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

Comparing trunk/multipoleSFPaper/multipoleSFPaper.tex (file contents):
Revision 3065 by chrisfen, Tue Oct 24 06:27:31 2006 UTC vs.
Revision 3066 by gezelter, Tue Oct 24 16:11:42 2006 UTC

# Line 1 | Line 1
1   %\documentclass[prb,aps,twocolumn,tabularx]{revtex4}
2 < \documentclass[12pt]{article}
2 > \documentclass[11pt]{article}
3   \usepackage{tabls}
4   \usepackage{endfloat}
5   \usepackage[tbtags]{amsmath}
# Line 20 | Line 20
20   \brokenpenalty=10000
21   \renewcommand{\baselinestretch}{1.2}
22   \renewcommand\citemid{\ } % no comma in optional reference note
23 + \AtBeginDelayedFloats{\renewcommand{\baselinestretch}{2}} %doublespace captions
24 + \let\Caption\caption
25 + \renewcommand\caption[1]{%
26 +        \Caption[#1]{}%
27 + }
28 + \setlength{\abovecaptionskip}{1.2in}
29 + \setlength{\belowcaptionskip}{1.2in}
30  
31   \begin{document}
32  
33 < \title{Pairwise electrostatic corrections for monopolar and multipolar systems}
33 > \title{Pairwise Alternatives to the Ewald Sum: Applications
34 > and Extension to Point Multipoles}
35  
36   \author{Christopher J. Fennell and J. Daniel Gezelter \\
37   Department of Chemistry and Biochemistry\\ University of Notre Dame\\
# Line 35 | Line 43 | Notre Dame, Indiana 46556}
43   %\doublespacing
44  
45   \begin{abstract}
46 + The damped, shifted-force electrostatic potential has been shown to
47 + give nearly quantitative agreement with smooth particle mesh Ewald for
48 + energy differences between configurations as well as for atomic force
49 + and molecular torque vectors.  In this paper, we extend this technique
50 + to handle interactions between electrostatic multipoles.  We also
51 + investigate the effects of damped and shifted electrostatics on the
52 + static, thermodynamic, and dynamic properties of liquid water and
53 + various polymorphs of ice.  We provide a way of choosing the optimal
54 + damping strength for a given cutoff radius that reproduces the static
55 + dielectric constant in a variety of water models.
56   \end{abstract}
57  
58   %\narrowtext
# Line 45 | Line 63 | Over the past several years, there has been increased
63  
64   \section{Introduction}
65  
66 < Over the past several years, there has been increased interest in
67 < pairwise methods for correcting electrostatic interaction in computer
66 > Over the past several years, there has been increasing interest in
67 > pairwise methods for correcting electrostatic interactions in computer
68   simulations of condensed molecular
69 < systems.\cite{Wolf99,Zahn02,Kast03,Fennell06}
69 > systems.\cite{Wolf99,Zahn02,Kast03,Beck05,Ma05,Fennell06} These
70 > techniques were developed from the observations and efforts of Wolf
71 > {\it et al.}  and their work towards an $\mathcal{O}(N)$ Coulombic
72 > sum.\cite{Wolf99} Wolf's method of cutoff neutralization is able to
73 > obtain excellent agreement with Madelung energies in ionic
74 > crystals.\cite{Wolf99}
75  
76 < \section{Shifted Force Pairwise Electrostatic Correction}
77 <
78 < In our previous look at pairwise electrostatic correction methods, we
79 < described the undamped and damped shifted potential {\sc sp} and
80 < shifted force {\sc sf} techniques.\cite{Fennell06} These techniques
81 < were developed from the observations and efforts of Wolf {\it et al.}
82 < and their work towards an $\mathcal{O}(N)$ Coulombic sum.\cite{Wolf99}
83 < The potential for the damped form of the {\sc sp} method is given by
76 > In a recent paper, we showed that simple modifications to Wolf's
77 > method could give nearly quantitative agreement with smooth particle
78 > mesh Ewald (SPME) for quantities of interest in Monte Carlo
79 > (i.e. configurational energy differences) and Molecular Dynamics
80 > (i.e. atomic force and molecular torque vectors).\cite{Fennell06} We
81 > described the undamped and damped shifted potential (SP) and shifted
82 > force (SF) techniques. The potential for damped form of the SF method
83 > is given by
84   \begin{equation}
62 V_{\textrm{DSP}}(r) = q_iq_j\left(\frac{\textrm{erfc}\left(\alpha r\right)}{r}
63 - \frac{\textrm{erfc}\left(\alpha R_\textrm{c}\right)}{R_\textrm{c}}\right)
64 \quad r\leqslant R_\textrm{c},
65 \label{eq:DSPPot}
66 \end{equation}
67 while the damped form of the {\sc sf} method is given by
68 \begin{equation}
85   \begin{split}
86   V_\mathrm{DSF}(r) = q_iq_j\Biggr{[}&
87   \frac{\mathrm{erfc}\left(\alpha r\right)}{r}
# Line 78 | Line 94 | Overall, the damped {\sc sf} method is considered an i
94   \label{eq:DSFPot}
95   \end{split}
96   \end{equation}
97 < Overall, the damped {\sc sf} method is considered an improvement over
98 < the {\sc sp} method because there is no discontinuity in the forces as
83 < particles move out of the cutoff radius ($R_\textrm{c}$). This is
84 < accomplished by shifting the forces (and potential) to zero at
85 < $R_\textrm{c}$. This is analogous to neutralizing the charge and the
86 < force effect of the charges within $R_\textrm{c}$.
97 > (In this potential and in all electrostatic quantities that follow,
98 > the standard $4 \pi \epsilon_{0}$ has been omitted for clarity.)
99  
100 < As we mentioned in the previous study, to complete the charge
101 < neutralization procedure, a self-neutralization term needs to be
102 < included in the potential. This term exists outside the pair-loop, and
103 < can be thought of as a charge opposite in sign and equal in magnitude
104 < to the central particle, spread out over the entire cutoff
105 < sphere. This term is calculated via a single loop over all charges in
106 < the system. For the undamped case, the self-term can be written as
100 > The damped SF method is an improvement over the SP method because
101 > there is no discontinuity in the forces as particles move out of the
102 > cutoff radius ($R_\textrm{c}$). This is accomplished by shifting the
103 > forces (and potential) to zero at $R_\textrm{c}$. This is analogous to
104 > neutralizing the charge as well as the force effect of the charges
105 > within $R_\textrm{c}$.
106 >
107 > To complete the charge neutralization procedure, a self-neutralization
108 > term needs to be included in the potential. This term is constant (as
109 > long as the charges and cutoff radius do not change), and exists
110 > outside the normal pair-loop.  It can be thought of as a contribution
111 > from a charge opposite in sign and equal in magnitude to the central
112 > charge, but which has been spread out over the surface of the cutoff
113 > sphere.  This term is calculated via a single loop over all charges in
114 > the system.  For the undamped case, the self term can be written as
115   \begin{equation}
116 < V_\textrm{self} = \sum_{i=1}^N\frac{q_iq_i}{2R_\textrm{c}},
116 > V_\textrm{self} = \frac{1}{2 R_\textrm{c}} \sum_{i=1}^N q_i^{2},
117   \label{eq:selfTerm}
118   \end{equation}
119   while for the damped case it can be written as
120   \begin{equation}
121 < V_\textrm{Dself} = \sum_{i=1}^Nq_iq_i\left(\frac{\alpha}{\sqrt{\pi}}
122 <        + \frac{\textrm{erfc}(\alpha R_\textrm{c})}{2R_\textrm{c}}\right).
121 > V_\textrm{self} = \left(\frac{\alpha}{\sqrt{\pi}}
122 >        + \frac{\textrm{erfc}(\alpha
123 > R_\textrm{c})}{2R_\textrm{c}}\right) \sum_{i=1}^N q_i^{2}.
124   \label{eq:dampSelfTerm}
125   \end{equation}
126   The first term within the parentheses of equation
127 < (\ref{eq:dampSelfTerm}) is identical to the self-term in the Ewald
127 > (\ref{eq:dampSelfTerm}) is identical to the self term in the Ewald
128   summation, and comes from the utilization of the complimentary error
129   function for electrostatic damping.\cite{deLeeuw80,Wolf99}
130  
131 < \section{Multipolar Electrostatic Damping}\label{sec:dampingMultipoles}
131 > The SF, SP, and Wolf methods operate by neutralizing the total charge
132 > contained within the cutoff sphere surrounding each particle.  This is
133 > accomplished by creating image charges on the surface of the cutoff
134 > sphere for each pair interaction computed within the sphere.  The
135 > damping function applied to the potential is also an important method
136 > for accelerating convergence.  In the case of systems involving
137 > electrostatic distributions of higher order than point charges, the
138 > question remains: How will the shifting and damping need to be
139 > modified in order to accommodate point multipoles?
140  
141 < As stated above, the {\sc sf} method operates by neutralizing the
142 < charge pair force interactions (and potential) within the cutoff
114 < sphere with shifting and by damping the electrostatic
115 < interactions. Now we would like to consider an extension of these
116 < techniques to include point multipole interactions. How will the
117 < shifting and damping need to be modified in order to accommodate point
118 < multipoles?
141 > \section{Electrostatic Damping for Point
142 > Multipoles}\label{sec:dampingMultipoles}
143  
144 < Of the two component techniques, the easiest to adapt is
145 < shifting. Shifting is employed to neutralize the cutoff sphere;
146 < however, in a system composed purely of point multipoles, the cutoff
123 < sphere is already neutralized. This means that shifting is not
124 < necessary between point multipoles. In a mixed system of monopoles and
125 < multipoles, the undamped {\sc sf} potential needs only to shift the
126 < force terms of the monopole and smoothly truncate the multipole
127 < interactions with a switching function. The switching function is
128 < required in order to conserve energy, because a discontinuity will
129 < exist in both the potential and forces at $R_\textrm{c}$ in the
130 < absence of shifting terms.
144 > To apply the SF method for systems involving point multipoles, we
145 > consider separately the two techniques (shifting and damping) which
146 > contribute to the effectiveness of the DSF potential.
