition to the computational cost, there have been some questions | 
| 102 |  |  | regarding possible artifacts caused by the inherent periodicity of the | 
| 103 |  |  | explicit Ewald summation.\cite{Tobias01} | 
| 104 | chrisfen | 2639 |  | 
| 105 | chrisfen | 2667 | In this paper, we focus on a new set of pairwise methods devised by | 
| 106 | gezelter | 2643 | Wolf {\it et al.},\cite{Wolf99} which we further extend.  These | 
| 107 |  |  | methods along with a few other mixed methods (i.e. reaction field) are | 
| 108 |  |  | compared with the smooth particle mesh Ewald | 
| 109 |  |  | sum,\cite{Onsager36,Essmann99} which is our reference method for | 
| 110 |  |  | handling long-range electrostatic interactions. The new methods for | 
| 111 |  |  | handling electrostatics have the potential to scale linearly with | 
| 112 |  |  | increasing system size since they involve only a simple modification | 
| 113 |  |  | to the direct pairwise sum.  They also lack the added periodicity of | 
| 114 |  |  | the Ewald sum, so they can be used for systems which are non-periodic | 
| 115 |  |  | or which have one- or two-dimensional periodicity.  Below, these | 
| 116 | chrisfen | 2740 | methods are evaluated using a variety of model systems to | 
| 117 |  |  | establish their usability in molecular simulations. | 
| 118 | chrisfen | 2639 |  | 
| 119 | chrisfen | 2608 | \subsection{The Ewald Sum} | 
| 120 | chrisfen | 2667 | The complete accumulation of the electrostatic interactions in a system with | 
| 121 | chrisfen | 2639 | periodic boundary conditions (PBC) requires the consideration of the | 
| 122 | gezelter | 2643 | effect of all charges within a (cubic) simulation box as well as those | 
| 123 |  |  | in the periodic replicas, | 
| 124 | chrisfen | 2636 | \begin{equation} | 
| 125 |  |  | V_\textrm{elec} = \frac{1}{2} {\sum_{\mathbf{n}}}^\prime \left[ \sum_{i=1}^N\sum_{j=1}^N \phi\left( \mathbf{r}_{ij} + L\mathbf{n},\bm{\Omega}_i,\bm{\Omega}_j\right) \right], | 
| 126 |  |  | \label{eq:PBCSum} | 
| 127 |  |  | \end{equation} | 
| 128 | chrisfen | 2639 | where the sum over $\mathbf{n}$ is a sum over all periodic box | 
| 129 |  |  | replicas with integer coordinates $\mathbf{n} = (l,m,n)$, and the | 
| 130 |  |  | prime indicates $i = j$ are neglected for $\mathbf{n} = | 
| 131 |  |  | 0$.\cite{deLeeuw80} Within the sum, $N$ is the number of electrostatic | 
| 132 |  |  | particles, $\mathbf{r}_{ij}$ is $\mathbf{r}_j - \mathbf{r}_i$, $L$ is | 
| 133 |  |  | the cell length, $\bm{\Omega}_{i,j}$ are the Euler angles for $i$ and | 
| 134 | gezelter | 2643 | $j$, and $\phi$ is the solution to Poisson's equation | 
| 135 |  |  | ($\phi(\mathbf{r}_{ij}) = q_i q_j |\mathbf{r}_{ij}|^{-1}$ for | 
| 136 |  |  | charge-charge interactions). In the case of monopole electrostatics, | 
| 137 |  |  | eq. (\ref{eq:PBCSum}) is conditionally convergent and is divergent for | 
| 138 |  |  | non-neutral systems. | 
| 139 | chrisfen | 2604 |  | 
| 140 | gezelter | 2643 | The electrostatic summation problem was originally studied by Ewald | 
| 141 | chrisfen | 2636 | for the case of an infinite crystal.\cite{Ewald21}. The approach he | 
| 142 |  |  | took was to convert this conditionally convergent sum into two | 
| 143 |  |  | absolutely convergent summations: a short-ranged real-space summation | 
| 144 |  |  | and a long-ranged reciprocal-space summation, | 
| 145 |  |  | \begin{equation} | 
| 146 |  |  | \begin{split} | 
| 147 | chrisfen | 2637 | V_\textrm{elec} = \frac{1}{2}& \sum_{i=1}^N\sum_{j=1}^N \Biggr[ q_i q_j\Biggr( {\sum_{\mathbf{n}}}^\prime \frac{\textrm{erfc}\left( \alpha|\mathbf{r}_{ij}+\mathbf{n}|\right)}{|\mathbf{r}_{ij}+\mathbf{n}|} \\ &+ \frac{1}{\pi L^3}\sum_{\mathbf{k}\ne 0}\frac{4\pi^2}{|\mathbf{k}|^2} \exp{\left(-\frac{\pi^2|\mathbf{k}|^2}{\alpha^2}\right) \cos\left(\mathbf{k}\cdot\mathbf{r}_{ij}\right)}\Biggr)\Biggr] \\ &- \frac{\alpha}{\pi^{1/2}}\sum_{i=1}^N q_i^2 + \frac{2\pi}{(2\epsilon_\textrm{S}+1)L^3}\left|\sum_{i=1}^N q_i\mathbf{r}_i\right|^2, | 
| 148 | chrisfen | 2636 | \end{split} | 
| 149 |  |  | \label{eq:EwaldSum} | 
| 150 |  |  | \end{equation} | 
| 151 | chrisfen | 2649 | where $\alpha$ is the damping or convergence parameter with units of | 
| 152 |  |  | \AA$^{-1}$, $\mathbf{k}$ are the reciprocal vectors and are equal to | 
| 153 |  |  | $2\pi\mathbf{n}/L^2$, and $\epsilon_\textrm{S}$ is the dielectric | 
| 154 |  |  | constant of the surrounding medium. The final two terms of | 
| 155 | chrisfen | 2636 | eq. (\ref{eq:EwaldSum}) are a particle-self term and a dipolar term | 
| 156 |  |  | for interacting with a surrounding dielectric.\cite{Allen87} This | 
| 157 |  |  | dipolar term was neglected in early applications in molecular | 
| 158 |  |  | simulations,\cite{Brush66,Woodcock71} until it was introduced by de | 
| 159 |  |  | Leeuw {\it et al.} to address situations where the unit cell has a | 
| 160 | gezelter | 2643 | dipole moment which is magnified through replication of the periodic | 
| 161 |  |  | images.\cite{deLeeuw80,Smith81} If this term is taken to be zero, the | 
| 162 |  |  | system is said to be using conducting (or ``tin-foil'') boundary | 
| 163 | chrisfen | 2637 | conditions, $\epsilon_{\rm S} = \infty$. Figure | 
| 164 | chrisfen | 2636 | \ref{fig:ewaldTime} shows how the Ewald sum has been applied over | 
| 165 | gezelter | 2653 | time.  Initially, due to the small system sizes that could be | 
| 166 |  |  | simulated feasibly, the entire simulation box was replicated to | 
| 167 |  |  | convergence.  In more modern simulations, the systems have grown large | 
| 168 |  |  | enough that a real-space cutoff could potentially give convergent | 
| 169 |  |  | behavior.  Indeed, it has been observed that with the choice of a | 
| 170 |  |  | small $\alpha$, the reciprocal-space portion of the Ewald sum can be | 
| 171 |  |  | rapidly convergent and small relative to the real-space | 
| 172 |  |  | portion.\cite{Karasawa89,Kolafa92} | 
| 173 | gezelter | 2643 |  | 
| 174 | chrisfen | 2610 | \begin{figure} | 
| 175 |  |  | \centering | 
| 176 | gezelter | 2656 | \includegraphics[width = \linewidth]{./ewaldProgression.pdf} | 
| 177 | gezelter | 2669 | \caption{The change in the need for the Ewald sum with | 
| 178 |  |  | increasing computational power.  A:~Initially, only small systems | 
| 179 |  |  | could be studied, and the Ewald sum replicated the simulation box to | 
| 180 |  |  | convergence.  B:~Now, radial cutoff methods should be able to reach | 
| 181 |  |  | convergence for the larger systems of charges that are common today.} | 
| 182 | chrisfen | 2610 | \label{fig:ewaldTime} | 
| 183 |  |  | \end{figure} | 
| 184 |  |  |  | 
| 185 | gezelter | 2643 | The original Ewald summation is an $\mathscr{O}(N^2)$ algorithm.  The | 
| 186 | chrisfen | 2649 | convergence parameter $(\alpha)$ plays an important role in balancing | 
| 187 | gezelter | 2643 | the computational cost between the direct and reciprocal-space | 
| 188 |  |  | portions of the summation.  The choice of this value allows one to | 
| 189 |  |  | select whether the real-space or reciprocal space portion of the | 
| 190 |  |  | summation is an $\mathscr{O}(N^2)$ calculation (with the other being | 
| 191 |  |  | $\mathscr{O}(N)$).\cite{Sagui99} With the appropriate choice of | 
| 192 |  |  | $\alpha$ and thoughtful algorithm development, this cost can be | 
| 193 |  |  | reduced to $\mathscr{O}(N^{3/2})$.\cite{Perram88} The typical route | 
| 194 |  |  | taken to reduce the cost of the Ewald summation even further is to set | 
| 195 |  |  | $\alpha$ such that the real-space interactions decay rapidly, allowing | 
| 196 |  |  | for a short spherical cutoff. Then the reciprocal space summation is | 
| 197 |  |  | optimized.  These optimizations usually involve utilization of the | 
| 198 |  |  | fast Fourier transform (FFT),\cite{Hockney81} leading to the | 
| 199 | chrisfen | 2637 | particle-particle particle-mesh (P3M) and particle mesh Ewald (PME) | 
| 200 |  |  | methods.\cite{Shimada93,Luty94,Luty95,Darden93,Essmann95} In these | 
| 201 |  |  | methods, the cost of the reciprocal-space portion of the Ewald | 
| 202 | gezelter | 2643 | summation is reduced from $\mathscr{O}(N^2)$ down to $\mathscr{O}(N | 
| 203 |  |  | \log N)$. | 
| 204 | chrisfen | 2636 |  | 
| 205 | gezelter | 2643 | These developments and optimizations have made the use of the Ewald | 
| 206 |  |  | summation routine in simulations with periodic boundary | 
| 207 |  |  | conditions. However, in certain systems, such as vapor-liquid | 
| 208 |  |  | interfaces and membranes, the intrinsic three-dimensional periodicity | 
| 209 |  |  | can prove problematic.  The Ewald sum has been reformulated to handle | 
| 210 |  |  | 2D systems,\cite{Parry75,Parry76,Heyes77,deLeeuw79,Rhee89}, but the | 
| 211 | chrisfen | 2740 | new methods are computationally expensive.\cite{Spohr97,Yeh99} More | 
| 212 |  |  | recently, there have been several successful efforts toward reducing | 
| 213 |  |  | the computational cost of 2D lattice summations, often enabling the | 
| 214 |  |  | use of the mentioned | 
| 215 |  |  | optimizations.\cite{Yeh99,Kawata01,Arnold02,deJoannis02,Brodka04} | 
| 216 | chrisfen | 2637 |  | 
| 217 |  |  | Several studies have recognized that the inherent periodicity in the | 
| 218 | gezelter | 2643 | Ewald sum can also have an effect on three-dimensional | 
| 219 |  |  | systems.\cite{Roberts94,Roberts95,Luty96,Hunenberger99a,Hunenberger99b,Weber00} | 
| 220 |  |  | Solvated proteins are essentially kept at high concentration due to | 
| 221 |  |  | the periodicity of the electrostatic summation method.  In these | 
| 222 | chrisfen | 2637 | systems, the more compact folded states of a protein can be | 
| 223 |  |  | artificially stabilized by the periodic replicas introduced by the | 
| 224 | gezelter | 2643 | Ewald summation.\cite{Weber00} Thus, care must be taken when | 