147  
148 < If we want to dampen the {\sc sf} potential, then we need to
149 < incorporate the complimentary error function term into the multipole
150 < potentials. The most direct way to do this is by replacing $r^{-1}$
151 < with erfc$(\alpha r)\cdot r^{-1}$ in the multipole expansion. In the
152 < multipole expansion, rather than considering only the interactions
153 < between single point charges, the electrostatic interaction is
154 < reformulated such that it describes the interaction between charge
155 < distributions about central sites of the respective sets of
156 < charges.\cite{Hirschfelder67} This procedure is what leads to the
157 < familiar charge-dipole,
148 > As noted above, shifting the potential and forces is employed to
149 > neutralize the total charge contained within each cutoff sphere;
150 > however, in a system composed purely of point multipoles, each cutoff
151 > sphere is already neutral, so shifting becomes unnecessary.
152 >
153 > In a mixed system of charges and multipoles, the undamped SF potential
154 > needs only to shift the force terms between charges and smoothly
155 > truncate the multipolar interactions with a switching function.  The
156 > switching function is required for energy conservation, because a
157 > discontinuity will exist in both the potential and forces at
158 > $R_\textrm{c}$ in the absence of shifting terms.
159 >
160 > To damp the SF potential for point multipoles, we need to incorporate
161 > the complimentary error function term into the standard forms of the
162 > multipolar potentials.  We can determine the necessary damping
163 > functions by replacing $1/r$ with $\mathrm{erfc}(\alpha r)/r$ in the
164 > multipole expansion.  This procedure quickly becomes quite complex
165 > with ``two-center'' systems, like the dipole-dipole potential, and is
166 > typically approached using spherical harmonics.\cite{Hirschfelder67} A
167 > simpler method for determining damped multipolar interaction
168 > potentials arises when we adopt the tensor formalism described by
169 > Stone.\cite{Stone02}
170 >
171 > The tensor formalism for electrostatic interactions involves obtaining
172 > the multipole interactions from successive gradients of the monopole
173 > potential. Thus, tensors of rank zero through two are
174   \begin{equation}
175 < V_\textrm{cd} = -q_i\frac{\boldsymbol{\mu}_j\cdot\mathbf{r}_{ij}}{r^3_{ij}}
176 <                = -q_i\mu_j\frac{\cos\theta}{r^2_{ij}},
145 < \label{eq:chargeDipole}
175 > T = \frac{1}{r_{ij}},
176 > \label{eq:tensorRank1}
177   \end{equation}
147 and dipole-dipole,
178   \begin{equation}
179 < V_\textrm{dd} = 3\frac{(\boldsymbol{\mu}_i\cdot\mathbf{r}_{ij})
180 <                        (\boldsymbol{\mu}_j\cdot\mathbf{r}_{ij})}{r^5_{ij}} -
151 <                \frac{\boldsymbol{\mu}_i\cdot\boldsymbol{\mu}_j}{r^3_{ij}},
152 < \label{eq:dipoleDipole}
179 > T_\alpha = \nabla_\alpha \frac{1}{r_{ij}},
180 > \label{eq:tensorRank2}
181   \end{equation}
182 < interaction potentials.
182 > \begin{equation}
183 > T_{\alpha\beta} = \nabla_\alpha\nabla_\beta \frac{1}{r_{ij}},
184 > \label{eq:tensorRank3}
185 > \end{equation}
186 > where the form of the first tensor is the charge-charge potential, the
187 > second gives the charge-dipole potential, and the third gives the
188 > charge-quadrupole and dipole-dipole potentials.\cite{Stone02} Since
189 > the force is $-\nabla V$, the forces for each potential come from the
190 > next higher tensor.
191  
192 < Using the charge-dipole interaction as an example, if we insert
193 < erfc$(\alpha r)\cdot r^{-1}$ in place of $r^{-1}$, a damped
194 < charge-dipole results,
192 > As one would do with the multipolar expansion, we can replace $r^{-1}$
193 > with $\mathrm{erfc}(\alpha r)\cdot r^{-1}$ to obtain damped forms of the
194 > electrostatic potentials.  Equation \ref{eq:tensorRank2} generates a
195 > damped charge-dipole potential,
196   \begin{equation}
197   V_\textrm{Dcd} = -q_i\mu_j\frac{\cos\theta}{r^2_{ij}} c_1(r_{ij}),
198   \label{eq:dChargeDipole}
# Line 166 | Line 203 | Thus, $c_1(r_{ij})$ is the resulting damping term that
203                  + \textrm{erfc}(\alpha r_{ij}).
204   \label{eq:c1Func}
205   \end{equation}
169 Thus, $c_1(r_{ij})$ is the resulting damping term that modifies the
170 standard charge-dipole potential (Eq. (\ref{eq:chargeDipole})).  Note
171 that this damping term is dependent upon distance and not upon
172 orientation, and that it is acting on what was originally an
173 $r^{-3}$ function. By writing the damped form in this manner, we
174 can collect the damping into one function and apply it to the original
175 potential when damping is desired. This works well for potentials that
176 have only one $r^{-n}$ term (where $n$ is an odd positive integer);
177 but in the case of the dipole-dipole potential, there is one part
178 dependent on $r^{-3}$ and another dependent on $r^{-5}$. In order to
179 properly damping this potential, each of these parts is dampened with
180 separate damping functions. We can determine the necessary damping
181 functions by continuing with the multipole expansion; however, it
182 quickly becomes more complex with ``two-center'' systems, like the
183 dipole-dipole potential, and is typically approached with a spherical
184 harmonic formalism.\cite{Hirschfelder67} A simpler method for
185 determining these functions arises from adopting the tensor formalism
186 for expressing the electrostatic interactions.\cite{Stone02}
206  
207 < The tensor formalism for electrostatic interactions involves obtaining
189 < the multipole interactions from successive gradients of the monopole
190 < potential. Thus, tensors of rank one through three are
191 < \begin{equation}
192 < T = \frac{1}{4\pi\epsilon_0r_{ij}},
193 < \label{eq:tensorRank1}
194 < \end{equation}
207 > Equation \ref{eq:tensorRank3} generates a damped dipole-dipole potential,
208   \begin{equation}
209 < T_\alpha = \frac{1}{4\pi\epsilon_0}\nabla_\alpha \frac{1}{r_{ij}},
210 < \label{eq:tensorRank2}
209 > V_\textrm{Ddd} = 3\frac{(\boldsymbol{\mu}_i\cdot\mathbf{r}_{ij})
210 >                (\boldsymbol{\mu}_j\cdot\mathbf{r}_{ij})}{r^5_{ij}}
211 >                c_2(r_{ij}) -
212 >                \frac{\boldsymbol{\mu}_i\cdot\boldsymbol{\mu}_j}{r^3_{ij}}
213 >                c_1(r_{ij}),
214 > \label{eq:dampDipoleDipole}
215   \end{equation}
216 + where
217   \begin{equation}
200 T_{\alpha\beta} = \frac{1}{4\pi\epsilon_0}
201                        \nabla_\alpha\nabla_\beta \frac{1}{r_{ij}},
202 \label{eq:tensorRank3}
203 \end{equation}
204 where the form of the first tensor gives the monopole-monopole
205 potential, the second gives the monopole-dipole potential, and the
206 third gives the monopole-quadrupole and dipole-dipole
207 potentials.\cite{Stone02} Since the force is $-\nabla V$, the forces
208 for each potential come from the next higher tensor.
209
210 To obtain the damped electrostatic forms, we replace $r^{-1}$ with
211 erfc$(\alpha r)\cdot r^{-1}$. Equation \ref{eq:tensorRank2} generates
212 $c_1(r_{ij})$, just like the multipole expansion, while equation
213 \ref{eq:tensorRank3} generates $c_2(r_{ij})$, where
214 \begin{equation}
218   c_2(r_{ij}) = \frac{4\alpha^3r^3_{ij}e^{-\alpha^2r^2_{ij}}}{3\sqrt{\pi}}
219                  + \frac{2\alpha r_{ij}e^{-\alpha^2r^2_{ij}}}{\sqrt{\pi}}
220                  + \textrm{erfc}(\alpha r_{ij}).
# Line 237 | Line 240 | is similar in form to those obtained by researchers fo
240   \label{eq:doubleFactorial}              
241   \end{equation}
242   and $c_0(r_{ij})$ is erfc$(\alpha r_{ij})$. This generating function
243 < is similar in form to those obtained by researchers for the
244 < application of the Ewald sum to
243 > is similar in form to those obtained by Smith and Aguado and Madden
244 > for the application of the Ewald sum to
245   multipoles.\cite{Smith82,Smith98,Aguado03}
246  
247   Returning to the dipole-dipole example, the potential consists of a
248   portion dependent upon $r^{-5}$ and another dependent upon
249 < $r^{-3}$. In the damped dipole-dipole potential,
247 < \begin{equation}
248 < V_\textrm{Ddd} = 3\frac{(\boldsymbol{\mu}_i\cdot\mathbf{r}_{ij})
249 <                (\boldsymbol{\mu}_j\cdot\mathbf{r}_{ij})}{r^5_{ij}}
250 <                c_2(r_{ij}) -
251 <                \frac{\boldsymbol{\mu}_i\cdot\boldsymbol{\mu}_j}{r^3_{ij}}
252 <                c_1(r_{ij}),
253 < \label{eq:dampDipoleDipole}
254 < \end{equation}
249 > $r^{-3}$.
250   $c_2(r_{ij})$ and $c_1(r_{ij})$ dampen these two parts
251 < respectively. The forces for the damped dipole-dipole interaction,
251 > respectively. The forces for the damped dipole-dipole interaction, are
252 > obtained from the next higher tensor, $T_{\alpha \beta \gamma}$,
253   \begin{equation}
254   \begin{split}
255   F_\textrm{Ddd} = &15\frac{(\boldsymbol{\mu}_i\cdot\mathbf{r}_{ij})
# Line 266 | Line 262 | rely on higher order damping functions because we perf
262   \end{split}
263   \label{eq:dampDipoleDipoleForces}
264   \end{equation}
265 < rely on higher order damping functions because we perform another
266 < gradient operation. In this manner, we can dampen higher order
267 < multipolar interactions along with the monopole interactions, allowing
268 < us to include multipoles in simulations involving damped electrostatic
273 < interactions.
265 > Using the tensor formalism, we can dampen higher order multipolar
266 > interactions using the same effective damping function that we use for
267 > charge-charge interactions.  This allows us to include multipoles in
268 > simulations involving damped electrostatic interactions.
269  
270 < \section{Applications of Pairwise Electrostatic Corrections: Water}
270 > \section{Applications of DSF Electrostatics}
271  
272 < In our earlier work, we performed a statistical comparison of both
273 < energetics and dynamics of the {\sc sf} and {\sc sp} methods, showing
274 < that these pairwise correction techniques are nearly equivalent to the
275 < Ewald summation in the simulation of general condensed
276 < systems.\cite{Fennell06} It would be beneficial to have some specific
277 < examples to better illustrate the similarities and differences of the
278 < pairwise (specifically {\sc sf}) and Ewald techniques. To address
279 < this, we decided to perform a detailed analysis involving water
280 < simulations. Water is a good test in that there is a great deal of
281 < detailed experimental information as well as an extensive library of
282 < results from a variety of theoretical models. In choosing a model for
288 < comparing these correction techniques, it is important to consider a
289 < model that has been optimized for use with corrected electrostatics so
290 < that the calculated properties are in better agreement with
291 < experiment.\cite{vanderSpoel98,Horn04,Rick04} For this reason, we
292 < chose the TIP5P-E water model.\cite{Rick04}
272 > Our earlier work on the SF method showed that it can give nearly
273 > quantitive agreement with SPME-derived configurational energy
274 > differences.  The force and torque vectors in identical configurations
275 > are also nearly equivalent under the damped SF potential and
276 > SPME.\cite{Fennell06} Although these measures bode well for the
277 > performance of the SF method in both Monte Carlo and Molecular
278 > Dynamics simulations, it would be helpful to have direct comparisons
279 > of structural, thermodynamic, and dynamic quantities.  To address
280 > this, we performed a detailed analysis of a group of simulations
281 > involving water models (both point charge and multipolar) under a
282 > number of different simulation conditions.
283  
284 < \subsection{TIP5P-E}
284 > To provide the most difficult test for the damped SF method, we have
285 > chosen a model that has been optimized for use with Ewald sum, and
286 > have compared the simulated properties to those computed via Ewald.
287 > It is well known that water models parametrized for use with the Ewald
288 > sum give calculated properties that are in better agreement with
289 > experiment.\cite{vanderSpoel98,Horn04,Rick04} For these reasons, we
290 > chose the TIP5P-E water model for our comparisons involving point
291 > charges.\cite{Rick04}  
292  
293 + The soft sticky dipole (SSD) family of water models is the perfect
294 + test case for the point-multipolar extension of damped electrostatics.
295 + SSD water models are single point molecules that consist of a ``soft''
296 + Lennard-Jones sphere, a point-dipole, and a tetrahedral function for
297 + capturing the hydrogen bonding nature of water - a spherical harmonic
298 + term for water-water tetrahedral interactions and a point-quadrupole
299 + for interactions with surrounding charges. Detailed descriptions of
300 + these models can be found in other
301 + studies.\cite{Liu96b,Chandra99,Tan03,Fennell04}
302 +
303 + In deciding which version of the SSD model to use, we need only
304 + consider that the SF technique was presented as a pairwise replacement
305 + for the Ewald summation. It has been suggested that models
306 + parametrized for the Ewald summation (like TIP5P-E) would be
307 + appropriate for use with a reaction field and vice versa.\cite{Rick04}
308 + Therefore, we decided to test the SSD/RF water model, which was
309 + parametrized for use with a reaction field, with the damped
310 + electrostatic technique to see how the calculated properties change.
311 +
312   The TIP5P-E water model is a variant of Mahoney and Jorgensen's
313   five-point transferable intermolecular potential (TIP5P) model for
314   water.\cite{Mahoney00} TIP5P was developed to reproduce the density
315 < maximum anomaly present in liquid water near 4$^\circ$C. As with many
316 < previous point charge water models (such as ST2, TIP3P, TIP4P, SPC,
317 < and SPC/E), TIP5P was parametrized using a simple cutoff with no
318 < long-range electrostatic
315 > maximum in liquid water near 4$^\circ$C. As with many previous point
316 > charge water models (such as ST2, TIP3P, TIP4P, SPC, and SPC/E), TIP5P
317 > was parametrized using a simple cutoff with no long-range
318 > electrostatic
319   correction.\cite{Stillinger74,Jorgensen83,Berendsen81,Berendsen87}
320   Without this correction, the pressure term on the central particle
321   from the surroundings is missing. When this correction is included,
322   systems of these particles expand to compensate for this added
323   pressure term and under-predict the density of water under standard
324 < conditions. When using any form of long-range electrostatic
325 < correction, it has become common practice to develop or utilize a
326 < reparametrized water model that corrects for this
311 < effect.\cite{vanderSpoel98,Fennell04,Horn04} The TIP5P-E model follows
312 < this practice and was optimized for use with the Ewald
313 < summation.\cite{Rick04} In his publication, Rick preserved the
314 < geometry and point charge magnitudes in TIP5P and focused on altering
315 < the Lennard-Jones parameters to correct the density at 298~K. With the
324 > conditions.  In developing TIP5P-E, Rick preserved the geometry and
325 > point charge magnitudes in TIP5P and focused on altering the
326 > Lennard-Jones parameters to correct the density at 298~K. With the
327   density corrected, he compared common water properties for TIP5P-E
328   using the Ewald sum with TIP5P using a 9~\AA\ cutoff.
329  
330 < In the following sections, we compared these same water properties
331 < calculated from TIP5P-E using the Ewald sum with TIP5P-E using the
332 < {\sc sf} technique.  In the above evaluation of the pairwise
333 < techniques, we observed some flexibility in the choice of parameters.
334 < Because of this, the following comparisons include the {\sc sf}
324 < technique with a 12~\AA\ cutoff and an $\alpha$ = 0.0, 0.1, and
325 < 0.2~\AA$^{-1}$, as well as a 9~\AA\ cutoff with an $\alpha$ =
326 < 0.2~\AA$^{-1}$.
330 > In the following sections, we compare these same properties calculated
331 > from TIP5P-E using the Ewald sum with TIP5P-E using the damped SF
332 > technique.  Our comparisons include the SF technique with a 12~\AA\
333 > cutoff and an $\alpha$ = 0.0, 0.1, and 0.2~\AA$^{-1}$, as well as a
334 > 9~\AA\ cutoff with an $\alpha$ = 0.2~\AA$^{-1}$.
335  
336 < \subsubsection{Density}\label{sec:t5peDensity}
336 > \subsection{The Density Maximum of TIP5P-E}\label{sec:t5peDensity}
337  
330 As stated previously, the property that prompted the development of
331 TIP5P-E was the density at 1 atm.  The density depends upon the
332 internal pressure of the system in the $NPT$ ensemble, and the
333 calculation of the pressure includes a components from both the
334 kinetic energy and the virial. More specifically, the instantaneous
335 molecular pressure ($p(t)$) is given by
336 \begin{equation}
337 p(t) =  \frac{1}{\textrm{d}V}\sum_\mu
338        \left[\frac{\mathbf{P}_{\mu}^2}{M_{\mu}}
339        + \mathbf{R}_{\mu}\cdot\sum_i\mathbf{F}_{\mu i}\right],
340 \label{eq:MolecularPressure}
341 \end{equation}
342 where d is the dimensionality of the system, $V$ is the volume,
343 $\mathbf{P}_{\mu}$ is the momentum of molecule $\mu$, $\mathbf{R}_\mu$
344 is the position of the center of mass ($M_\mu$) of molecule $\mu$, and
345 $\mathbf{F}_{\mu i}$ is the force on atom $i$ of molecule
346 $\mu$.\cite{Melchionna93} The virial term (the right term in the
347 brackets of equation \ref{eq:MolecularPressure}) is directly dependent
348 on the interatomic forces.  Since the {\sc sp} method does not modify
349 the forces, the pressure using {\sc sp} will be identical to that
350 obtained without an electrostatic correction.  The {\sc sf} method
351 does alter the virial component and, by way of the modified pressures,
352 should provide densities more in line with those obtained using the
353 Ewald summation.
354
338   To compare densities, $NPT$ simulations were performed with the same
339   temperatures as those selected by Rick in his Ewald summation
340   simulations.\cite{Rick04} In order to improve statistics around the
# Line 367 | Line 350 | using the Ewald summation (Ref. \citen{Rick04}) and th
350   \begin{figure}
351   \includegraphics[width=\linewidth]{./figures/tip5peDensities.pdf}
352   \caption{Density versus temperature for the TIP5P-E water model when
353 < using the Ewald summation (Ref. \citen{Rick04}) and the {\sc sf} method
354 < with various parameters. The pressure term from the image-charge shell
355 < is larger than that provided by the reciprocal-space portion of the
356 < Ewald summation, leading to slightly lower densities. This effect is
357 < more visible with the 9~\AA\ cutoff, where the image charges exert a
358 < greater force on the central particle. The error bars for the {\sc sf}
359 < methods show the average one-sigma uncertainty of the density
360 < measurement, and this uncertainty is the same for all the {\sc sf}
361 < curves.}
353 > using the Ewald summation (Ref. \citen{Rick04}) and the SF method with
354 > varying cutoff radii and damping coefficients. The pressure term from
355 > the image-charge shell is larger than that provided by the
356 > reciprocal-space portion of the Ewald summation, leading to slightly
357 > lower densities. This effect is more visible with the 9~\AA\ cutoff,
358 > where the image charges exert a greater force on the central
359 > particle. The error bars for the SF methods show the average one-sigma
360 > uncertainty of the density measurement, and this uncertainty is the
361 > same for all the SF curves.}
362   \label{fig:t5peDensities}
363   \end{figure}
364   Figure \ref{fig:t5peDensities} shows the densities calculated for
365 < TIP5P-E using differing electrostatic corrections overlaid on the
366 < experimental values.\cite{CRC80} The densities when using the {\sc sf}
367 < technique are close to, though typically lower than, those calculated
365 > TIP5P-E using differing electrostatic corrections overlaid with the
366 > experimental values.\cite{CRC80} The densities when using the SF
367 > technique are close to, but typically lower than, those calculated
368   using the Ewald summation. These slightly reduced densities indicate
369   that the pressure component from the image charges at R$_\textrm{c}$
370   is larger than that exerted by the reciprocal-space portion of the
371 < Ewald summation. Bringing the image charges closer to the central
371 > Ewald summation.  Bringing the image charges closer to the central
372   particle by choosing a 9~\AA\ R$_\textrm{c}$ (rather than the
373   preferred 12~\AA\ R$_\textrm{c}$) increases the strength of the image
374   charge interactions on the central particle and results in a further
# Line 398 | Line 381 | the damped {\sc sf} sums showing slightly higher densi
381   is distance-dependent, force components from the image charges will be
382   reduced more than those from particles close the the central
383   charge. This effect is visible in figure \ref{fig:t5peDensities} with
384 < the damped {\sc sf} sums showing slightly higher densities; however,
385 < it is apparent that the choice of cutoff radius plays a much more
386 < important role in the resulting densities.