| 225 |  |  | considering the use of the Ewald summation where the assumed | 
| 226 |  |  | periodicity would introduce spurious effects in the system dynamics. | 
| 227 | chrisfen | 2637 |  | 
| 228 | chrisfen | 2608 | \subsection{The Wolf and Zahn Methods} | 
| 229 | gezelter | 2617 | In a recent paper by Wolf \textit{et al.}, a procedure was outlined | 
| 230 | gezelter | 2624 | for the accurate accumulation of electrostatic interactions in an | 
| 231 | gezelter | 2643 | efficient pairwise fashion.  This procedure lacks the inherent | 
| 232 |  |  | periodicity of the Ewald summation.\cite{Wolf99} Wolf \textit{et al.} | 
| 233 |  |  | observed that the electrostatic interaction is effectively | 
| 234 |  |  | short-ranged in condensed phase systems and that neutralization of the | 
| 235 |  |  | charge contained within the cutoff radius is crucial for potential | 
| 236 |  |  | stability. They devised a pairwise summation method that ensures | 
| 237 |  |  | charge neutrality and gives results similar to those obtained with the | 
| 238 | chrisfen | 2667 | Ewald summation.  The resulting shifted Coulomb potential includes | 
| 239 |  |  | image-charges subtracted out through placement on the cutoff sphere | 
| 240 |  |  | and a distance-dependent damping function (identical to that seen in | 
| 241 |  |  | the real-space portion of the Ewald sum) to aid convergence | 
| 242 | chrisfen | 2601 | \begin{equation} | 
| 243 | chrisfen | 2640 | V_{\textrm{Wolf}}(r_{ij})= \frac{q_i q_j \textrm{erfc}(\alpha r_{ij})}{r_{ij}}-\lim_{r_{ij}\rightarrow R_\textrm{c}}\left\{\frac{q_iq_j \textrm{erfc}(\alpha r_{ij})}{r_{ij}}\right\}. | 
| 244 | chrisfen | 2601 | \label{eq:WolfPot} | 
| 245 |  |  | \end{equation} | 
| 246 | gezelter | 2617 | Eq. (\ref{eq:WolfPot}) is essentially the common form of a shifted | 
| 247 |  |  | potential.  However, neutralizing the charge contained within each | 
| 248 |  |  | cutoff sphere requires the placement of a self-image charge on the | 
| 249 |  |  | surface of the cutoff sphere.  This additional self-term in the total | 
| 250 | gezelter | 2624 | potential enabled Wolf {\it et al.}  to obtain excellent estimates of | 
| 251 | gezelter | 2617 | Madelung energies for many crystals. | 
| 252 |  |  |  | 
| 253 |  |  | In order to use their charge-neutralized potential in molecular | 
| 254 |  |  | dynamics simulations, Wolf \textit{et al.} suggested taking the | 
| 255 |  |  | derivative of this potential prior to evaluation of the limit.  This | 
| 256 |  |  | procedure gives an expression for the forces, | 
| 257 | chrisfen | 2601 | \begin{equation} | 
| 258 | chrisfen | 2636 | F_{\textrm{Wolf}}(r_{ij}) = q_i q_j\left\{\left[\frac{\textrm{erfc}\left(\alpha r_{ij}\right)}{r^2_{ij}}+\frac{2\alpha}{\pi^{1/2}}\frac{\exp{\left(-\alpha^2r_{ij}^2\right)}}{r_{ij}}\right]-\left[\frac{\textrm{erfc}\left(\alpha R_{\textrm{c}}\right)}{R_{\textrm{c}}^2}+\frac{2\alpha}{\pi^{1/2}}\frac{\exp{\left(-\alpha^2R_{\textrm{c}}^2\right)}}{R_{\textrm{c}}}\right]\right\}, | 
| 259 | chrisfen | 2601 | \label{eq:WolfForces} | 
| 260 |  |  | \end{equation} | 
| 261 | gezelter | 2617 | that incorporates both image charges and damping of the electrostatic | 
| 262 |  |  | interaction. | 
| 263 |  |  |  | 
| 264 |  |  | More recently, Zahn \textit{et al.} investigated these potential and | 
| 265 |  |  | force expressions for use in simulations involving water.\cite{Zahn02} | 
| 266 | gezelter | 2624 | In their work, they pointed out that the forces and derivative of | 
| 267 |  |  | the potential are not commensurate.  Attempts to use both | 
| 268 | gezelter | 2643 | eqs. (\ref{eq:WolfPot}) and (\ref{eq:WolfForces}) together will lead | 
| 269 | gezelter | 2624 | to poor energy conservation.  They correctly observed that taking the | 
| 270 |  |  | limit shown in equation (\ref{eq:WolfPot}) {\it after} calculating the | 
| 271 |  |  | derivatives gives forces for a different potential energy function | 
| 272 | gezelter | 2643 | than the one shown in eq. (\ref{eq:WolfPot}). | 
| 273 | gezelter | 2617 |  | 
| 274 | gezelter | 2643 | Zahn \textit{et al.} introduced a modified form of this summation | 
| 275 |  |  | method as a way to use the technique in Molecular Dynamics | 
| 276 |  |  | simulations.  They proposed a new damped Coulomb potential, | 
| 277 | chrisfen | 2601 | \begin{equation} | 
| 278 | gezelter | 2643 | V_{\textrm{Zahn}}(r_{ij}) = q_iq_j\left\{\frac{\mathrm{erfc}\left(\alpha r_{ij}\right)}{r_{ij}}-\left[\frac{\mathrm{erfc}\left(\alpha R_\mathrm{c}\right)}{R_\mathrm{c}^2}+\frac{2\alpha}{\pi^{1/2}}\frac{\exp\left(-\alpha^2R_\mathrm{c}^2\right)}{R_\mathrm{c}}\right]\left(r_{ij}-R_\mathrm{c}\right)\right\}, | 
| 279 | chrisfen | 2601 | \label{eq:ZahnPot} | 
| 280 |  |  | \end{equation} | 
| 281 | gezelter | 2643 | and showed that this potential does fairly well at capturing the | 
| 282 | gezelter | 2617 | structural and dynamic properties of water compared the same | 
| 283 |  |  | properties obtained using the Ewald sum. | 
| 284 | chrisfen | 2601 |  | 
| 285 | chrisfen | 2608 | \subsection{Simple Forms for Pairwise Electrostatics} | 
| 286 |  |  |  | 
| 287 | gezelter | 2617 | The potentials proposed by Wolf \textit{et al.} and Zahn \textit{et | 
| 288 |  |  | al.} are constructed using two different (and separable) computational | 
| 289 | gezelter | 2624 | tricks: \begin{enumerate} | 
| 290 | gezelter | 2617 | \item shifting through the use of image charges, and | 
| 291 |  |  | \item damping the electrostatic interaction. | 
| 292 | gezelter | 2624 | \end{enumerate}  Wolf \textit{et al.} treated the | 
| 293 | gezelter | 2617 | development of their summation method as a progressive application of | 
| 294 |  |  | these techniques,\cite{Wolf99} while Zahn \textit{et al.} founded | 
| 295 |  |  | their damped Coulomb modification (Eq. (\ref{eq:ZahnPot})) on the | 
| 296 |  |  | post-limit forces (Eq. (\ref{eq:WolfForces})) which were derived using | 
| 297 |  |  | both techniques.  It is possible, however, to separate these | 
| 298 |  |  | tricks and study their effects independently. | 
| 299 |  |  |  | 
| 300 |  |  | Starting with the original observation that the effective range of the | 
| 301 |  |  | electrostatic interaction in condensed phases is considerably less | 
| 302 |  |  | than $r^{-1}$, either the cutoff sphere neutralization or the | 
| 303 |  |  | distance-dependent damping technique could be used as a foundation for | 
| 304 |  |  | a new pairwise summation method.  Wolf \textit{et al.} made the | 
| 305 |  |  | observation that charge neutralization within the cutoff sphere plays | 
| 306 |  |  | a significant role in energy convergence; therefore we will begin our | 
| 307 |  |  | analysis with the various shifted forms that maintain this charge | 
| 308 |  |  | neutralization.  We can evaluate the methods of Wolf | 
| 309 |  |  | \textit{et al.}  and Zahn \textit{et al.} by considering the standard | 
| 310 |  |  | shifted potential, | 
| 311 | chrisfen | 2601 | \begin{equation} | 
| 312 | gezelter | 2643 | V_\textrm{SP}(r) =      \begin{cases} | 
| 313 | gezelter | 2617 | v(r)-v_\textrm{c} &\quad r\leqslant R_\textrm{c} \\ 0 &\quad r > | 
| 314 |  |  | R_\textrm{c} | 
| 315 |  |  | \end{cases}, | 
| 316 |  |  | \label{eq:shiftingPotForm} | 
| 317 |  |  | \end{equation} | 
| 318 |  |  | and shifted force, | 
| 319 |  |  | \begin{equation} | 
| 320 | gezelter | 2643 | V_\textrm{SF}(r) =      \begin{cases} | 
| 321 | gezelter | 2624 | v(r)-v_\textrm{c}-\left(\frac{d v(r)}{d r}\right)_{r=R_\textrm{c}}(r-R_\textrm{c}) | 
| 322 | gezelter | 2617 | &\quad r\leqslant R_\textrm{c} \\ 0 &\quad r > R_\textrm{c} | 
| 323 | chrisfen | 2601 | \end{cases}, | 
| 324 | chrisfen | 2612 | \label{eq:shiftingForm} | 
| 325 | chrisfen | 2601 | \end{equation} | 
| 326 | gezelter | 2617 | functions where $v(r)$ is the unshifted form of the potential, and | 
| 327 |  |  | $v_c$ is $v(R_\textrm{c})$.  The Shifted Force ({\sc sf}) form ensures | 
| 328 |  |  | that both the potential and the forces goes to zero at the cutoff | 
| 329 |  |  | radius, while the Shifted Potential ({\sc sp}) form only ensures the | 
| 330 |  |  | potential is smooth at the cutoff radius | 
| 331 |  |  | ($R_\textrm{c}$).\cite{Allen87} | 
| 332 |  |  |  | 
| 333 | gezelter | 2624 | The forces associated with the shifted potential are simply the forces | 
| 334 |  |  | of the unshifted potential itself (when inside the cutoff sphere), | 
| 335 | chrisfen | 2601 | \begin{equation} | 
| 336 | gezelter | 2643 | F_{\textrm{SP}} = -\left( \frac{d v(r)}{dr} \right), | 
| 337 | chrisfen | 2612 | \end{equation} | 
| 338 | gezelter | 2624 | and are zero outside.  Inside the cutoff sphere, the forces associated | 
| 339 |  |  | with the shifted force form can be written, | 
| 340 | chrisfen | 2612 | \begin{equation} | 
| 341 | gezelter | 2643 | F_{\textrm{SF}} = -\left( \frac{d v(r)}{dr} \right) + \left(\frac{d | 
| 342 | gezelter | 2624 | v(r)}{dr} \right)_{r=R_\textrm{c}}. | 
| 343 |  |  | \end{equation} | 
| 344 |  |  |  | 
| 345 | gezelter | 2643 | If the potential, $v(r)$, is taken to be the normal Coulomb potential, | 
| 346 | gezelter | 2624 | \begin{equation} | 
| 347 |  |  | v(r) = \frac{q_i q_j}{r}, | 
| 348 |  |  | \label{eq:Coulomb} | 
| 349 |  |  | \end{equation} | 
| 350 |  |  | then the Shifted Potential ({\sc sp}) forms will give Wolf {\it et | 
| 351 |  |  | al.}'s undamped prescription: | 
| 352 |  |  | \begin{equation} | 
| 353 | gezelter | 2643 | V_\textrm{SP}(r) = | 
| 354 | gezelter | 2624 | q_iq_j\left(\frac{1}{r}-\frac{1}{R_\textrm{c}}\right) \quad | 
| 355 |  |  | r\leqslant R_\textrm{c}, | 
| 356 | chrisfen | 2636 | \label{eq:SPPot} | 
| 357 | gezelter | 2624 | \end{equation} | 
| 358 |  |  | with associated forces, | 
| 359 |  |  | \begin{equation} | 
| 360 | gezelter | 2643 | F_\textrm{SP}(r) = q_iq_j\left(\frac{1}{r^2}\right) \quad r\leqslant R_\textrm{c}. | 