384 > the damped SF sums showing slightly higher densities; however, it is
385 > clear that the choice of cutoff radius plays a much more important
386 > role in the resulting densities.
387  
388 < As a final note, all of the above density calculations were performed
389 < with systems of 512 water molecules. Rick observed a system size
390 < dependence of the computed densities when using the Ewald summation,
391 < most likely due to his tying of the convergence parameter to the box
388 > All of the above density calculations were performed with systems of
389 > 512 water molecules. Rick observed a system size dependence of the
390 > computed densities when using the Ewald summation, most likely due to
391 > his tying of the convergence parameter to the box
392   dimensions.\cite{Rick04} For systems of 256 water molecules, the
393   calculated densities were on average 0.002 to 0.003~g/cm$^3$ lower. A
394   system size of 256 molecules would force the use of a shorter
395 < R$_\textrm{c}$ when using the {\sc sf} technique, and this would also
396 < lower the densities. Moving to larger systems, as long as the
397 < R$_\textrm{c}$ remains at a fixed value, we would expect the densities
398 < to remain constant.
395 > R$_\textrm{c}$ when using the SF technique, and this would also lower
396 > the densities. Moving to larger systems, as long as the R$_\textrm{c}$
397 > remains at a fixed value, we would expect the densities to remain
398 > constant.
399  
400 < \subsubsection{Liquid Structure}\label{sec:t5peLiqStructure}
400 > \subsection{Liquid Structure of TIP5P-E}\label{sec:t5peLiqStructure}
401  
419 A common function considered when developing and comparing water
420 models is the oxygen-oxygen radial distribution function
421 ($g_\textrm{OO}(r)$). The $g_\textrm{OO}(r)$ is the probability of
422 finding a pair of oxygen atoms some distance ($r$) apart relative to a
423 random distribution at the same density.\cite{Allen87} It is
424 calculated via
425 \begin{equation}
426 g_\textrm{OO}(r) = \frac{V}{N^2}\left\langle\sum_i\sum_{j\ne i}
427        \delta(\mathbf{r}-\mathbf{r}_{ij})\right\rangle,
428 \label{eq:GOOofR}
429 \end{equation}
430 where the double sum is over all $i$ and $j$ pairs of $N$ oxygen
431 atoms. The $g_\textrm{OO}(r)$ can be directly compared to X-ray or
432 neutron scattering experiments through the oxygen-oxygen structure
433 factor ($S_\textrm{OO}(k)$) by the following relationship:
434 \begin{equation}
435 S_\textrm{OO}(k) = 1 + 4\pi\rho
436        \int_0^\infty r^2\frac{\sin kr}{kr}g_\textrm{OO}(r)\textrm{d}r.
437 \label{eq:SOOofK}
438 \end{equation}
439 Thus, $S_\textrm{OO}(k)$ is related to the Fourier-Laplace transform
440 of $g_\textrm{OO}(r)$.
441
402   The experimentally determined $g_\textrm{OO}(r)$ for liquid water has
403   been compared in great detail with the various common water models,
404   and TIP5P was found to be in better agreement than other rigid,
405   non-polarizable models.\cite{Sorenson00} This excellent agreement with
406   experiment was maintained when Rick developed TIP5P-E.\cite{Rick04} To
407 < check whether the choice of using the Ewald summation or the {\sc sf}
407 > check whether the choice of using the Ewald summation or the SF
408   technique alters the liquid structure, the $g_\textrm{OO}(r)$s at
409   298~K and 1~atm were determined for the systems compared in the
410   previous section.
# Line 454 | Line 414 | using the {\sc sf} technique, the $g_\textrm{OO}(r)$s
414   \caption{The $g_\textrm{OO}(r)$s calculated for TIP5P-E at 298~K and
415   1atm while using the Ewald summation (Ref. \cite{Rick04}) and the {\sc
416   sf} technique with varying parameters. Even with the reduced densities
417 < using the {\sc sf} technique, the $g_\textrm{OO}(r)$s are essentially
417 > using the SF technique, the $g_\textrm{OO}(r)$s are essentially
418   identical.}
419   \label{fig:t5peGofRs}
420   \end{figure}
# Line 463 | Line 423 | are indistinguishable. These results indicate that the
423   $g_\textrm{OO}(r)$ while using the Ewald summation in
424   figure~\ref{fig:t5peGofRs}. The differences in density do not appear
425   to have any effect on the liquid structure as the $g_\textrm{OO}(r)$s
426 < are indistinguishable. These results indicate that the
426 > are indistinguishable. These results do indicate that
427   $g_\textrm{OO}(r)$ is insensitive to the choice of electrostatic
428   correction.
429  
430 < \subsubsection{Thermodynamic Properties}\label{sec:t5peThermo}
430 > \subsection{Thermodynamic Properties of TIP5P-E}\label{sec:t5peThermo}
431  
432 < In addition to the density, there are a variety of thermodynamic
433 < quantities that can be calculated for water and compared directly to
434 < experimental values. Some of these additional quantities include the
435 < latent heat of vaporization ($\Delta H_\textrm{vap}$), the constant
436 < pressure heat capacity ($C_p$), the isothermal compressibility
437 < ($\kappa_T$), the thermal expansivity ($\alpha_p$), and the static
438 < dielectric constant ($\epsilon$). All of these properties were
439 < calculated for TIP5P-E with the Ewald summation, so they provide a
440 < good set for comparisons involving the {\sc sf} technique.
432 > In addition to the density and structual features of the liquid, there
433 > are a variety of thermodynamic quantities that can be calculated for
434 > water and compared directly to experimental values. Some of these
435 > additional quantities include the latent heat of vaporization ($\Delta
436 > H_\textrm{vap}$), the constant pressure heat capacity ($C_p$), the
437 > isothermal compressibility ($\kappa_T$), the thermal expansivity
438 > ($\alpha_p$), and the static dielectric constant ($\epsilon$). All of
439 > these properties were calculated for TIP5P-E with the Ewald summation,
440 > so they provide a good set for comparisons involving the SF technique.
441  
442   The $\Delta H_\textrm{vap}$ is the enthalpy change required to
443   transform one mole of substance from the liquid phase to the gas
# Line 485 | Line 445 | determined via
445   determined via
446   \begin{equation}
447   \begin{split}
448 < \Delta H_\textrm{vap} &= H_\textrm{gas} - H_\textrm{liq.} \\
449 <                      &= E_\textrm{gas} - E_\textrm{liq.}
450 <                        + p(V_\textrm{gas} - V_\textrm{liq.}) \\
451 <                      &\approx -\frac{\langle U_\textrm{liq.}\rangle}{N}+ RT,
448 > \Delta H_\textrm{vap} &= H_\textrm{gas} - H_\textrm{liq} \\
449 >                      &= E_\textrm{gas} - E_\textrm{liq}
450 >                        + P(V_\textrm{gas} - V_\textrm{liq}) \\
451 >                      &\approx -\frac{\langle U_\textrm{liq}\rangle}{N}+ RT,
452   \end{split}
453   \label{eq:DeltaHVap}
454   \end{equation}
455 < where $E$ is the total energy, $U$ is the potential energy, $p$ is the
455 > where $E$ is the total energy, $U$ is the potential energy, $P$ is the
456   pressure, $V$ is the volume, $N$ is the number of molecules, $R$ is
457   the gas constant, and $T$ is the temperature.\cite{Jorgensen98b} As
458   seen in the last line of equation (\ref{eq:DeltaHVap}), we can
459   approximate $\Delta H_\textrm{vap}$ by using the ideal gas for the gas
460   state. This allows us to cancel the kinetic energy terms, leaving only
461 < the $U_\textrm{liq.}$ term. Additionally, the $pV$ term for the gas is
461 > the $U_\textrm{liq}$ term. Additionally, the $pV$ term for the gas is
462   several orders of magnitude larger than that of the liquid, so we can
463   neglect the liquid $pV$ term.
464  
# Line 508 | Line 468 | C_p    = \left(\frac{\partial H}{\partial T}\right)_{N,p
468   enthalpy in constant pressure simulations via
469   \begin{equation}
470   \begin{split}
471 < C_p     = \left(\frac{\partial H}{\partial T}\right)_{N,p}
471 > C_p     = \left(\frac{\partial H}{\partial T}\right)_{N,P}
472          = \frac{(\langle H^2\rangle - \langle H\rangle^2)}{Nk_BT^2},
473   \end{split}
474   \label{eq:Cp}
# Line 517 | Line 477 | where $k_B$ is Boltzmann's constant.  $\kappa_T$ can b
477   \begin{equation}
478   \begin{split}
479   \kappa_T = \frac{1}{V}\left(\frac{\partial V}{\partial p}\right)_{N,T}
480 <         = \frac{(\langle V^2\rangle_{N,P,T} - \langle V\rangle^2_{N,P,T})}
481 <                {k_BT\langle V\rangle_{N,P,T}},
480 >         = \frac{(\langle V^2\rangle_{NPT} - \langle V\rangle^2_{NPT})}
481 >                {k_BT\langle V\rangle_{NPT}},
482   \end{split}
483   \label{eq:kappa}
484   \end{equation}
# Line 526 | Line 486 | and $\alpha_p$ can be calculated via
486   \begin{equation}
487   \begin{split}
488   \alpha_p        = \frac{1}{V}\left(\frac{\partial V}{\partial T}\right)_{N,P}
489 <                = \frac{(\langle VH\rangle_{N,P,T}
490 <                        - \langle V\rangle_{N,P,T}\langle H\rangle_{N,P,T})}
491 <                        {k_BT^2\langle V\rangle_{N,P,T}}.
489 >                = \frac{(\langle VH\rangle_{NPT}
490 >                        - \langle V\rangle_{NPT}\langle H\rangle_{NPT})}
491 >                        {k_BT^2\langle V\rangle_{NPT}}.