| 361 | chrisfen | 2636 | \label{eq:SPForces} | 
| 362 | chrisfen | 2612 | \end{equation} | 
| 363 | gezelter | 2624 | These forces are identical to the forces of the standard Coulomb | 
| 364 |  |  | interaction, and cutting these off at $R_c$ was addressed by Wolf | 
| 365 |  |  | \textit{et al.} as undesirable.  They pointed out that the effect of | 
| 366 |  |  | the image charges is neglected in the forces when this form is | 
| 367 |  |  | used,\cite{Wolf99} thereby eliminating any benefit from the method in | 
| 368 |  |  | molecular dynamics.  Additionally, there is a discontinuity in the | 
| 369 |  |  | forces at the cutoff radius which results in energy drift during MD | 
| 370 |  |  | simulations. | 
| 371 | chrisfen | 2612 |  | 
| 372 | gezelter | 2624 | The shifted force ({\sc sf}) form using the normal Coulomb potential | 
| 373 |  |  | will give, | 
| 374 | chrisfen | 2612 | \begin{equation} | 
| 375 | gezelter | 2643 | V_\textrm{SF}(r) = q_iq_j\left[\frac{1}{r}-\frac{1}{R_\textrm{c}}+\left(\frac{1}{R_\textrm{c}^2}\right)(r-R_\textrm{c})\right] \quad r\leqslant R_\textrm{c}. | 
| 376 | chrisfen | 2612 | \label{eq:SFPot} | 
| 377 |  |  | \end{equation} | 
| 378 | gezelter | 2624 | with associated forces, | 
| 379 | chrisfen | 2612 | \begin{equation} | 
| 380 | gezelter | 2643 | F_\textrm{SF}(r) =  q_iq_j\left(\frac{1}{r^2}-\frac{1}{R_\textrm{c}^2}\right) \quad r\leqslant R_\textrm{c}. | 
| 381 | chrisfen | 2612 | \label{eq:SFForces} | 
| 382 |  |  | \end{equation} | 
| 383 | gezelter | 2624 | This formulation has the benefits that there are no discontinuities at | 
| 384 | gezelter | 2643 | the cutoff radius, while the neutralizing image charges are present in | 
| 385 |  |  | both the energy and force expressions.  It would be simple to add the | 
| 386 |  |  | self-neutralizing term back when computing the total energy of the | 
| 387 | gezelter | 2624 | system, thereby maintaining the agreement with the Madelung energies. | 
| 388 |  |  | A side effect of this treatment is the alteration in the shape of the | 
| 389 |  |  | potential that comes from the derivative term.  Thus, a degree of | 
| 390 |  |  | clarity about agreement with the empirical potential is lost in order | 
| 391 |  |  | to gain functionality in dynamics simulations. | 
| 392 | chrisfen | 2612 |  | 
| 393 | chrisfen | 2620 | Wolf \textit{et al.} originally discussed the energetics of the | 
| 394 | gezelter | 2643 | shifted Coulomb potential (Eq. \ref{eq:SPPot}) and found that it was | 
| 395 |  |  | insufficient for accurate determination of the energy with reasonable | 
| 396 |  |  | cutoff distances.  The calculated Madelung energies fluctuated around | 
| 397 |  |  | the expected value as the cutoff radius was increased, but the | 
| 398 |  |  | oscillations converged toward the correct value.\cite{Wolf99} A | 
| 399 | gezelter | 2624 | damping function was incorporated to accelerate the convergence; and | 
| 400 | gezelter | 2643 | though alternative forms for the damping function could be | 
| 401 | gezelter | 2624 | used,\cite{Jones56,Heyes81} the complimentary error function was | 
| 402 |  |  | chosen to mirror the effective screening used in the Ewald summation. | 
| 403 |  |  | Incorporating this error function damping into the simple Coulomb | 
| 404 |  |  | potential, | 
| 405 | chrisfen | 2612 | \begin{equation} | 
| 406 | gezelter | 2624 | v(r) = \frac{\mathrm{erfc}\left(\alpha r\right)}{r}, | 
| 407 | chrisfen | 2601 | \label{eq:dampCoulomb} | 
| 408 |  |  | \end{equation} | 
| 409 | gezelter | 2643 | the shifted potential (eq. (\ref{eq:SPPot})) becomes | 
| 410 | chrisfen | 2601 | \begin{equation} | 
| 411 | gezelter | 2643 | V_{\textrm{DSP}}(r) = q_iq_j\left(\frac{\textrm{erfc}\left(\alpha r\right)}{r}-\frac{\textrm{erfc}\left(\alpha R_\textrm{c}\right)}{R_\textrm{c}}\right) \quad r\leqslant R_\textrm{c}, | 
| 412 | chrisfen | 2612 | \label{eq:DSPPot} | 
| 413 | chrisfen | 2629 | \end{equation} | 
| 414 | gezelter | 2624 | with associated forces, | 
| 415 | chrisfen | 2612 | \begin{equation} | 
| 416 | gezelter | 2643 | F_{\textrm{DSP}}(r) = q_iq_j\left(\frac{\textrm{erfc}\left(\alpha r\right)}{r^2}+\frac{2\alpha}{\pi^{1/2}}\frac{\exp{\left(-\alpha^2r^2\right)}}{r}\right) \quad r\leqslant R_\textrm{c}. | 
| 417 | chrisfen | 2612 | \label{eq:DSPForces} | 
| 418 |  |  | \end{equation} | 
| 419 | gezelter | 2643 | Again, this damped shifted potential suffers from a | 
| 420 |  |  | force-discontinuity at the cutoff radius, and the image charges play | 
| 421 |  |  | no role in the forces.  To remedy these concerns, one may derive a | 
| 422 |  |  | {\sc sf} variant by including the derivative term in | 
| 423 |  |  | eq. (\ref{eq:shiftingForm}), | 
| 424 | chrisfen | 2612 | \begin{equation} | 
| 425 | chrisfen | 2620 | \begin{split} | 
| 426 | gezelter | 2643 | V_\mathrm{DSF}(r) = q_iq_j\Biggr{[}&\frac{\mathrm{erfc}\left(\alpha r\right)}{r}-\frac{\mathrm{erfc}\left(\alpha R_\mathrm{c}\right)}{R_\mathrm{c}} \\ &\left.+\left(\frac{\mathrm{erfc}\left(\alpha R_\mathrm{c}\right)}{R_\mathrm{c}^2}+\frac{2\alpha}{\pi^{1/2}}\frac{\exp\left(-\alpha^2R_\mathrm{c}^2\right)}{R_\mathrm{c}}\right)\left(r-R_\mathrm{c}\right)\ \right] \quad r\leqslant R_\textrm{c}. | 
| 427 | chrisfen | 2612 | \label{eq:DSFPot} | 
| 428 | chrisfen | 2620 | \end{split} | 
| 429 | chrisfen | 2612 | \end{equation} | 
| 430 | chrisfen | 2636 | The derivative of the above potential will lead to the following forces, | 
| 431 | chrisfen | 2612 | \begin{equation} | 
| 432 | chrisfen | 2620 | \begin{split} | 
| 433 | gezelter | 2643 | F_\mathrm{DSF}(r) = q_iq_j\Biggr{[}&\left(\frac{\textrm{erfc}\left(\alpha r\right)}{r^2}+\frac{2\alpha}{\pi^{1/2}}\frac{\exp{\left(-\alpha^2r^2\right)}}{r}\right) \\ &\left.-\left(\frac{\textrm{erfc}\left(\alpha R_{\textrm{c}}\right)}{R_{\textrm{c}}^2}+\frac{2\alpha}{\pi^{1/2}}\frac{\exp{\left(-\alpha^2R_{\textrm{c}}^2\right)}}{R_{\textrm{c}}}\right)\right] \quad r\leqslant R_\textrm{c}. | 
| 434 | chrisfen | 2612 | \label{eq:DSFForces} | 
| 435 | chrisfen | 2620 | \end{split} | 
| 436 | chrisfen | 2612 | \end{equation} | 
| 437 | gezelter | 2643 | If the damping parameter $(\alpha)$ is set to zero, the undamped case, | 
| 438 |  |  | eqs. (\ref{eq:SPPot} through \ref{eq:SFForces}) are correctly | 
| 439 |  |  | recovered from eqs. (\ref{eq:DSPPot} through \ref{eq:DSFForces}). | 
| 440 | chrisfen | 2601 |  | 
| 441 | chrisfen | 2636 | This new {\sc sf} potential is similar to equation \ref{eq:ZahnPot} | 
| 442 |  |  | derived by Zahn \textit{et al.}; however, there are two important | 
| 443 |  |  | differences.\cite{Zahn02} First, the $v_\textrm{c}$ term from | 
| 444 |  |  | eq. (\ref{eq:shiftingForm}) is equal to eq. (\ref{eq:dampCoulomb}) | 
| 445 |  |  | with $r$ replaced by $R_\textrm{c}$.  This term is {\it not} present | 
| 446 |  |  | in the Zahn potential, resulting in a potential discontinuity as | 
| 447 |  |  | particles cross $R_\textrm{c}$.  Second, the sign of the derivative | 
| 448 |  |  | portion is different.  The missing $v_\textrm{c}$ term would not | 
| 449 |  |  | affect molecular dynamics simulations (although the computed energy | 
| 450 |  |  | would be expected to have sudden jumps as particle distances crossed | 
| 451 | gezelter | 2643 | $R_c$).  The sign problem is a potential source of errors, however. | 
| 452 |  |  | In fact, it introduces a discontinuity in the forces at the cutoff, | 
| 453 |  |  | because the force function is shifted in the wrong direction and | 
| 454 |  |  | doesn't cross zero at $R_\textrm{c}$. | 
| 455 | chrisfen | 2602 |  | 
| 456 | gezelter | 2624 | Eqs. (\ref{eq:DSFPot}) and (\ref{eq:DSFForces}) result in an | 
| 457 | gezelter | 2643 | electrostatic summation method in which the potential and forces are | 
| 458 |  |  | continuous at the cutoff radius and which incorporates the damping | 
| 459 |  |  | function proposed by Wolf \textit{et al.}\cite{Wolf99} In the rest of | 
| 460 |  |  | this paper, we will evaluate exactly how good these methods ({\sc sp}, | 
| 461 |  |  | {\sc sf}, damping) are at reproducing the correct electrostatic | 
| 462 |  |  | summation performed by the Ewald sum. | 
| 463 | gezelter | 2624 |  | 
| 464 |  |  | \subsection{Other alternatives} | 
| 465 | gezelter | 2643 | In addition to the methods described above, we considered some other | 
| 466 |  |  | techniques that are commonly used in molecular simulations.  The | 
| 467 | chrisfen | 2629 | simplest of these is group-based cutoffs.  Though of little use for | 
| 468 | gezelter | 2643 | charged molecules, collecting atoms into neutral groups takes | 
| 469 | chrisfen | 2629 | advantage of the observation that the electrostatic interactions decay | 
| 470 |  |  | faster than those for monopolar pairs.\cite{Steinbach94} When | 
| 471 | gezelter | 2643 | considering these molecules as neutral groups, the relative | 
| 472 |  |  | orientations of the molecules control the strength of the interactions | 
| 473 |  |  | at the cutoff radius.  Consequently, as these molecular particles move | 
| 474 |  |  | through $R_\textrm{c}$, the energy will drift upward due to the | 
| 475 |  |  | anisotropy of the net molecular dipole interactions.\cite{Rahman71} To | 
| 476 |  |  | maintain good energy conservation, both the potential and derivative | 
| 477 |  |  | need to be smoothly switched to zero at $R_\textrm{c}$.\cite{Adams79} | 
| 478 |  |  | This is accomplished using a standard switching function.  If a smooth | 
| 479 |  |  | second derivative is desired, a fifth (or higher) order polynomial can | 
| 480 |  |  | be used.\cite{Andrea83} | 
| 481 | gezelter | 2624 |  | 
| 482 | chrisfen | 2629 | Group-based cutoffs neglect the surroundings beyond $R_\textrm{c}$, | 
| 483 | gezelter | 2643 | and to incorporate the effects of the surroundings, a method like | 
| 484 |  |  | Reaction Field ({\sc rf}) can be used.  The original theory for {\sc | 
| 485 |  |  | rf} was originally developed by Onsager,\cite{Onsager36} and it was | 
| 486 |  |  | applied in simulations for the study of water by Barker and | 
| 487 |  |  | Watts.\cite{Barker73} In modern simulation codes, {\sc rf} is simply | 
| 488 |  |  | an extension of the group-based cutoff method where the net dipole | 
| 489 |  |  | within the cutoff sphere polarizes an external dielectric, which | 
| 490 |  |  | reacts back on the central dipole.  The same switching function | 
| 491 |  |  | considerations for group-based cutoffs need to made for {\sc rf}, with | 
| 492 |  |  | the additional pre-specification of a dielectric constant. | 
| 493 | gezelter | 2624 |  | 
| 494 | chrisfen | 2608 | \section{Methods} | 
| 495 |  |  |  | 
| 496 | chrisfen | 2620 | In classical molecular mechanics simulations, there are two primary | 
| 497 |  |  | techniques utilized to obtain information about the system of | 
| 498 |  |  | interest: Monte Carlo (MC) and Molecular Dynamics (MD).  Both of these | 
| 499 |  |  | techniques utilize pairwise summations of interactions between | 
| 500 |  |  | particle sites, but they use these summations in different ways. | 
| 501 | chrisfen | 2608 |  | 
| 502 | gezelter | 2645 | In MC, the potential energy difference between configurations dictates | 
| 503 |  |  | the progression of MC sampling.  Going back to the origins of this | 
| 504 |  |  | method, the acceptance criterion for the canonical ensemble laid out | 
| 505 |  |  | by Metropolis \textit{et al.} states that a subsequent configuration | 
| 506 |  |  | is accepted if $\Delta E < 0$ or if $\xi < \exp(-\Delta E/kT)$, where | 
| 507 |  |  | $\xi$ is a random number between 0 and 1.\cite{Metropolis53} | 
| 508 |  |  | Maintaining the correct $\Delta E$ when using an alternate method for | 
| 509 |  |  | handling the long-range electrostatics will ensure proper sampling | 
| 510 |  |  | from the ensemble. | 
| 511 | chrisfen | 2608 |  | 
| 512 | gezelter | 2624 | In MD, the derivative of the potential governs how the system will | 
| 513 | chrisfen | 2620 | progress in time.  Consequently, the force and torque vectors on each | 
| 514 | gezelter | 2624 | body in the system dictate how the system evolves.  If the magnitude | 
| 515 |  |  | and direction of these vectors are similar when using alternate | 
| 516 |  |  | electrostatic summation techniques, the dynamics in the short term | 
| 517 |  |  | will be indistinguishable.  Because error in MD calculations is | 
| 518 |  |  | cumulative, one should expect greater deviation at longer times, | 
| 519 |  |  | although methods which have large differences in the force and torque | 
| 520 |  |  | vectors will diverge from each other more rapidly. | 
| 521 | chrisfen | 2608 |  | 
| 522 | chrisfen | 2609 | \subsection{Monte Carlo and the Energy Gap}\label{sec:MCMethods} | 
| 523 | gezelter | 2645 |  | 
| 524 | gezelter | 2624 | The pairwise summation techniques (outlined in section | 
| 525 |  |  | \ref{sec:ESMethods}) were evaluated for use in MC simulations by | 
| 526 |  |  | studying the energy differences between conformations.  We took the | 
| 527 | gezelter | 2656 | {\sc spme}-computed energy difference between two conformations to be the | 
| 528 | gezelter | 2624 | correct behavior. An ideal performance by an alternative method would | 
| 529 | gezelter | 2645 | reproduce these energy differences exactly (even if the absolute | 
| 530 |  |  | energies calculated by the methods are different).  Since none of the | 
| 531 |  |  | methods provide exact energy differences, we used linear least squares | 
| 532 |  |  | regressions of energy gap data to evaluate how closely the methods | 
| 533 |  |  | mimicked the Ewald energy gaps.  Unitary results for both the | 
| 534 |  |  | correlation (slope) and correlation coefficient for these regressions | 
| 535 | gezelter | 2656 | indicate perfect agreement between the alternative method and {\sc spme}. | 
| 536 | gezelter | 2645 | Sample correlation plots for two alternate methods are shown in | 
| 537 |  |  | Fig. \ref{fig:linearFit}. | 
| 538 | chrisfen | 2608 |  | 
| 539 | chrisfen | 2609 | \begin{figure} | 
| 540 |  |  | \centering | 
| 541 | chrisfen | 2741 | \includegraphics[width = \linewidth]{./dualLinear.pdf} | 
| 542 | gezelter | 2645 | \caption{Example least squares regressions of the configuration energy | 
| 543 |  |  | differences for SPC/E water systems. The upper plot shows a data set | 
| 544 |  |  | with a poor correlation coefficient ($R^2$), while the lower plot | 
| 545 |  |  | shows a data set with a good correlation coefficient.} | 
| 546 |  |  | \label{fig:linearFit} | 
| 547 | chrisfen | 2609 | \end{figure} | 
| 548 |  |  |  | 
| 549 | chrisfen | 2740 | Each of the seven system types (detailed in section \ref{sec:RepSims}) | 
| 550 |  |  | were represented using 500 independent configurations.  Thus, each of | 
| 551 |  |  | the alternative (non-Ewald) electrostatic summation methods was | 
| 552 |  |  | evaluated using an accumulated 873,250 configurational energy | 
| 553 |  |  | differences. | 
| 554 | chrisfen | 2609 |  | 
| 555 | gezelter | 2624 | Results and discussion for the individual analysis of each of the | 
| 556 | chrisfen | 2742 | system types appear in the supporting information,\cite{EPAPSdeposit} | 
| 557 |  |  | while the cumulative results over all the investigated systems appears | 
| 558 |  |  | below in section \ref{sec:EnergyResults}. | 
| 559 | gezelter | 2624 |  | 
| 560 | chrisfen | 2609 | \subsection{Molecular Dynamics and the Force and Torque Vectors}\label{sec:MDMethods} | 
| 561 | gezelter | 2624 | We evaluated the pairwise methods (outlined in section | 
| 562 |  |  | \ref{sec:ESMethods}) for use in MD simulations by | 
| 563 |  |  | comparing the force and torque vectors with those obtained using the | 
| 564 | gezelter | 2656 | reference Ewald summation ({\sc spme}).  Both the magnitude and the | 
| 565 | gezelter | 2624 | direction of these vectors on each of the bodies in the system were | 
| 566 |  |  | analyzed.  For the magnitude of these vectors, linear least squares | 
| 567 |  |  | regression analyses were performed as described previously for | 
| 568 |  |  | comparing $\Delta E$ values.  Instead of a single energy difference | 
| 569 |  |  | between two system configurations, we compared the magnitudes of the | 
| 570 |  |  | forces (and torques) on each molecule in each configuration.  For a | 
| 571 |  |  | system of 1000 water molecules and 40 ions, there are 1040 force | 
| 572 |  |  | vectors and 1000 torque vectors.  With 500 configurations, this | 
| 573 |  |  | results in 520,000 force and 500,000 torque vector comparisons. | 
| 574 |  |  | Additionally, data from seven different system types was aggregated | 
| 575 |  |  | before the comparison was made. | 
| 576 | chrisfen | 2609 |  | 
| 577 | gezelter | 2624 | The {\it directionality} of the force and torque vectors was | 
| 578 |  |  | investigated through measurement of the angle ($\theta$) formed | 
| 579 | gezelter | 2656 | between those computed from the particular method and those from {\sc spme}, | 
| 580 | chrisfen | 2610 | \begin{equation} | 
| 581 | gezelter | 2645 | \theta_f = \cos^{-1} \left(\hat{F}_\textrm{SPME} \cdot \hat{F}_\textrm{M}\right), | 
| 582 | chrisfen | 2610 | \end{equation} | 
| 583 | gezelter | 2656 | where $\hat{F}_\textrm{M}$ is the unit vector pointing along the force | 
| 584 | chrisfen | 2651 | vector computed using method M.  Each of these $\theta$ values was | 
| 585 |  |  | accumulated in a distribution function and weighted by the area on the | 
| 586 | chrisfen | 2652 | unit sphere.  Since this distribution is a measure of angular error | 
| 587 |  |  | between two different electrostatic summation methods, there is no | 
| 588 |  |  | {\it a priori} reason for the profile to adhere to any specific | 
| 589 |  |  | shape. Thus, gaussian fits were used to measure the width of the | 
| 590 | chrisfen | 2667 | resulting distributions. The variance ($\sigma^2$) was extracted from | 
| 591 |  |  | each of these fits and was used to compare distribution widths. | 
| 592 |  |  | Values of $\sigma^2$ near zero indicate vector directions | 
| 593 |  |  | indistinguishable from those calculated when using the reference | 
| 594 |  |  | method ({\sc spme}). | 
| 595 | gezelter | 2624 |  | 
| 596 |  |  | \subsection{Short-time Dynamics} | 
| 597 | gezelter | 2645 |  | 
| 598 |  |  | The effects of the alternative electrostatic summation methods on the | 
| 599 |  |  | short-time dynamics of charged systems were evaluated by considering a | 
| 600 |  |  | NaCl crystal at a temperature of 1000 K.  A subset of the best | 
| 601 |  |  | performing pairwise methods was used in this comparison.  The NaCl | 
| 602 |  |  | crystal was chosen to avoid possible complications from the treatment | 
| 603 |  |  | of orientational motion in molecular systems.  All systems were | 
| 604 |  |  | started with the same initial positions and velocities.  Simulations | 
| 605 |  |  | were performed under the microcanonical ensemble, and velocity | 
| 606 | chrisfen | 2638 | autocorrelation functions (Eq. \ref{eq:vCorr}) were computed for each | 
| 607 |  |  | of the trajectories, | 
| 608 | chrisfen | 2609 | \begin{equation} | 
| 609 | gezelter | 2656 | C_v(t) = \frac{\langle v(0)\cdot v(t)\rangle}{\langle v^2\rangle}. | 
| 610 | chrisfen | 2609 | \label{eq:vCorr} | 
| 611 |  |  | \end{equation} | 
| 612 | chrisfen | 2638 | Velocity autocorrelation functions require detailed short time data, | 
| 613 |  |  | thus velocity information was saved every 2 fs over 10 ps | 
| 614 |  |  | trajectories. Because the NaCl crystal is composed of two different | 
| 615 |  |  | atom types, the average of the two resulting velocity autocorrelation | 
| 616 |  |  | functions was used for comparisons. | 