492   \end{split}
493   \label{eq:alpha}
494   \end{equation}
# Line 566 | Line 526 | equal to zero. Since the {\sc sf} method also lacks th
526   however, it converges rather slowly, thus requiring multi-nanosecond
527   simulation times.\cite{Horn04} In the case of tin-foil boundary
528   conditions, the dielectric/surface term of the Ewald summation is
529 < equal to zero. Since the {\sc sf} method also lacks this
529 > equal to zero. Since the SF method also lacks this
530   dielectric/surface term, equation (\ref{eq:staticDielectric}) is still
531   valid for determining static dielectric constants.
532  
# Line 581 | Line 541 | and the {\sc sf} techniques along with the experimenta
541   \centering
542   \includegraphics[width=4.5in]{./figures/t5peThermo.pdf}
543   \caption{Thermodynamic properties for TIP5P-E using the Ewald summation
544 < and the {\sc sf} techniques along with the experimental values. Units
544 > and the SF techniques along with the experimental values. Units
545   for the properties are kcal mol$^{-1}$ for $\Delta H_\textrm{vap}$,
546   cal mol$^{-1}$ K$^{-1}$ for $C_p$, 10$^6$ atm$^{-1}$ for $\kappa_T$,
547   and 10$^5$ K$^{-1}$ for $\alpha_p$. Ewald summation results are from
# Line 593 | Line 553 | As observed for the density in section \ref{sec:t5peDe
553   \label{fig:t5peThermo}
554   \end{figure}
555  
556 < As observed for the density in section \ref{sec:t5peDensity}, the
557 < property trends with temperature seen when using the Ewald summation
558 < are reproduced with the {\sc sf} technique. One noticeable difference
559 < between the properties calculated using the two methods are the lower
560 < $\Delta H_\textrm{vap}$ values when using {\sc sf}. This is to be
561 < expected due to the direct weakening of the electrostatic interaction
562 < through forced neutralization. This results in an increase of the
563 < intermolecular potential producing lower values from equation
564 < (\ref{eq:DeltaHVap}). The slopes of these values with temperature are
565 < similar to that seen using the Ewald summation; however, they are both
566 < steeper than the experimental trend, indirectly resulting in the
607 < inflated $C_p$ values at all temperatures.
556 > For all of the properties computed, the trends with temperature
557 > obtained when using the Ewald summation are reproduced with the SF
558 > technique. One noticeable difference between the properties calculated
559 > using the two methods are the lower $\Delta H_\textrm{vap}$ values
560 > when using SF.  This is to be expected due to the direct weakening of
561 > the electrostatic interaction through forced neutralization. This
562 > results in an increase of the intermolecular potential producing lower
563 > values from equation (\ref{eq:DeltaHVap}). The slopes of these values
564 > with temperature are similar to that seen using the Ewald summation;
565 > however, they are both steeper than the experimental trend, indirectly
566 > resulting in the inflated $C_p$ values at all temperatures.
567  
568   Above the supercooled regime, $C_p$, $\kappa_T$, and $\alpha_p$ values
569   all overlap within error. As indicated for the $\Delta H_\textrm{vap}$
570 < and $C_p$ results discussed in the previous paragraph, the deviations
571 < between experiment and simulation in this region are not the fault of
572 < the electrostatic summation methods but are due to the geometry and
573 < parameters of the TIP5P class of water models. Like most rigid,
574 < non-polarizable, point-charge water models, the density decreases with
575 < temperature at a much faster rate than experiment (see figure
576 < \ref{fig:t5peDensities}). This reduced density leads to the inflated
577 < compressibility and expansivity values at higher temperatures seen
578 < here in figure \ref{fig:t5peThermo}. Incorporation of polarizability
579 < and many-body effects are required in order for water models to
580 < overcome differences between simulation-based and experimentally
622 < determined densities at these higher
570 > and $C_p$ results, the deviations between experiment and simulation in
571 > this region are not the fault of the electrostatic summation methods
572 > but are due to the geometry and parameters of the TIP5P class of water
573 > models. Like most rigid, non-polarizable, point-charge water models,
574 > the density decreases with temperature at a much faster rate than
575 > experiment (see figure \ref{fig:t5peDensities}). This reduced density
576 > leads to the inflated compressibility and expansivity values at higher
577 > temperatures seen here in figure \ref{fig:t5peThermo}. Incorporation
578 > of polarizability and many-body effects are required in order for
579 > water models to overcome differences between simulation-based and
580 > experimentally determined densities at these higher
581   temperatures.\cite{Laasonen93,Donchev06}
582  
583   At temperatures below the freezing point for experimental water, the
584 < differences between {\sc sf} and the Ewald summation results are more
584 > differences between SF and the Ewald summation results are more
585   apparent. The larger $C_p$ and lower $\alpha_p$ values in this region
586   indicate a more pronounced transition in the supercooled regime,
587 < particularly in the case of {\sc sf} without damping. This points to
630 < the onset of a more frustrated or glassy behavior for TIP5P-E at
631 < temperatures below 250~K in the {\sc sf} simulations, indicating that
632 < disorder in the reciprocal-space term of the Ewald summation might act
633 < to loosen up the local structure more than the image-charges in {\sc
634 < sf}. The damped {\sc sf} actually makes a better comparison with
635 < experiment in this region, particularly for the $\alpha_p$ values. The
636 < local interactions in the undamped {\sc sf} technique appear to be too
637 < strong since the property change is much more dramatic than the damped
638 < forms, while the Ewald summation appears to weight the
639 < reciprocal-space interactions at the expense the local interactions,
640 < disagreeing with the experimental results. This observation is
641 < explored further in section \ref{sec:t5peDynamics}.
587 > particularly in the case of SF without damping.
588  
589 + This points to the onset of a more frustrated or glassy behavior for
590 + the undamped and weakly-damped SF simulations of TIP5P-E at
591 + temperatures below 250~K than is seen from the Ewald sum.  Undamped SF
592 + electrostatics has a stronger contribution from nearby charges.
593 + Damping these local interactions or using a reciprocal-space method
594 + makes the water less sensitive to ordering on a short length scale.
595 + We can recover nearly quantitative agreement with the Ewald results by
596 + increasing the damping constant.
597 +
598   The final thermodynamic property displayed in figure
599   \ref{fig:t5peThermo}, $\epsilon$, shows the greatest discrepancy
600 < between the Ewald summation and the {\sc sf} technique (and experiment
601 < for that matter). It is known that the dielectric constant is
602 < dependent upon and quite sensitive to the imposed boundary
603 < conditions.\cite{Neumann80,Neumann83} This is readily apparent in the
604 < converged $\epsilon$ values accumulated for the {\sc sf}
650 < simulations. Lack of a damping function results in dielectric
600 > between the Ewald and SF methods (and with experiment). It is known
601 > that the dielectric constant is dependent upon and is quite sensitive
602 > to the imposed boundary conditions.\cite{Neumann80,Neumann83} This is
603 > readily apparent in the converged $\epsilon$ values accumulated for
604 > the SF simulations. Lack of a damping function results in dielectric
605   constants significantly smaller than those obtained using the Ewald
606   sum. Increasing the damping coefficient to 0.2~\AA$^{-1}$ improves the
607   agreement considerably. It should be noted that the choice of the
608 < ``Ewald coefficient'' value also has a significant effect on the
609 < calculated value when using the Ewald summation. In the simulations of
610 < TIP5P-E with the Ewald sum, this screening parameter was tethered to
611 < the simulation box size (as was the $R_\textrm{c}$).\cite{Rick04} In
612 < general, systems with larger screening parameters reported larger
613 < dielectric constant values, the same behavior we see here with {\sc
614 < sf}; however, the choice of cutoff radius also plays an important
615 < role. This connection is further explored below as optimal damping
662 < coefficients for different choices of $R_\textrm{c}$ are determined
663 < for {\sc sf} in order to best capture the dielectric behavior.
608 > ``Ewald coefficient'' ($\kappa$) and real-space cutoff values also
609 > have a significant effect on the calculated static dielectric constant
610 > when using the Ewald summation. In the simulations of TIP5P-E with the
611 > Ewald sum, this screening parameter was tethered to the simulation box
612 > size (as was the $R_\textrm{c}$).\cite{Rick04} In general, systems
613 > with larger screening parameters reported larger dielectric constant
614 > values, the same behavior we see here with {\sc sf}; however, the
615 > choice of cutoff radius also plays an important role.
616  
617 < \subsubsection{Dielectric Constant}\label{sec:t5peDielectric}
617 > \subsubsection{Dielectric Constants for TIP5P-E and SSD/RF}\label{sec:t5peDielectric}
618  
619 < In the previous, we observed that the choice of damping coefficient
620 < plays a major role in the calculated dielectric constant.  This is not
621 < too surprising given the results for damping parameter influence on
622 < the long-time correlated motions of the NaCl crystal in our previous
623 < study.\cite{Fennell06} The static dielectric constant is calculated
624 < from the long-time fluctuations of the system's accumulated dipole
625 < moment (Eq. (\ref{eq:staticDielectric})), so it is going to be quite
626 < sensitive to the choice of damping parameter.  We would like to choose
627 < optimal damping constants such that any arbitrary choice of cutoff
628 < radius will properly capture the dielectric behavior of the liquid.
619 > In the previous section, we observed that the choice of damping
620 > coefficient plays a major role in the calculated dielectric constant
621 > for the SF method.  Similar damping parameter behavior was observed in
622 > the long-time correlated motions of the NaCl crystal.\cite{Fennell06}
623 > The static dielectric constant is calculated from the long-time
624 > fluctuations of the system's accumulated dipole moment
625 > (Eq. (\ref{eq:staticDielectric})), so it is quite sensitive to the
626 > choice of damping parameter.  Since $\alpha$ is an adjustable
627 > parameter, it would be best to choose optimal damping constants such
628 > that any arbitrary choice of cutoff radius will properly capture the
629 > dielectric behavior of the liquid.
630  
631   In order to find these optimal values, we mapped out the static
632   dielectric constant as a function of both the damping parameter and
633 < cutoff radius.  To calculate the static dielectric constant, we
634 < performed 5~ns $NPT$ calculations on systems of 512 TIP5P-E water
635 < molecules and averaged over the converged region (typically the later
636 < 2.5~ns) of equation (\ref{eq:staticDielectric}).  The selected cutoff
637 < radii ranged from 9, 10, 11, and 12~\AA , and the damping parameter
638 < values ranged from 0.1 to 0.35~\AA$^{-1}$.