| 617 |  |  |  | 
| 618 |  |  | \subsection{Long-Time and Collective Motion}\label{sec:LongTimeMethods} | 
| 619 | gezelter | 2645 |  | 
| 620 |  |  | The effects of the same subset of alternative electrostatic methods on | 
| 621 |  |  | the {\it long-time} dynamics of charged systems were evaluated using | 
| 622 | chrisfen | 2667 | the same model system (NaCl crystals at 1000~K).  The power spectrum | 
| 623 | gezelter | 2645 | ($I(\omega)$) was obtained via Fourier transform of the velocity | 
| 624 |  |  | autocorrelation function, \begin{equation} I(\omega) = | 
| 625 |  |  | \frac{1}{2\pi}\int^{\infty}_{-\infty}C_v(t)e^{-i\omega t}dt, | 
| 626 | chrisfen | 2609 | \label{eq:powerSpec} | 
| 627 |  |  | \end{equation} | 
| 628 | chrisfen | 2638 | where the frequency, $\omega=0,\ 1,\ ...,\ N-1$. Again, because the | 
| 629 |  |  | NaCl crystal is composed of two different atom types, the average of | 
| 630 | gezelter | 2645 | the two resulting power spectra was used for comparisons. Simulations | 
| 631 |  |  | were performed under the microcanonical ensemble, and velocity | 
| 632 | chrisfen | 2740 | information was saved every 5~fs over 100~ps trajectories. | 
| 633 | chrisfen | 2609 |  | 
| 634 |  |  | \subsection{Representative Simulations}\label{sec:RepSims} | 
| 635 | chrisfen | 2740 | A variety of representative molecular simulations were analyzed to | 
| 636 |  |  | determine the relative effectiveness of the pairwise summation | 
| 637 |  |  | techniques in reproducing the energetics and dynamics exhibited by | 
| 638 |  |  | {\sc spme}.  We wanted to span the space of typical molecular | 
| 639 |  |  | simulations (i.e. from liquids of neutral molecules to ionic | 
| 640 |  |  | crystals), so the systems studied were: | 
| 641 | chrisfen | 2599 | \begin{enumerate} | 
| 642 | gezelter | 2645 | \item liquid water (SPC/E),\cite{Berendsen87} | 
| 643 |  |  | \item crystalline water (Ice I$_\textrm{c}$ crystals of SPC/E), | 
| 644 |  |  | \item NaCl crystals, | 
| 645 |  |  | \item NaCl melts, | 
| 646 |  |  | \item a low ionic strength solution of NaCl in water (0.11 M), | 
| 647 |  |  | \item a high ionic strength solution of NaCl in water (1.1 M), and | 
| 648 |  |  | \item a 6 \AA\  radius sphere of Argon in water. | 
| 649 | chrisfen | 2599 | \end{enumerate} | 
| 650 | chrisfen | 2620 | By utilizing the pairwise techniques (outlined in section | 
| 651 |  |  | \ref{sec:ESMethods}) in systems composed entirely of neutral groups, | 
| 652 | gezelter | 2645 | charged particles, and mixtures of the two, we hope to discern under | 
| 653 |  |  | which conditions it will be possible to use one of the alternative | 
| 654 |  |  | summation methodologies instead of the Ewald sum. | 
| 655 | chrisfen | 2586 |  | 
| 656 | gezelter | 2645 | For the solid and liquid water configurations, configurations were | 
| 657 |  |  | taken at regular intervals from high temperature trajectories of 1000 | 
| 658 |  |  | SPC/E water molecules.  Each configuration was equilibrated | 
| 659 |  |  | independently at a lower temperature (300~K for the liquid, 200~K for | 
| 660 |  |  | the crystal).  The solid and liquid NaCl systems consisted of 500 | 
| 661 |  |  | $\textrm{Na}^{+}$ and 500 $\textrm{Cl}^{-}$ ions.  Configurations for | 
| 662 |  |  | these systems were selected and equilibrated in the same manner as the | 
| 663 | chrisfen | 2667 | water systems. In order to introduce measurable fluctuations in the | 
| 664 |  |  | configuration energy differences, the crystalline simulations were | 
| 665 |  |  | equilibrated at 1000~K, near the $T_\textrm{m}$ for NaCl. The liquid | 
| 666 |  |  | NaCl configurations needed to represent a fully disordered array of | 
| 667 |  |  | point charges, so the high temperature of 7000~K was selected for | 
| 668 |  |  | equilibration. The ionic solutions were made by solvating 4 (or 40) | 
| 669 |  |  | ions in a periodic box containing 1000 SPC/E water molecules.  Ion and | 
| 670 |  |  | water positions were then randomly swapped, and the resulting | 
| 671 |  |  | configurations were again equilibrated individually.  Finally, for the | 
| 672 |  |  | Argon / Water ``charge void'' systems, the identities of all the SPC/E | 
| 673 |  |  | waters within 6 \AA\ of the center of the equilibrated water | 
| 674 |  |  | configurations were converted to argon. | 
| 675 | chrisfen | 2586 |  | 
| 676 | gezelter | 2645 | These procedures guaranteed us a set of representative configurations | 
| 677 | gezelter | 2653 | from chemically-relevant systems sampled from appropriate | 
| 678 |  |  | ensembles. Force field parameters for the ions and Argon were taken | 
| 679 | gezelter | 2645 | from the force field utilized by {\sc oopse}.\cite{Meineke05} | 
| 680 |  |  |  | 
| 681 |  |  | \subsection{Comparison of Summation Methods}\label{sec:ESMethods} | 
| 682 |  |  | We compared the following alternative summation methods with results | 
| 683 | gezelter | 2656 | from the reference method ({\sc spme}): | 
| 684 | gezelter | 2645 | \begin{itemize} | 
| 685 |  |  | \item {\sc sp} with damping parameters ($\alpha$) of 0.0, 0.1, 0.2, | 
| 686 |  |  | and 0.3 \AA$^{-1}$, | 
| 687 |  |  | \item {\sc sf} with damping parameters ($\alpha$) of 0.0, 0.1, 0.2, | 
| 688 |  |  | and 0.3 \AA$^{-1}$, | 
| 689 |  |  | \item reaction field with an infinite dielectric constant, and | 
| 690 |  |  | \item an unmodified cutoff. | 
| 691 |  |  | \end{itemize} | 
| 692 |  |  | Group-based cutoffs with a fifth-order polynomial switching function | 
| 693 |  |  | were utilized for the reaction field simulations.  Additionally, we | 
| 694 | gezelter | 2656 | investigated the use of these cutoffs with the {\sc sp}, {\sc sf}, and pure | 
| 695 |  |  | cutoff.  The {\sc spme} electrostatics were performed using the {\sc tinker} | 
| 696 |  |  | implementation of {\sc spme},\cite{Ponder87} while all other calculations | 
| 697 | gezelter | 2653 | were performed using the {\sc oopse} molecular mechanics | 
| 698 | gezelter | 2645 | package.\cite{Meineke05} All other portions of the energy calculation | 
| 699 |  |  | (i.e. Lennard-Jones interactions) were handled in exactly the same | 
| 700 |  |  | manner across all systems and configurations. | 
| 701 | chrisfen | 2586 |  | 
| 702 | chrisfen | 2667 | The alternative methods were also evaluated with three different | 
| 703 | chrisfen | 2649 | cutoff radii (9, 12, and 15 \AA).  As noted previously, the | 
| 704 |  |  | convergence parameter ($\alpha$) plays a role in the balance of the | 
| 705 |  |  | real-space and reciprocal-space portions of the Ewald calculation. | 
| 706 |  |  | Typical molecular mechanics packages set this to a value dependent on | 
| 707 |  |  | the cutoff radius and a tolerance (typically less than $1 \times | 
| 708 |  |  | 10^{-4}$ kcal/mol).  Smaller tolerances are typically associated with | 
| 709 | gezelter | 2653 | increasing accuracy at the expense of computational time spent on the | 
| 710 |  |  | reciprocal-space portion of the summation.\cite{Perram88,Essmann95} | 
| 711 | gezelter | 2656 | The default {\sc tinker} tolerance of $1 \times 10^{-8}$ kcal/mol was used | 
| 712 |  |  | in all {\sc spme} calculations, resulting in Ewald coefficients of 0.4200, | 
| 713 | gezelter | 2653 | 0.3119, and 0.2476 \AA$^{-1}$ for cutoff radii of 9, 12, and 15 \AA\ | 
| 714 |  |  | respectively. | 
| 715 | chrisfen | 2609 |  | 
| 716 | chrisfen | 2575 | \section{Results and Discussion} | 
| 717 |  |  |  | 
| 718 | chrisfen | 2609 | \subsection{Configuration Energy Differences}\label{sec:EnergyResults} | 
| 719 | chrisfen | 2620 | In order to evaluate the performance of the pairwise electrostatic | 
| 720 |  |  | summation methods for Monte Carlo simulations, the energy differences | 
| 721 |  |  | between configurations were compared to the values obtained when using | 
| 722 | gezelter | 2656 | {\sc spme}.  The results for the subsequent regression analysis are shown in | 
| 723 | chrisfen | 2620 | figure \ref{fig:delE}. | 
| 724 | chrisfen | 2590 |  | 
| 725 |  |  | \begin{figure} | 
| 726 |  |  | \centering | 
| 727 | chrisfen | 2741 | \includegraphics[width=5.5in]{./delEplot.pdf} | 
| 728 | gezelter | 2645 | \caption{Statistical analysis of the quality of configurational energy | 
| 729 |  |  | differences for a given electrostatic method compared with the | 
| 730 |  |  | reference Ewald sum.  Results with a value equal to 1 (dashed line) | 
| 731 |  |  | indicate $\Delta E$ values indistinguishable from those obtained using | 
| 732 | gezelter | 2656 | {\sc spme}.  Different values of the cutoff radius are indicated with | 
| 733 | gezelter | 2645 | different symbols (9\AA\ = circles, 12\AA\ = squares, and 15\AA\ = | 
| 734 |  |  | inverted triangles).} | 
| 735 | chrisfen | 2601 | \label{fig:delE} | 
| 736 | chrisfen | 2594 | \end{figure} | 
| 737 |  |  |  | 
| 738 | gezelter | 2645 | The most striking feature of this plot is how well the Shifted Force | 
| 739 |  |  | ({\sc sf}) and Shifted Potential ({\sc sp}) methods capture the energy | 
| 740 |  |  | differences.  For the undamped {\sc sf} method, and the | 
| 741 |  |  | moderately-damped {\sc sp} methods, the results are nearly | 
| 742 |  |  | indistinguishable from the Ewald results.  The other common methods do | 
| 743 |  |  | significantly less well. | 
| 744 | chrisfen | 2594 |  | 
| 745 | gezelter | 2645 | The unmodified cutoff method is essentially unusable.  This is not | 
| 746 |  |  | surprising since hard cutoffs give large energy fluctuations as atoms | 
| 747 |  |  | or molecules move in and out of the cutoff | 
| 748 |  |  | radius.\cite{Rahman71,Adams79} These fluctuations can be alleviated to | 
| 749 |  |  | some degree by using group based cutoffs with a switching | 
| 750 |  |  | function.\cite{Adams79,Steinbach94,Leach01} However, we do not see | 
| 751 |  |  | significant improvement using the group-switched cutoff because the | 