633 > cutoff radius for TIP5P-E and for a point-dipolar water model
634 > (SSD/RF).  To calculate the static dielectric constant, we performed
635 > 5~ns $NPT$ calculations on systems of 512 water molecules and averaged
636 > over the converged region (typically the later 2.5~ns) of equation
637 > (\ref{eq:staticDielectric}).  The selected cutoff radii ranged from 9,
638 > 10, 11, and 12~\AA , and the damping parameter values ranged from 0.1
639 > to 0.35~\AA$^{-1}$.
640  
641   \begin{table}
642   \centering
643 < \caption{TIP5P-E dielectric constant as a function of $\alpha$ and
644 < $R_\textrm{c}$. The color scale ranges from blue (10) to red (100).}
643 > \caption{Static dielectric constants for the TIP5P-E and SSD/RF water models at 298~K and 1~atm as a function of damping coefficient $\alpha$ and
644 > cutoff radius $R_\textrm{c}$. The color scale ranges from blue (10) to red (100).}
645   \vspace{6pt}
646 < \begin{tabular}{ lcccc }
646 > \begin{tabular}{ lccccccccc }
647   \toprule
648   \toprule
649 < & \multicolumn{4}{c}{$R_\textrm{c}$ (\AA )} \\
650 < \cmidrule(lr){2-5}
651 < $\alpha$ (\AA$^{-1}$) & 9 & 10 & 11 & 12 \\
649 > & \multicolumn{4}{c}{TIP5P-E} & & \multicolumn{4}{c}{SSD/RF} \\
650 > & \multicolumn{4}{c}{$R_\textrm{c}$ (\AA )} & & \multicolumn{4}{c}{$R_\textrm{c}$ (\AA )} \\
651 > \cmidrule(lr){2-5}  \cmidrule(lr){7-10}
652 > $\alpha$ (\AA$^{-1}$) & 9 & 10 & 11 & 12 & & 9 & 10 & 11 & 12 \\
653   \midrule
654 < 0.35 & \cellcolor[rgb]{1, 0.788888888888889, 0.5} 87.0 & \cellcolor[rgb]{1, 0.695555555555555, 0.5} 91.2 & \cellcolor[rgb]{1, 0.717777777777778, 0.5} 90.2 & \cellcolor[rgb]{1, 0.686666666666667, 0.5} 91.6 \\
655 < & \cellcolor[rgb]{1, 0.892222222222222, 0.5} & \cellcolor[rgb]{1, 0.704444444444444, 0.5} & \cellcolor[rgb]{1, 0.72, 0.5} & \cellcolor[rgb]{1, 0.6666666666667, 0.5} \\
656 < 0.3 & \cellcolor[rgb]{1, 0.995555555555556, 0.5} 77.7 & \cellcolor[rgb]{1, 0.713333333333333, 0.5} 90.4 & \cellcolor[rgb]{1, 0.713333333333333, 0.5} 90.4 & \cellcolor[rgb]{1, 0.646666666666667, 0.5} 93.4 \\
657 < 0.275 & \cellcolor[rgb]{0.653333333333333, 1, 0.5} 61.9 & \cellcolor[rgb]{1, 0.933333333333333, 0.5} 80.5 & \cellcolor[rgb]{1, 0.811111111111111, 0.5} 86.0 & \cellcolor[rgb]{1, 0.766666666666667, 0.5} 88 \\
658 < 0.25 & \cellcolor[rgb]{0.537777777777778, 1, 0.5} 56.7 & \cellcolor[rgb]{0.795555555555555, 1, 0.5} 68.3 & \cellcolor[rgb]{1, 0.966666666666667, 0.5} 79 & \cellcolor[rgb]{1, 0.704444444444445, 0.5} 90.8 \\
659 < 0.225 & \cellcolor[rgb]{0.5, 1, 0.768888888888889} 42.9 & \cellcolor[rgb]{0.566666666666667, 1, 0.5} 58.0 & \cellcolor[rgb]{0.693333333333333, 1, 0.5} 63.7 & \cellcolor[rgb]{1, 0.937777777777778, 0.5} 80.3 \\
660 < 0.2 & \cellcolor[rgb]{0.5, 0.973333333333333, 1} 31.3 & \cellcolor[rgb]{0.5, 1, 0.842222222222222} 39.6 & \cellcolor[rgb]{0.54, 1, 0.5} 56.8 & \cellcolor[rgb]{0.735555555555555, 1, 0.5} 65.6 \\
661 < & \cellcolor[rgb]{0.5, 0.848888888888889, 1} & \cellcolor[rgb]{0.5, 0.973333333333333, 1} & \cellcolor[rgb]{0.5, 1, 0.793333333333333} & \cellcolor[rgb]{0.5, 1, 0.624444444444445} \\
662 < 0.15 & \cellcolor[rgb]{0.5, 0.724444444444444, 1} 20.1 & \cellcolor[rgb]{0.5, 0.788888888888889, 1} 23.0 & \cellcolor[rgb]{0.5, 0.873333333333333, 1} 26.8 & \cellcolor[rgb]{0.5, 1, 0.984444444444444} 33.2 \\
663 < & \cellcolor[rgb]{0.5, 0.696111111111111, 1} & \cellcolor[rgb]{0.5, 0.736333333333333, 1} & \cellcolor[rgb]{0.5, 0.775222222222222, 1} & \cellcolor[rgb]{0.5, 0.860666666666667, 1} \\
664 < 0.1 & \cellcolor[rgb]{0.5, 0.667777777777778, 1} 17.55 & \cellcolor[rgb]{0.5, 0.683777777777778, 1} 18.27 & \cellcolor[rgb]{0.5, 0.677111111111111, 1} 17.97 & \cellcolor[rgb]{0.5, 0.705777777777778, 1} 19.26 \\
654 > 0.35 & \cellcolor[rgb]{1, 0.788888888888889, 0.5} 87.0 & \cellcolor[rgb]{1, 0.695555555555555, 0.5} 91.2 & \cellcolor[rgb]{1, 0.717777777777778, 0.5} 90.2 & \cellcolor[rgb]{1, 0.686666666666667, 0.5} 91.6      & & \cellcolor[rgb]{1, 0.5, 0.5} 116 & \cellcolor[rgb]{1, 0.5, 0.5} 119.2 & \cellcolor[rgb]{1, 0.5, 0.5} 131.4 & \cellcolor[rgb]{1, 0.5, 0.5} 130 \\                                                            
655 > & \cellcolor[rgb]{1, 0.892222222222222, 0.5} & \cellcolor[rgb]{1, 0.704444444444444, 0.5} & \cellcolor[rgb]{1, 0.72, 0.5} & \cellcolor[rgb]{1, 0.6666666666667, 0.5}                                              & & \cellcolor[rgb]{1, 0.5, 0.5} & \cellcolor[rgb]{1, 0.5, 0.5} & \cellcolor[rgb]{1, 0.5, 0.5} & \cellcolor[rgb]{1, 0.5, 0.5} \\                                                                                
656 > 0.3 & \cellcolor[rgb]{1, 0.995555555555556, 0.5} 77.7 & \cellcolor[rgb]{1, 0.713333333333333, 0.5} 90.4 & \cellcolor[rgb]{1, 0.713333333333333, 0.5} 90.4 & \cellcolor[rgb]{1, 0.646666666666667, 0.5} 93.4       & & \cellcolor[rgb]{1, 0.5, 0.5} 100 & \cellcolor[rgb]{1, 0.5, 0.5} 118.8 & \cellcolor[rgb]{1, 0.5, 0.5} 116 & \cellcolor[rgb]{1, 0.5, 0.5} 122 \\                                                              
657 > 0.275 & \cellcolor[rgb]{0.653333333333333, 1, 0.5} 61.9 & \cellcolor[rgb]{1, 0.933333333333333, 0.5} 80.5 & \cellcolor[rgb]{1, 0.811111111111111, 0.5} 86.0 & \cellcolor[rgb]{1, 0.766666666666667, 0.5} 88       & & \cellcolor[rgb]{1, 1, 0.5} 77.5 & \cellcolor[rgb]{1, 0.5, 0.5} 105 & \cellcolor[rgb]{1, 0.5, 0.5} 118 & \cellcolor[rgb]{1, 0.5, 0.5} 125.2 \\                                                              
658 > 0.25 & \cellcolor[rgb]{0.537777777777778, 1, 0.5} 56.7 & \cellcolor[rgb]{0.795555555555555, 1, 0.5} 68.3 & \cellcolor[rgb]{1, 0.966666666666667, 0.5} 79 & \cellcolor[rgb]{1, 0.704444444444445, 0.5} 90.8        & & \cellcolor[rgb]{0.5, 1, 0.582222222222222} 51.3 & \cellcolor[rgb]{1, 0.993333333333333, 0.5} 77.8 & \cellcolor[rgb]{1, 0.522222222222222, 0.5} 99 & \cellcolor[rgb]{1, 0.5, 0.5} 113 \\                    
659 > 0.225 & \cellcolor[rgb]{0.5, 1, 0.768888888888889} 42.9 & \cellcolor[rgb]{0.566666666666667, 1, 0.5} 58.0 & \cellcolor[rgb]{0.693333333333333, 1, 0.5} 63.7 & \cellcolor[rgb]{1, 0.937777777777778, 0.5} 80.3     & & \cellcolor[rgb]{0.5, 0.984444444444444, 1} 31.8 & \cellcolor[rgb]{0.5, 1, 0.586666666666667} 51.1 & \cellcolor[rgb]{1, 0.995555555555556, 0.5} 77.7 & \cellcolor[rgb]{1, 0.5, 0.5} 108.1 \\                
660 > 0.2 & \cellcolor[rgb]{0.5, 0.973333333333333, 1} 31.3 & \cellcolor[rgb]{0.5, 1, 0.842222222222222} 39.6 & \cellcolor[rgb]{0.54, 1, 0.5} 56.8 & \cellcolor[rgb]{0.735555555555555, 1, 0.5} 65.6                    & & \cellcolor[rgb]{0.5, 0.698666666666667, 1} 18.94 & \cellcolor[rgb]{0.5, 0.946666666666667, 1} 30.1 & \cellcolor[rgb]{0.5, 1, 0.704444444444445} 45.8 & \cellcolor[rgb]{0.893333333333333, 1, 0.5} 72.7 \\  
661 > & \cellcolor[rgb]{0.5, 0.848888888888889, 1} & \cellcolor[rgb]{0.5, 0.973333333333333, 1} & \cellcolor[rgb]{0.5, 1, 0.793333333333333} & \cellcolor[rgb]{0.5, 1, 0.624444444444445}                               & & \cellcolor[rgb]{0.5, 0.599333333333333, 1} & \cellcolor[rgb]{0.5, 0.732666666666667, 1} & \cellcolor[rgb]{0.5, 0.942111111111111, 1} & \cellcolor[rgb]{0.5, 1, 0.695555555555556} \\                        
662 > 0.15 & \cellcolor[rgb]{0.5, 0.724444444444444, 1} 20.1 & \cellcolor[rgb]{0.5, 0.788888888888889, 1} 23.0 & \cellcolor[rgb]{0.5, 0.873333333333333, 1} 26.8 & \cellcolor[rgb]{0.5, 1, 0.984444444444444} 33.2      & & \cellcolor[rgb]{0.5, 0.5, 1} 8.29 & \cellcolor[rgb]{0.5, 0.518666666666667, 1} 10.84 & \cellcolor[rgb]{0.5, 0.588666666666667, 1} 13.99 & \cellcolor[rgb]{0.5, 0.715555555555556, 1} 19.7 \\                
663 > & \cellcolor[rgb]{0.5, 0.696111111111111, 1} & \cellcolor[rgb]{0.5, 0.736333333333333, 1} & \cellcolor[rgb]{0.5, 0.775222222222222, 1} & \cellcolor[rgb]{0.5, 0.860666666666667, 1}                               & & \cellcolor[rgb]{0.5, 0.5, 1} & \cellcolor[rgb]{0.5, 0.509333333333333, 1} & \cellcolor[rgb]{0.5, 0.544333333333333, 1} & \cellcolor[rgb]{0.5, 0.607777777777778, 1} \\                                      
664 > 0.1 & \cellcolor[rgb]{0.5, 0.667777777777778, 1} 17.55 & \cellcolor[rgb]{0.5, 0.683777777777778, 1} 18.27 & \cellcolor[rgb]{0.5, 0.677111111111111, 1} 17.97 & \cellcolor[rgb]{0.5, 0.705777777777778, 1} 19.26   & & \cellcolor[rgb]{0.5, 0.5, 1} 4.96 & \cellcolor[rgb]{0.5, 0.5, 1} 5.46 & \cellcolor[rgb]{0.5, 0.5, 1} 6.04 & \cellcolor[rgb]{0.5,0.5, 1} 6.82 \\                                                            
665   \bottomrule
666   \end{tabular}
667 < \label{tab:tip5peDielectricMap}
667 > \label{tab:DielectricMap}
668   \end{table}
669 +