| 752 |  |  | salt and salt solution systems contain non-neutral groups.  Interested | 
| 753 |  |  | readers can consult the accompanying supporting information for a | 
| 754 | chrisfen | 2742 | comparison where all groups are neutral.\cite{EPAPSdeposit} | 
| 755 | gezelter | 2645 |  | 
| 756 | gezelter | 2653 | For the {\sc sp} method, inclusion of electrostatic damping improves | 
| 757 |  |  | the agreement with Ewald, and using an $\alpha$ of 0.2 \AA $^{-1}$ | 
| 758 | gezelter | 2656 | shows an excellent correlation and quality of fit with the {\sc spme} | 
| 759 | gezelter | 2653 | results, particularly with a cutoff radius greater than 12 | 
| 760 | gezelter | 2645 | \AA .  Use of a larger damping parameter is more helpful for the | 
| 761 |  |  | shortest cutoff shown, but it has a detrimental effect on simulations | 
| 762 |  |  | with larger cutoffs. | 
| 763 | chrisfen | 2609 |  | 
| 764 | gezelter | 2653 | In the {\sc sf} sets, increasing damping results in progressively {\it | 
| 765 |  |  | worse} correlation with Ewald.  Overall, the undamped case is the best | 
| 766 | gezelter | 2645 | performing set, as the correlation and quality of fits are | 
| 767 |  |  | consistently superior regardless of the cutoff distance.  The undamped | 
| 768 |  |  | case is also less computationally demanding (because no evaluation of | 
| 769 |  |  | the complementary error function is required). | 
| 770 |  |  |  | 
| 771 |  |  | The reaction field results illustrates some of that method's | 
| 772 |  |  | limitations, primarily that it was developed for use in homogenous | 
| 773 |  |  | systems; although it does provide results that are an improvement over | 
| 774 |  |  | those from an unmodified cutoff. | 
| 775 |  |  |  | 
| 776 | chrisfen | 2608 | \subsection{Magnitudes of the Force and Torque Vectors} | 
| 777 | chrisfen | 2599 |  | 
| 778 | chrisfen | 2620 | Evaluation of pairwise methods for use in Molecular Dynamics | 
| 779 |  |  | simulations requires consideration of effects on the forces and | 
| 780 | gezelter | 2653 | torques.  Figures \ref{fig:frcMag} and \ref{fig:trqMag} show the | 
| 781 |  |  | regression results for the force and torque vector magnitudes, | 
| 782 |  |  | respectively.  The data in these figures was generated from an | 
| 783 |  |  | accumulation of the statistics from all of the system types. | 
| 784 | chrisfen | 2594 |  | 
| 785 |  |  | \begin{figure} | 
| 786 |  |  | \centering | 
| 787 | chrisfen | 2741 | \includegraphics[width=5.5in]{./frcMagplot.pdf} | 
| 788 | chrisfen | 2651 | \caption{Statistical analysis of the quality of the force vector | 
| 789 |  |  | magnitudes for a given electrostatic method compared with the | 
| 790 |  |  | reference Ewald sum.  Results with a value equal to 1 (dashed line) | 
| 791 |  |  | indicate force magnitude values indistinguishable from those obtained | 
| 792 | gezelter | 2656 | using {\sc spme}.  Different values of the cutoff radius are indicated with | 
| 793 | chrisfen | 2651 | different symbols (9\AA\ = circles, 12\AA\ = squares, and 15\AA\ = | 
| 794 |  |  | inverted triangles).} | 
| 795 | chrisfen | 2601 | \label{fig:frcMag} | 
| 796 | chrisfen | 2594 | \end{figure} | 
| 797 |  |  |  | 
| 798 | gezelter | 2653 | Again, it is striking how well the Shifted Potential and Shifted Force | 
| 799 | gezelter | 2656 | methods are doing at reproducing the {\sc spme} forces.  The undamped and | 
| 800 | gezelter | 2653 | weakly-damped {\sc sf} method gives the best agreement with Ewald. | 
| 801 |  |  | This is perhaps expected because this method explicitly incorporates a | 
| 802 |  |  | smooth transition in the forces at the cutoff radius as well as the | 
| 803 |  |  | neutralizing image charges. | 
| 804 |  |  |  | 
| 805 | chrisfen | 2620 | Figure \ref{fig:frcMag}, for the most part, parallels the results seen | 
| 806 |  |  | in the previous $\Delta E$ section.  The unmodified cutoff results are | 
| 807 |  |  | poor, but using group based cutoffs and a switching function provides | 
| 808 | gezelter | 2653 | an improvement much more significant than what was seen with $\Delta | 
| 809 |  |  | E$. | 
| 810 |  |  |  | 
| 811 |  |  | With moderate damping and a large enough cutoff radius, the {\sc sp} | 
| 812 |  |  | method is generating usable forces.  Further increases in damping, | 
| 813 | chrisfen | 2620 | while beneficial for simulations with a cutoff radius of 9 \AA\ , is | 
| 814 | gezelter | 2653 | detrimental to simulations with larger cutoff radii. | 
| 815 |  |  |  | 
| 816 |  |  | The reaction field results are surprisingly good, considering the poor | 
| 817 | chrisfen | 2620 | quality of the fits for the $\Delta E$ results.  There is still a | 
| 818 | gezelter | 2653 | considerable degree of scatter in the data, but the forces correlate | 
| 819 |  |  | well with the Ewald forces in general.  We note that the reaction | 
| 820 |  |  | field calculations do not include the pure NaCl systems, so these | 
| 821 | chrisfen | 2620 | results are partly biased towards conditions in which the method | 
| 822 |  |  | performs more favorably. | 
| 823 | chrisfen | 2594 |  | 
| 824 |  |  | \begin{figure} | 
| 825 |  |  | \centering | 
| 826 | chrisfen | 2741 | \includegraphics[width=5.5in]{./trqMagplot.pdf} | 
| 827 | chrisfen | 2651 | \caption{Statistical analysis of the quality of the torque vector | 
| 828 |  |  | magnitudes for a given electrostatic method compared with the | 
| 829 |  |  | reference Ewald sum.  Results with a value equal to 1 (dashed line) | 
| 830 |  |  | indicate torque magnitude values indistinguishable from those obtained | 
| 831 | gezelter | 2656 | using {\sc spme}.  Different values of the cutoff radius are indicated with | 
| 832 | chrisfen | 2651 | different symbols (9\AA\ = circles, 12\AA\ = squares, and 15\AA\ = | 
| 833 |  |  | inverted triangles).} | 
| 834 | chrisfen | 2601 | \label{fig:trqMag} | 
| 835 | chrisfen | 2594 | \end{figure} | 
| 836 |  |  |  | 
| 837 | gezelter | 2653 | Molecular torques were only available from the systems which contained | 
| 838 |  |  | rigid molecules (i.e. the systems containing water).  The data in | 
| 839 |  |  | fig. \ref{fig:trqMag} is taken from this smaller sampling pool. | 
| 840 | chrisfen | 2594 |  | 
| 841 | gezelter | 2653 | Torques appear to be much more sensitive to charges at a longer | 
| 842 |  |  | distance.   The striking feature in comparing the new electrostatic | 
| 843 | gezelter | 2656 | methods with {\sc spme} is how much the agreement improves with increasing | 
| 844 | gezelter | 2653 | cutoff radius.  Again, the weakly damped and undamped {\sc sf} method | 
| 845 | gezelter | 2656 | appears to be reproducing the {\sc spme} torques most accurately. | 
| 846 | gezelter | 2653 |  | 
| 847 |  |  | Water molecules are dipolar, and the reaction field method reproduces | 
| 848 |  |  | the effect of the surrounding polarized medium on each of the | 
| 849 |  |  | molecular bodies. Therefore it is not surprising that reaction field | 
| 850 |  |  | performs best of all of the methods on molecular torques. | 
| 851 |  |  |  | 
| 852 | chrisfen | 2608 | \subsection{Directionality of the Force and Torque Vectors} | 
| 853 | chrisfen | 2599 |  | 
| 854 | gezelter | 2653 | It is clearly important that a new electrostatic method can reproduce | 
| 855 |  |  | the magnitudes of the force and torque vectors obtained via the Ewald | 
| 856 |  |  | sum. However, the {\it directionality} of these vectors will also be | 
| 857 |  |  | vital in calculating dynamical quantities accurately.  Force and | 
| 858 |  |  | torque directionalities were investigated by measuring the angles | 
| 859 |  |  | formed between these vectors and the same vectors calculated using | 
| 860 | gezelter | 2656 | {\sc spme}.  The results (Fig. \ref{fig:frcTrqAng}) are compared through the | 
| 861 | gezelter | 2653 | variance ($\sigma^2$) of the Gaussian fits of the angle error | 
| 862 |  |  | distributions of the combined set over all system types. | 
| 863 | chrisfen | 2594 |  | 
| 864 |  |  | \begin{figure} | 
| 865 |  |  | \centering | 
| 866 | chrisfen | 2741 | \includegraphics[width=5.5in]{./frcTrqAngplot.pdf} | 
| 867 | gezelter | 2653 | \caption{Statistical analysis of the width of the angular distribution | 
| 868 |  |  | that the force and torque vectors from a given electrostatic method | 
| 869 |  |  | make with their counterparts obtained using the reference Ewald sum. | 
| 870 |  |  | Results with a variance ($\sigma^2$) equal to zero (dashed line) | 
| 871 |  |  | indicate force and torque directions indistinguishable from those | 
| 872 | gezelter | 2656 | obtained using {\sc spme}.  Different values of the cutoff radius are | 
| 873 | gezelter | 2653 | indicated with different symbols (9\AA\ = circles, 12\AA\ = squares, | 
| 874 |  |  | and 15\AA\ = inverted triangles).} | 
| 875 | chrisfen | 2601 | \label{fig:frcTrqAng} | 
| 876 | chrisfen | 2594 | \end{figure} | 
| 877 |  |  |  | 
| 878 | chrisfen | 2620 | Both the force and torque $\sigma^2$ results from the analysis of the | 
| 879 |  |  | total accumulated system data are tabulated in figure | 
| 880 | gezelter | 2653 | \ref{fig:frcTrqAng}. Here it is clear that the Shifted Potential ({\sc | 
| 881 | gezelter | 2656 | sp}) method would be essentially unusable for molecular dynamics | 
| 882 |  |  | unless the damping function is added.  The Shifted Force ({\sc sf}) | 
| 883 |  |  | method, however, is generating force and torque vectors which are | 
| 884 |  |  | within a few degrees of the Ewald results even with weak (or no) | 
| 885 |  |  | damping. | 
| 886 | chrisfen | 2594 |  | 
| 887 | gezelter | 2653 | All of the sets (aside from the over-damped case) show the improvement | 
| 888 |  |  | afforded by choosing a larger cutoff radius.  Increasing the cutoff | 
| 889 |  |  | from 9 to 12 \AA\ typically results in a halving of the width of the | 
| 890 | gezelter | 2656 | distribution, with a similar improvement when going from 12 to 15 | 
| 891 | gezelter | 2653 | \AA . | 
| 892 |  |  |  | 
| 893 |  |  | The undamped {\sc sf}, group-based cutoff, and reaction field methods | 
| 894 |  |  | all do equivalently well at capturing the direction of both the force | 
| 895 | gezelter | 2656 | and torque vectors.  Using the electrostatic damping improves the | 
| 896 |  |  | angular behavior significantly for the {\sc sp} and moderately for the | 
| 897 |  |  | {\sc sf} methods.  Overdamping is detrimental to both methods.  Again | 
| 898 |  |  | it is important to recognize that the force vectors cover all | 
| 899 |  |  | particles in all seven systems, while torque vectors are only | 
| 900 |  |  | available for neutral molecular groups.  Damping is more beneficial to | 
| 901 | gezelter | 2653 | charged bodies, and this observation is investigated further in the | 
| 902 | chrisfen | 2742 | accompanying supporting information.\cite{EPAPSdeposit} | 
| 903 | gezelter | 2653 |  | 
| 904 |  |  | Although not discussed previously, group based cutoffs can be applied | 
| 905 | gezelter | 2656 | to both the {\sc sp} and {\sc sf} methods.  The group-based cutoffs | 
| 906 |  |  | will reintroduce small discontinuities at the cutoff radius, but the | 
| 907 |  |  | effects of these can be minimized by utilizing a switching function. | 
| 908 |  |  | Though there are no significant benefits or drawbacks observed in | 
| 909 |  |  | $\Delta E$ and the force and torque magnitudes when doing this, there | 
| 910 |  |  | is a measurable improvement in the directionality of the forces and | 
| 911 |  |  | torques. Table \ref{tab:groupAngle} shows the angular variances | 
| 912 |  |  | obtained using group based cutoffs along with the results seen in | 
| 913 |  |  | figure \ref{fig:frcTrqAng}.  The {\sc sp} (with an $\alpha$ of 0.2 | 
| 914 |  |  | \AA$^{-1}$ or smaller) shows much narrower angular distributions when | 
| 915 |  |  | using group-based cutoffs. The {\sc sf} method likewise shows | 
| 916 |  |  | improvement in the undamped and lightly damped cases. | 
| 917 | gezelter | 2653 |  | 
| 918 | chrisfen | 2595 | \begin{table}[htbp] | 
| 919 | gezelter | 2656 | \centering | 
| 920 |  |  | \caption{Statistical analysis of the angular | 
| 921 |  |  | distributions that the force (upper) and torque (lower) vectors | 
| 922 |  |  | from a given electrostatic method make with their counterparts | 
| 923 |  |  | obtained using the reference Ewald sum.  Calculations were | 
| 924 |  |  | performed both with (Y) and without (N) group based cutoffs and a | 
| 925 |  |  | switching function.  The $\alpha$ values have units of \AA$^{-1}$ | 
| 926 |  |  | and the variance values have units of degrees$^2$.} | 
| 927 |  |  |  | 
| 928 | chrisfen | 2599 | \begin{tabular}{@{} ccrrrrrrrr @{}} | 
| 929 | chrisfen | 2595 | \\ | 
| 930 |  |  | \toprule | 
| 931 |  |  | & & \multicolumn{4}{c}{Shifted Potential} & \multicolumn{4}{c}{Shifted Force} \\ | 
| 932 |  |  | \cmidrule(lr){3-6} | 
| 933 |  |  | \cmidrule(l){7-10} | 
| 934 | chrisfen | 2599 | $R_\textrm{c}$ & Groups & $\alpha = 0$ & $\alpha = 0.1$ & $\alpha = 0.2$ & $\alpha = 0.3$ & $\alpha = 0$ & $\alpha = 0.1$ & $\alpha = 0.2$ & $\alpha = 0.3$\\ | 
| 935 | chrisfen | 2595 | \midrule | 
| 936 | chrisfen | 2599 |  | 
| 937 |  |  | 9 \AA   & N & 29.545 & 12.003 & 5.489 & 0.610 & 2.323 & 2.321 & 0.429 & 0.603 \\ | 
| 938 |  |  | & \textbf{Y} & \textbf{2.486} & \textbf{2.160} & \textbf{0.667} & \textbf{0.608} & \textbf{1.768} & \textbf{1.766} & \textbf{0.676} & \textbf{0.609} \\ | 
| 939 |  |  | 12 \AA  & N & 19.381 & 3.097 & 0.190 & 0.608 & 0.920 & 0.736 & 0.133 & 0.612 \\ | 
| 940 |  |  | & \textbf{Y} & \textbf{0.515} & \textbf{0.288} & \textbf{0.127} & \textbf{0.586} & \textbf{0.308} & \textbf{0.249} & \textbf{0.127} & \textbf{0.586} \\ | 
| 941 |  |  | 15 \AA  & N & 12.700 & 1.196 & 0.123 & 0.601 & 0.339 & 0.160 & 0.123 & 0.601 \\ | 
| 942 |  |  | & \textbf{Y} & \textbf{0.228} & \textbf{0.099} & \textbf{0.121} & \textbf{0.598} & \textbf{0.144} & \textbf{0.090} & \textbf{0.121} & \textbf{0.598} \\ | 
| 943 | chrisfen | 2594 |  | 
| 944 | chrisfen | 2595 | \midrule | 
| 945 |  |  |  | 
| 946 | chrisfen | 2599 | 9 \AA   & N & 262.716 & 116.585 & 5.234 & 5.103 & 2.392 & 2.350 & 1.770 & 5.122 \\ | 
| 947 |  |  | & \textbf{Y} & \textbf{2.115} & \textbf{1.914} & \textbf{1.878} & \textbf{5.142} & \textbf{2.076} & \textbf{2.039} & \textbf{1.972} & \textbf{5.146} \\ | 
| 948 |  |  | 12 \AA  & N & 129.576 & 25.560 & 1.369 & 5.080 & 0.913 & 0.790 & 1.362 & 5.124 \\ | 
| 949 |  |  | & \textbf{Y} & \textbf{0.810} & \textbf{0.685} & \textbf{1.352} & \textbf{5.082} & \textbf{0.765} & \textbf{0.714} & \textbf{1.360} & \textbf{5.082} \\ | 
| 950 |  |  | 15 \AA  & N & 87.275 & 4.473 & 1.271 & 5.000 & 0.372 & 0.312 & 1.271 & 5.000 \\ | 
| 951 |  |  | & \textbf{Y} & \textbf{0.282} & \textbf{0.294} & \textbf{1.272} & \textbf{4.999} & \textbf{0.324} & \textbf{0.318} & \textbf{1.272} & \textbf{4.999} \\ | 
| 952 | chrisfen | 2595 |  | 
| 953 |  |  | \bottomrule | 
| 954 |  |  | \end{tabular} | 
| 955 | chrisfen | 2601 | \label{tab:groupAngle} | 
| 956 | chrisfen | 2595 | \end{table} | 
| 957 |  |  |  | 
| 958 | gezelter | 2656 | One additional trend in table \ref{tab:groupAngle} is that the | 
| 959 |  |  | $\sigma^2$ values for both {\sc sp} and {\sc sf} converge as $\alpha$ | 
| 960 |  |  | increases, something that is more obvious with group-based cutoffs. | 
| 961 |  |  | The complimentary error function inserted into the potential weakens | 
| 962 |  |  | the electrostatic interaction as the value of $\alpha$ is increased. | 
| 963 |  |  | However, at larger values of $\alpha$, it is possible to overdamp the | 
| 964 |  |  | electrostatic interaction and to remove it completely.  Kast | 
| 965 | gezelter | 2653 | \textit{et al.}  developed a method for choosing appropriate $\alpha$ | 
| 966 |  |  | values for these types of electrostatic summation methods by fitting | 
| 967 |  |  | to $g(r)$ data, and their methods indicate optimal values of 0.34, | 
| 968 |  |  | 0.25, and 0.16 \AA$^{-1}$ for cutoff values of 9, 12, and 15 \AA\ | 
| 969 |  |  | respectively.\cite{Kast03} These appear to be reasonable choices to | 
| 970 |  |  | obtain proper MC behavior (Fig. \ref{fig:delE}); however, based on | 
| 971 |  |  | these findings, choices this high would introduce error in the | 
| 972 | gezelter | 2656 | molecular torques, particularly for the shorter cutoffs.  Based on our | 
| 973 |  |  | observations, empirical damping up to 0.2 \AA$^{-1}$ is beneficial, | 
| 974 |  |  | but damping may be unnecessary when using the {\sc sf} method. | 
| 975 | chrisfen | 2595 |  | 
| 976 | chrisfen | 2638 | \subsection{Short-Time Dynamics: Velocity Autocorrelation Functions of NaCl Crystals} | 
| 977 | chrisfen | 2601 |  | 
| 978 | gezelter | 2653 | Zahn {\it et al.} investigated the structure and dynamics of water | 
| 979 |  |  | using eqs. (\ref{eq:ZahnPot}) and | 
| 980 |  |  | (\ref{eq:WolfForces}).\cite{Zahn02,Kast03} Their results indicated | 
| 981 |  |  | that a method similar (but not identical with) the damped {\sc sf} | 
| 982 |  |  | method resulted in properties very similar to those obtained when | 
| 983 |  |  | using the Ewald summation.  The properties they studied (pair | 
| 984 |  |  | distribution functions, diffusion constants, and velocity and | 
| 985 |  |  | orientational correlation functions) may not be particularly sensitive | 
| 986 |  |  | to the long-range and collective behavior that governs the | 
| 987 | gezelter | 2656 | low-frequency behavior in crystalline systems.  Additionally, the | 
| 988 |  |  | ionic crystals are the worst case scenario for the pairwise methods | 
| 989 |  |  | because they lack the reciprocal space contribution contained in the | 
| 990 |  |  | Ewald summation. | 
| 991 | chrisfen | 2601 |  | 
| 992 | gezelter | 2653 | We are using two separate measures to probe the effects of these | 
| 993 |  |  | alternative electrostatic methods on the dynamics in crystalline | 
| 994 |  |  | materials.  For short- and intermediate-time dynamics, we are | 
| 995 |  |  | computing the velocity autocorrelation function, and for long-time | 
| 996 |  |  | and large length-scale collective motions, we are looking at the | 
| 997 |  |  | low-frequency portion of the power spectrum. | 
| 998 |  |  |  | 
| 999 | chrisfen | 2601 | \begin{figure} | 
| 1000 |  |  | \centering | 
| 1001 | chrisfen | 2741 | \includegraphics[width = \linewidth]{./vCorrPlot.pdf} | 
| 1002 | gezelter | 2656 | \caption{Velocity autocorrelation functions of NaCl crystals at | 
| 1003 |  |  | 1000 K using {\sc spme}, {\sc sf} ($\alpha$ = 0.0, 0.1, \& 0.2), and {\sc | 
| 1004 | gezelter | 2653 | sp} ($\alpha$ = 0.2). The inset is a magnification of the area around | 
| 1005 |  |  | the first minimum.  The times to first collision are nearly identical, | 
| 1006 |  |  | but differences can be seen in the peaks and troughs, where the | 
| 1007 |  |  | undamped and weakly damped methods are stiffer than the moderately | 
| 1008 | gezelter | 2656 | damped and {\sc spme} methods.} | 
| 1009 | chrisfen | 2638 | \label{fig:vCorrPlot} | 
| 1010 |  |  | \end{figure} | 
| 1011 |  |  |  | 
| 1012 | gezelter | 2656 | The short-time decay of the velocity autocorrelation function through | 
| 1013 |  |  | the first collision are nearly identical in figure | 
| 1014 |  |  | \ref{fig:vCorrPlot}, but the peaks and troughs of the functions show | 
| 1015 |  |  | how the methods differ.  The undamped {\sc sf} method has deeper | 
| 1016 |  |  | troughs (see inset in Fig. \ref{fig:vCorrPlot}) and higher peaks than | 
| 1017 |  |  | any of the other methods.  As the damping parameter ($\alpha$) is | 
| 1018 |  |  | increased, these peaks are smoothed out, and the {\sc sf} method | 