670   The results of these calculations are displayed in table
671 < \ref{tab:tip5peDielectricMap}. Coloring of the table is based on the
672 < numerical values within the cells and was added to better facilitate
673 < interpretation of the displayed data and highlight trends in the
674 < dielectric constant behavior. This makes it easy to see that the
675 < dielectric constant is linear with respect to $\alpha$ and
676 < $R_\textrm{c}$ in the low to moderate damping regions, and the slope
677 < is 0.025~\AA$^{-1}\cdot R_\textrm{c}^{-1}$. Another point to note is
678 < that choosing $\alpha$ and $R_\textrm{c}$ identical to those used with
679 < the Ewald summation results in the same calculated dielectric
680 < constant. As an example, in the paper outlining the development of
681 < TIP5P-E, the real-space cutoff and Ewald coefficient were tethered to
726 < the system size, and for a 512 molecule system are approximately
727 < 12~\AA\ and 0.25~\AA$^{-1}$ respectively.\cite{Rick04} These
728 < parameters resulted in a dielectric constant of 92$\pm$14, while with
729 < {\sc sf} these parameters give a dielectric constant of
671 > \ref{tab:DielectricMap}.  The dielectric constants for both models
672 > decrease linearly with increasing cutoff radii ($R_\textrm{c}$) and
673 > increase linearly with increasing damping ($\alpha$).  Another point
674 > to note is that choosing $\alpha$ and $R_\textrm{c}$ identical to
675 > those used with the Ewald summation results in the same calculated
676 > dielectric constant. As an example, in the paper outlining the
677 > development of TIP5P-E, the real-space cutoff and Ewald coefficient
678 > were tethered to the system size, and for a 512 molecule system are
679 > approximately 12~\AA\ and 0.25~\AA$^{-1}$ respectively.\cite{Rick04}
680 > These parameters resulted in a dielectric constant of 92$\pm$14, while
681 > with SF these parameters give a dielectric constant of
682   90.8$\pm$0.9. Another example comes from the TIP4P-Ew paper where
683   $\alpha$ and $R_\textrm{c}$ were chosen to be 9.5~\AA\ and
684   0.35~\AA$^{-1}$, and these parameters resulted in a dielectric
685 < constant equal to 63$\pm$1.\cite{Horn04} Calculations using {\sc sf}
686 < with these parameters and this water model give a dielectric constant
687 < of 61$\pm$1. Since the dielectric constant is dependent on $\alpha$
688 < and $R_\textrm{c}$ with the {\sc sf} technique, it might be
689 < interesting to investigate the dielectric dependence of the real-space
690 < Ewald parameters.
685 > constant equal to 63$\pm$1.\cite{Horn04} Calculations using SF with
686 > these parameters and this water model give a dielectric constant of
687 > 61$\pm$1. Since the dielectric constant is dependent on $\alpha$ and
688 > $R_\textrm{c}$ with the SF technique, it might be interesting to
689 > investigate the dielectric dependence of the real-space Ewald
690 > parameters.
691  
692 < Although it is tempting to choose damping parameters equivalent to
693 < these Ewald examples, the results of our previous work indicate that
694 < values this high are destructive to both the energetics and
692 > We have also mapped out the static dielectric constant of SSD/RF as a
693 > function of $R_\textrm{c}$ and $\alpha$. It is apparent from this
694 > table that electrostatic damping has a more pronounced effect on the
695 > dipolar interactions of SSD/RF than the monopolar interactions of
696 > TIP5P-E. The dielectric constant covers a much wider range and has a
697 > steeper slope with increasing damping parameter.
698 >
699 > Although it is tempting to choose damping parameters equivalent to the
700 > Ewald examples, the results of our previous work indicate that values
701 > this high are destructive to both the energetics and
702   dynamics.\cite{Fennell06} Ideally, $\alpha$ should not exceed
703   0.3~\AA$^{-1}$ for any of the cutoff values in this range. If the
704   optimal damping parameter is chosen to be midway between 0.275 and
# Line 748 | Line 707 | behavior for the damped {\sc sf} technique will result
707   0.3~\AA$^{-1}$ for the studied cutoff radii. This linear progression
708   would give values of 0.2875, 0.2625, 0.2375, and 0.2125~\AA$^{-1}$ for
709   cutoff radii of 9, 10, 11, and 12~\AA. Setting this to be the default
710 < behavior for the damped {\sc sf} technique will result in consistent
710 > behavior for the damped SF technique will result in consistent
711   dielectric behavior for these and other condensed molecular systems,
712   regardless of the chosen cutoff radius. The static dielectric
713   constants for TIP5P-E simulations with 9 and 12\AA\ $R_\textrm{c}$
# Line 757 | Line 716 | field (RF) with an infinite RF dielectric constant
716   with the Ewald sum; however, they are more in line with the values
717   reported by Mahoney {\it et al.} for TIP5P while using a reaction
718   field (RF) with an infinite RF dielectric constant
719 < (81.5$\pm$1.6).\cite{Mahoney00} As a final note, aside from a slight
719 > (81.5$\pm$1.6).\cite{Mahoney00}
720 >
721 > We can use the same trend of $\alpha$ with $R_\textrm{c}$ for SSD/RF
722 > and for a 12~\AA\ $R_\textrm{c}$, the resulting dielectric constant is
723 > 82.6$\pm$0.6. This value compares very favorably with the experimental
724 > value of 78.3.\cite{Malmberg56} This is not surprising given that
725 > early studies of the SSD model indicate a static dielectric constant
726 > around 81.\cite{Liu96}  The static dielectric constants for SSD/RF
727 > simulations with 9 and 12\AA\ $R_\textrm{c}$ values using their
728 > respective damping parameters are 82.6$\pm$0.6 and 75$\pm$2.
729 >
730 > As a final note, aside from a slight
731   lowering of $\Delta H_\textrm{vap}$, using these $\alpha$ values does
732   not alter the other other thermodynamic properties.
733  
734   \subsubsection{Dynamic Properties}\label{sec:t5peDynamics}
735  
736 < To look at the dynamic properties of TIP5P-E when using the {\sc sf}
736 > To look at the dynamic properties of TIP5P-E when using the SF
737   method, 200~ps $NVE$ simulations were performed for each temperature
738   at the average density reported by the $NPT$ simulations using 9 and
739   12~\AA\ $R_\textrm{c}$ values using the ideal $\alpha$ values
# Line 829 | Line 799 | constants ({\it lower}) for TIP5P-E using the Ewald su
799   \centering
800   \includegraphics[width=3.5in]{./figures/t5peDynamics.pdf}
801   \caption{Diffusion constants ({\it upper}) and reorientational time
802 < constants ({\it lower}) for TIP5P-E using the Ewald sum and {\sc sf}
802 > constants ({\it lower}) for TIP5P-E using the Ewald sum and SF
803   technique compared with experiment. Data at temperatures less than
804   0$^\circ$C were not included in the $\tau_2^y$ plot to allow for
805   easier comparisons in the more relevant temperature regime.}
# Line 838 | Line 808 | the Ewald sum are reproduced with the {\sc sf} techniq
808   Results for the diffusion constants and orientational relaxation times
809   are shown in figure \ref{fig:t5peDynamics}. From this figure, it is
810   apparent that the trends for both $D$ and $\tau_2^y$ of TIP5P-E using
811 < the Ewald sum are reproduced with the {\sc sf} technique. The enhanced
811 > the Ewald sum are reproduced with the SF technique. The enhanced
812   diffusion at high temperatures are again the product of the lower
813   densities in comparison with experiment and do not provide any special
814   insight into differences between the electrostatic summation
815 < techniques. With the undamped {\sc sf} technique, TIP5P-E tends to
815 > techniques. With the undamped SF technique, TIP5P-E tends to
816   diffuse a little faster than with the Ewald sum; however, use of light
817   to moderate damping results in indistinguishable $D$ values. Though
818 < not apparent in this figure, {\sc sf} values at the lowest temperature
818 > not apparent in this figure, SF values at the lowest temperature
819   are approximately twice as slow as $D$ values obtained using the Ewald
820   sum. These values support the observation from section
821   \ref{sec:t5peThermo} that there appeared to be a change to a more
822 < glassy-like phase with the {\sc sf} technique at these lower
822 > glassy-like phase with the SF technique at these lower
823   temperatures, though this change seems to be more prominent with the
824 < {\it undamped} {\sc sf} method, which has stronger local pairwise
824 > {\it undamped} SF method, which has stronger local pairwise
825   electrostatic interactions.
826  
827   The $\tau_2^y$ results in the lower frame of figure
828   \ref{fig:t5peDynamics} show a much greater difference between the {\sc
829   sf} results and the Ewald results. At all temperatures shown, TIP5P-E
830   relaxes faster than experiment with the Ewald sum while tracking
831 < experiment fairly well when using the {\sc sf} technique, independent
831 > experiment fairly well when using the SF technique, independent
832   of the choice of damping constant. Their are several possible reasons
833   for this deviation between techniques. The Ewald results were
834 < calculated using shorter (10ps) trajectories than the {\sc sf} results
835 < (25ps). A quick calculation from a 10~ps trajectory with {\sc sf} with
834 > calculated using shorter (10ps) trajectories than the SF results
835 > (25ps). A quick calculation from a 10~ps trajectory with SF with
836   an $\alpha$ of 0.2~\AA$^{-1}$ at 25$^\circ$C showed a 0.4~ps drop in
837   $\tau_2^y$, placing the result more in line with that obtained using
838   the Ewald sum. This example supports this explanation; however,
# Line 877 | Line 847 | ensemble, while the $NVE$ ensemble was used for {\sc s
847   technique, if the convergence tolerances are raised for increased
848   performance, error will accumulate in the orientational
849   motion. Finally, the Ewald results were calculated using the $NVT$
850 < ensemble, while the $NVE$ ensemble was used for {\sc sf}
850 > ensemble, while the $NVE$ ensemble was used for SF
851   calculations. The additional mode of motion due to the thermostat will
852   alter the dynamics, resulting in differences between $NVT$ and $NVE$
853   results. These differences are increasingly noticeable as the
# Line 887 | Line 857 | applying the damped {\sc sf} technique to systems cont
857   \subsection{SSD/RF}
858  
859   In section \ref{sec:dampingMultipoles}, we described a method for
860 < applying the damped {\sc sf} technique to systems containing point
860 > applying the damped SF technique to systems containing point
861   multipoles. The soft sticky dipole (SSD) family of water models is the
862   perfect test case because of the dipole-dipole (and
863   charge-dipole/quadrupole) interactions that are present. As alluded to
# Line 900 | Line 870 | consider that the {\sc sf} technique was presented as
870   found in other studies.\cite{Liu96b,Chandra99,Tan03,Fennell04}
871  
872   In deciding which version of the SSD model to use, we need only
873 < consider that the {\sc sf} technique was presented as a pairwise
873 > consider that the SF technique was presented as a pairwise
874   replacement for the Ewald summation. It has been suggested that models
875   parametrized for the Ewald summation (like TIP5P-E) would be
876   appropriate for use with a reaction field and vice versa.\cite{Rick04}
# Line 951 | Line 921 | ions, dissolved organic molecules, or even proteins.
921   SSD/RF can be used effectively with mixed systems, such as dissolved
922   ions, dissolved organic molecules, or even proteins.
923  
954 \subsubsection{Dielectric Constant}
955
956 \begin{table}
957 \centering
958 \caption{SSD/RF dielectric constant as a function of $\alpha$ and $R_\textrm{c}$. The color scale ranges from blue (10) to red (100).}
959 \vspace{6pt}
960 \begin{tabular}{ lcccc }
961 \toprule
962 \toprule
963 & \multicolumn{4}{c}{$R_\textrm{c}$ (\AA )} \\
964 \cmidrule(lr){2-5}
965 $\alpha$ (\AA$^{-1}$) & 9 & 10 & 11 & 12 \\
966 \midrule
967 0.35 & \cellcolor[rgb]{1, 0.5, 0.5} 116 & \cellcolor[rgb]{1, 0.5, 0.5} 119.2 & \cellcolor[rgb]{1, 0.5, 0.5} 131.4 & \cellcolor[rgb]{1, 0.5, 0.5} 130 \\
968 & \cellcolor[rgb]{1, 0.5, 0.5} & \cellcolor[rgb]{1, 0.5, 0.5} & \cellcolor[rgb]{1, 0.5, 0.5} & \cellcolor[rgb]{1, 0.5, 0.5} \\
969 0.3 & \cellcolor[rgb]{1, 0.5, 0.5} 100 & \cellcolor[rgb]{1, 0.5, 0.5} 118.8 & \cellcolor[rgb]{1, 0.5, 0.5} 116 & \cellcolor[rgb]{1, 0.5, 0.5} 122 \\
970 0.275 & \cellcolor[rgb]{1, 1, 0.5} 77.5 & \cellcolor[rgb]{1, 0.5, 0.5} 105 & \cellcolor[rgb]{1, 0.5, 0.5} 118 & \cellcolor[rgb]{1, 0.5, 0.5} 125.2 \\
971 0.25 & \cellcolor[rgb]{0.5, 1, 0.582222222222222} 51.3 & \cellcolor[rgb]{1, 0.993333333333333, 0.5} 77.8 & \cellcolor[rgb]{1, 0.522222222222222, 0.5} 99 & \cellcolor[rgb]{1, 0.5, 0.5} 113 \\
972 0.225 & \cellcolor[rgb]{0.5, 0.984444444444444, 1} 31.8 & \cellcolor[rgb]{0.5, 1, 0.586666666666667} 51.1 & \cellcolor[rgb]{1, 0.995555555555556, 0.5} 77.7 & \cellcolor[rgb]{1, 0.5, 0.5} 108.1 \\
973 0.2 & \cellcolor[rgb]{0.5, 0.698666666666667, 1} 18.94 & \cellcolor[rgb]{0.5, 0.946666666666667, 1} 30.1 & \cellcolor[rgb]{0.5, 1, 0.704444444444445} 45.8 & \cellcolor[rgb]{0.893333333333333, 1, 0.5} 72.7 \\
974 & \cellcolor[rgb]{0.5, 0.599333333333333, 1} & \cellcolor[rgb]{0.5, 0.732666666666667, 1} & \cellcolor[rgb]{0.5, 0.942111111111111, 1} & \cellcolor[rgb]{0.5, 1, 0.695555555555556} \\
975 0.15 & \cellcolor[rgb]{0.5, 0.5, 1} 8.29 & \cellcolor[rgb]{0.5, 0.518666666666667, 1} 10.84 & \cellcolor[rgb]{0.5, 0.588666666666667, 1} 13.99 & \cellcolor[rgb]{0.5, 0.715555555555556, 1} 19.7 \\
976 & \cellcolor[rgb]{0.5, 0.5, 1} & \cellcolor[rgb]{0.5, 0.509333333333333, 1} & \cellcolor[rgb]{0.5, 0.544333333333333, 1} & \cellcolor[rgb]{0.5, 0.607777777777778, 1} \\
977 0.1 & \cellcolor[rgb]{0.5, 0.5, 1} 4.96 & \cellcolor[rgb]{0.5, 0.5, 1} 5.46 & \cellcolor[rgb]{0.5, 0.5, 1} 6.04 & \cellcolor[rgb]{0.5, 0.5, 1} 6.82 \\
978 \bottomrule
979 \end{tabular}
980 \label{tab:ssdrfDielectricMap}
981 \end{table}
982 In addition to the properties tabulated in table
983 \ref{tab:dampedSSDRF}, we mapped out the static dielectric constant of
984 SSD/RF as a function of $R_\textrm{c}$ and $\alpha$ in the same
985 fashion as in section \ref{sec:t5peDielectric} with TIP5P-E. It is
986 apparent from this table that electrostatic damping has a more
987 pronounced effect on the dipolar interactions of SSD/RF than the
988 monopolar interactions of TIP5P-E. The dielectric constant covers a
989 much wider range and has a steeper slope with increasing damping
990 parameter. We can use the same trend of $\alpha$ with $R_\textrm{c}$
991 used with TIP5P-E, and for a 12~\AA\ $R_\textrm{c}$, the resulting
992 dielectric constant is 82.6$\pm$0.6. This value compares very
993 favorably with the experimental value of 78.3.\cite{Malmberg56} This
994 is not surprising given that early studies of the SSD model indicate a
995 static dielectric constant around 81.\cite{Liu96}
996
924   \section{Application of Pairwise Electrostatic Corrections: Imaginary Ice}
925  
926   In an earlier work, we performed a series of free energy calculations
# Line 1002 | Line 929 | observed. Being that the {\sc sf} technique can be use
929   crystallizations of an SSD type single point water
930   model.\cite{Fennell05} In this study, a distinct dependence of the
931   free energies on the interaction cutoff and correction technique was
932 < observed. Being that the {\sc sf} technique can be used as a simple
932 > observed. Being that the SF technique can be used as a simple
933   and efficient replacement for the Ewald summation, it would be
934   interesting to apply it in addressing the question of the free
935   energies of these ice polymorphs with varying water models. To this
936   end, we set up thermodynamic integrations of all of the previously
937 < discussed ice polymorphs using the {\sc sf} technique with a cutoff
937 > discussed ice polymorphs using the SF technique with a cutoff
938   radius of 12~\AA\ and an $\alpha$ of 0.2125~\AA . These calculations
939   were performed on TIP5P-E and TIP4P-Ew (variants of the root models
940   optimized for the Ewald summation) as well as SPC/E, and SSD/RF.
# Line 1015 | Line 942 | using the damped {\sc sf} electrostatic correction met
942   \begin{table}
943   \centering
944   \caption{Helmholtz free energies of ice polymorphs at 1~atm and 200~K
945 < using the damped {\sc sf} electrostatic correction method with a
945 > using the damped SF electrostatic correction method with a
946   variety of water models.}
947   \begin{tabular}{ lccccc }
948   \toprule

Diff Legend

Removed lines
+ Added lines
< Changed lines
> Changed lines