| 1019 |  |  | approaches the {\sc spme} results.  With $\alpha$ values of 0.2 \AA$^{-1}$, | 
| 1020 |  |  | the {\sc sf} and {\sc sp} functions are nearly identical and track the | 
| 1021 |  |  | {\sc spme} features quite well.  This is not surprising because the {\sc sf} | 
| 1022 |  |  | and {\sc sp} potentials become nearly identical with increased | 
| 1023 |  |  | damping.  However, this appears to indicate that once damping is | 
| 1024 |  |  | utilized, the details of the form of the potential (and forces) | 
| 1025 |  |  | constructed out of the damped electrostatic interaction are less | 
| 1026 |  |  | important. | 
| 1027 | chrisfen | 2638 |  | 
| 1028 |  |  | \subsection{Collective Motion: Power Spectra of NaCl Crystals} | 
| 1029 |  |  |  | 
| 1030 | gezelter | 2656 | To evaluate how the differences between the methods affect the | 
| 1031 |  |  | collective long-time motion, we computed power spectra from long-time | 
| 1032 |  |  | traces of the velocity autocorrelation function. The power spectra for | 
| 1033 |  |  | the best-performing alternative methods are shown in | 
| 1034 |  |  | fig. \ref{fig:methodPS}.  Apodization of the correlation functions via | 
| 1035 |  |  | a cubic switching function between 40 and 50 ps was used to reduce the | 
| 1036 |  |  | ringing resulting from data truncation.  This procedure had no | 
| 1037 |  |  | noticeable effect on peak location or magnitude. | 
| 1038 | chrisfen | 2638 |  | 
| 1039 |  |  | \begin{figure} | 
| 1040 |  |  | \centering | 
| 1041 | chrisfen | 2741 | \includegraphics[width = \linewidth]{./spectraSquare.pdf} | 
| 1042 | chrisfen | 2651 | \caption{Power spectra obtained from the velocity auto-correlation | 
| 1043 | gezelter | 2656 | functions of NaCl crystals at 1000 K while using {\sc spme}, {\sc sf} | 
| 1044 |  |  | ($\alpha$ = 0, 0.1, \& 0.2), and {\sc sp} ($\alpha$ = 0.2).  The inset | 
| 1045 |  |  | shows the frequency region below 100 cm$^{-1}$ to highlight where the | 
| 1046 |  |  | spectra differ.} | 
| 1047 | chrisfen | 2610 | \label{fig:methodPS} | 
| 1048 | chrisfen | 2601 | \end{figure} | 
| 1049 |  |  |  | 
| 1050 | gezelter | 2656 | While the high frequency regions of the power spectra for the | 
| 1051 |  |  | alternative methods are quantitatively identical with Ewald spectrum, | 
| 1052 |  |  | the low frequency region shows how the summation methods differ. | 
| 1053 |  |  | Considering the low-frequency inset (expanded in the upper frame of | 
| 1054 |  |  | figure \ref{fig:dampInc}), at frequencies below 100 cm$^{-1}$, the | 
| 1055 |  |  | correlated motions are blue-shifted when using undamped or weakly | 
| 1056 |  |  | damped {\sc sf}.  When using moderate damping ($\alpha = 0.2$ | 
| 1057 |  |  | \AA$^{-1}$) both the {\sc sf} and {\sc sp} methods give nearly identical | 
| 1058 |  |  | correlated motion to the Ewald method (which has a convergence | 
| 1059 |  |  | parameter of 0.3119 \AA$^{-1}$).  This weakening of the electrostatic | 
| 1060 |  |  | interaction with increased damping explains why the long-ranged | 
| 1061 |  |  | correlated motions are at lower frequencies for the moderately damped | 
| 1062 |  |  | methods than for undamped or weakly damped methods. | 
| 1063 |  |  |  | 
| 1064 |  |  | To isolate the role of the damping constant, we have computed the | 
| 1065 |  |  | spectra for a single method ({\sc sf}) with a range of damping | 
| 1066 |  |  | constants and compared this with the {\sc spme} spectrum. | 
| 1067 |  |  | Fig. \ref{fig:dampInc} shows more clearly that increasing the | 
| 1068 |  |  | electrostatic damping red-shifts the lowest frequency phonon modes. | 
| 1069 |  |  | However, even without any electrostatic damping, the {\sc sf} method | 
| 1070 |  |  | has at most a 10 cm$^{-1}$ error in the lowest frequency phonon mode. | 
| 1071 |  |  | Without the {\sc sf} modifications, an undamped (pure cutoff) method | 
| 1072 |  |  | would predict the lowest frequency peak near 325 cm$^{-1}$.  {\it | 
| 1073 |  |  | Most} of the collective behavior in the crystal is accurately captured | 
| 1074 |  |  | using the {\sc sf} method.  Quantitative agreement with Ewald can be | 
| 1075 |  |  | obtained using moderate damping in addition to the shifting at the | 
| 1076 |  |  | cutoff distance. | 
| 1077 |  |  |  | 
| 1078 | chrisfen | 2601 | \begin{figure} | 
| 1079 |  |  | \centering | 
| 1080 | chrisfen | 2741 | \includegraphics[width = \linewidth]{./increasedDamping.pdf} | 
| 1081 | gezelter | 2656 | \caption{Effect of damping on the two lowest-frequency phonon modes in | 
| 1082 | chrisfen | 2667 | the NaCl crystal at 1000~K.  The undamped shifted force ({\sc sf}) | 
| 1083 | gezelter | 2656 | method is off by less than 10 cm$^{-1}$, and increasing the | 
| 1084 |  |  | electrostatic damping to 0.25 \AA$^{-1}$ gives quantitative agreement | 
| 1085 |  |  | with the power spectrum obtained using the Ewald sum.  Overdamping can | 
| 1086 |  |  | result in underestimates of frequencies of the long-wavelength | 
| 1087 |  |  | motions.} | 
| 1088 | chrisfen | 2601 | \label{fig:dampInc} | 
| 1089 |  |  | \end{figure} | 
| 1090 |  |  |  | 
| 1091 | chrisfen | 2575 | \section{Conclusions} | 
| 1092 |  |  |  | 
| 1093 | chrisfen | 2620 | This investigation of pairwise electrostatic summation techniques | 
| 1094 | gezelter | 2656 | shows that there are viable and computationally efficient alternatives | 
| 1095 |  |  | to the Ewald summation.  These methods are derived from the damped and | 
| 1096 |  |  | cutoff-neutralized Coulombic sum originally proposed by Wolf | 
| 1097 |  |  | \textit{et al.}\cite{Wolf99} In particular, the {\sc sf} | 
| 1098 |  |  | method, reformulated above as eqs. (\ref{eq:DSFPot}) and | 
| 1099 |  |  | (\ref{eq:DSFForces}), shows a remarkable ability to reproduce the | 
| 1100 |  |  | energetic and dynamic characteristics exhibited by simulations | 
| 1101 |  |  | employing lattice summation techniques.  The cumulative energy | 
| 1102 |  |  | difference results showed the undamped {\sc sf} and moderately damped | 
| 1103 |  |  | {\sc sp} methods produced results nearly identical to {\sc spme}.  Similarly | 
| 1104 |  |  | for the dynamic features, the undamped or moderately damped {\sc sf} | 
| 1105 |  |  | and moderately damped {\sc sp} methods produce force and torque vector | 
| 1106 |  |  | magnitude and directions very similar to the expected values.  These | 
| 1107 |  |  | results translate into long-time dynamic behavior equivalent to that | 
| 1108 |  |  | produced in simulations using {\sc spme}. | 
| 1109 | chrisfen | 2604 |  | 
| 1110 | gezelter | 2656 | As in all purely-pairwise cutoff methods, these methods are expected | 
| 1111 |  |  | to scale approximately {\it linearly} with system size, and they are | 
| 1112 |  |  | easily parallelizable.  This should result in substantial reductions | 
| 1113 |  |  | in the computational cost of performing large simulations. | 
| 1114 |  |  |  | 
| 1115 | chrisfen | 2620 | Aside from the computational cost benefit, these techniques have | 
| 1116 |  |  | applicability in situations where the use of the Ewald sum can prove | 
| 1117 | gezelter | 2656 | problematic.  Of greatest interest is their potential use in | 
| 1118 |  |  | interfacial systems, where the unmodified lattice sum techniques | 
| 1119 |  |  | artificially accentuate the periodicity of the system in an | 
| 1120 |  |  | undesirable manner.  There have been alterations to the standard Ewald | 
| 1121 |  |  | techniques, via corrections and reformulations, to compensate for | 
| 1122 |  |  | these systems; but the pairwise techniques discussed here require no | 
| 1123 |  |  | modifications, making them natural tools to tackle these problems. | 
| 1124 |  |  | Additionally, this transferability gives them benefits over other | 
| 1125 |  |  | pairwise methods, like reaction field, because estimations of physical | 
| 1126 |  |  | properties (e.g. the dielectric constant) are unnecessary. | 
| 1127 | chrisfen | 2605 |  | 
| 1128 | gezelter | 2656 | If a researcher is using Monte Carlo simulations of large chemical | 
| 1129 |  |  | systems containing point charges, most structural features will be | 
| 1130 |  |  | accurately captured using the undamped {\sc sf} method or the {\sc sp} | 
| 1131 |  |  | method with an electrostatic damping of 0.2 \AA$^{-1}$.  These methods | 
| 1132 |  |  | would also be appropriate for molecular dynamics simulations where the | 
| 1133 |  |  | data of interest is either structural or short-time dynamical | 
| 1134 |  |  | quantities.  For long-time dynamics and collective motions, the safest | 
| 1135 |  |  | pairwise method we have evaluated is the {\sc sf} method with an | 
| 1136 |  |  | electrostatic damping between 0.2 and 0.25 | 
| 1137 |  |  | \AA$^{-1}$. | 
| 1138 | chrisfen | 2605 |  | 
| 1139 | gezelter | 2656 | We are not suggesting that there is any flaw with the Ewald sum; in | 
| 1140 |  |  | fact, it is the standard by which these simple pairwise sums have been | 
| 1141 |  |  | judged.  However, these results do suggest that in the typical | 
| 1142 |  |  | simulations performed today, the Ewald summation may no longer be | 
| 1143 |  |  | required to obtain the level of accuracy most researchers have come to | 
| 1144 |  |  | expect. | 
| 1145 |  |  |  | 
| 1146 | chrisfen | 2575 | \section{Acknowledgments} | 
| 1147 | gezelter | 2656 | Support for this project was provided by the National Science | 
| 1148 |  |  | Foundation under grant CHE-0134881.  The authors would like to thank | 
| 1149 |  |  | Steve Corcelli and Ed Maginn for helpful discussions and comments. | 
| 1150 |  |  |  | 
| 1151 | chrisfen | 2594 | \newpage | 
| 1152 |  |  |  | 
| 1153 | gezelter | 2617 | \bibliographystyle{jcp2} | 
| 1154 | chrisfen | 2575 | \bibliography{electrostaticMethods} | 
| 1155 |  |  |  | 
| 1156 |  |  |  | 
| 1157 |  |  | \end{document} |