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

Comparing trunk/electrostaticMethodsPaper/electrostaticMethods.tex (file contents):
Revision 2623 by chrisfen, Wed Mar 15 04:34:51 2006 UTC vs.
Revision 2624 by gezelter, Wed Mar 15 17:09:09 2006 UTC

# Line 98 | Line 98 | for an accurate accumulation of electrostatic interact
98  
99   \subsection{The Wolf and Zahn Methods}
100   In a recent paper by Wolf \textit{et al.}, a procedure was outlined
101 < for an accurate accumulation of electrostatic interactions in an
101 > for the accurate accumulation of electrostatic interactions in an
102   efficient pairwise fashion.\cite{Wolf99} Wolf \textit{et al.} observed
103   that the electrostatic interaction is effectively short-ranged in
104   condensed phase systems and that neutralization of the charge
# Line 111 | Line 111 | V^{\textrm{Wolf}}(r_{ij})= \frac{q_iq_j \textrm{erfc}(
111   function (identical to that seen in the real-space portion of the
112   Ewald sum) to aid convergence
113   \begin{equation}
114 < V^{\textrm{Wolf}}(r_{ij})= \frac{q_iq_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\}.
114 > V_{\textrm{Wolf}}(r_{ij})= \frac{q_iq_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\}.
115   \label{eq:WolfPot}
116   \end{equation}
117   Eq. (\ref{eq:WolfPot}) is essentially the common form of a shifted
118   potential.  However, neutralizing the charge contained within each
119   cutoff sphere requires the placement of a self-image charge on the
120   surface of the cutoff sphere.  This additional self-term in the total
121 < potential enables Wolf {\it et al.}  to obtain excellent estimates of
121 > potential enabled Wolf {\it et al.}  to obtain excellent estimates of
122   Madelung energies for many crystals.
123  
124   In order to use their charge-neutralized potential in molecular
# Line 126 | Line 126 | F^{\textrm{Wolf}}(r_{ij}) = q_i q_j\left\{-\left[\frac
126   derivative of this potential prior to evaluation of the limit.  This
127   procedure gives an expression for the forces,
128   \begin{equation}
129 < F^{\textrm{Wolf}}(r_{ij}) = q_i q_j\left\{-\left[\frac{\textrm{erfc}(\alpha r_{ij})}{r^2_{ij}}+\frac{2\alpha}{\pi^{1/2}}\frac{\exp{(-\alpha^2r_{ij}^2)}}{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\},
129 > F_{\textrm{Wolf}}(r_{ij}) = q_i q_j\left\{-\left[\frac{\textrm{erfc}(\alpha r_{ij})}{r^2_{ij}}+\frac{2\alpha}{\pi^{1/2}}\frac{\exp{(-\alpha^2r_{ij}^2)}}{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\},
130   \label{eq:WolfForces}
131   \end{equation}
132   that incorporates both image charges and damping of the electrostatic
# Line 134 | Line 134 | In their work, they pointed out that the method that t
134  
135   More recently, Zahn \textit{et al.} investigated these potential and
136   force expressions for use in simulations involving water.\cite{Zahn02}
137 < In their work, they pointed out that the method that the forces and
138 < derivative of the potential are not commensurate.  Attempts to use
139 < both Eqs. (\ref{eq:WolfPot}) and (\ref{eq:WolfForces}) together will
140 < lead to poor energy conservation.  They correctly observed that taking
141 < the limit shown in equation (\ref{eq:WolfPot}) {\it after} calculating
142 < the derivatives is mathematically invalid.
137 > In their work, they pointed out that the forces and derivative of
138 > the potential are not commensurate.  Attempts to use both
139 > Eqs. (\ref{eq:WolfPot}) and (\ref{eq:WolfForces}) together will lead
140 > to poor energy conservation.  They correctly observed that taking the
141 > limit shown in equation (\ref{eq:WolfPot}) {\it after} calculating the
142 > derivatives gives forces for a different potential energy function
143 > than the one shown in Eq. (\ref{eq:WolfPot}).
144  
145   Zahn \textit{et al.} proposed a modified form of this ``Wolf summation
146   method'' as a way to use this technique in Molecular Dynamics
# Line 147 | Line 148 | V^{\textrm{Zahn}}(r_{ij}) = q_iq_j\left\{\frac{\mathrm
148   \ref{eq:WolfForces}, they proposed a new damped Coulomb
149   potential,
150   \begin{equation}
151 < 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\}.
151 > 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\}.
152   \label{eq:ZahnPot}
153   \end{equation}
154   They showed that this potential does fairly well at capturing the
# Line 158 | Line 159 | tricks: \begin{itemize}
159  
160   The potentials proposed by Wolf \textit{et al.} and Zahn \textit{et
161   al.} are constructed using two different (and separable) computational
162 < tricks: \begin{itemize}
162 > tricks: \begin{enumerate}
163   \item shifting through the use of image charges, and
164   \item damping the electrostatic interaction.
165 < \end{itemize}  Wolf \textit{et al.} treated the
165 > \end{enumerate}  Wolf \textit{et al.} treated the
166   development of their summation method as a progressive application of
167   these techniques,\cite{Wolf99} while Zahn \textit{et al.} founded
168   their damped Coulomb modification (Eq. (\ref{eq:ZahnPot})) on the
# Line 181 | Line 182 | v^\textrm{SP}(r) =     \begin{cases}
182   \textit{et al.}  and Zahn \textit{et al.} by considering the standard
183   shifted potential,
184   \begin{equation}
185 < v^\textrm{SP}(r) =      \begin{cases}
185 > v_\textrm{SP}(r) =      \begin{cases}
186   v(r)-v_\textrm{c} &\quad r\leqslant R_\textrm{c} \\ 0 &\quad r >
187   R_\textrm{c}  
188   \end{cases},
# Line 189 | Line 190 | v^\textrm{SF}(r) =     \begin{cases}
190   \end{equation}
191   and shifted force,
192   \begin{equation}
193 < v^\textrm{SF}(r) =      \begin{cases}
194 < v(r)-v_\textrm{c}-\left(\frac{\textrm{d}v(r)}{\textrm{d}r}\right)_{r=R_\textrm{c}}(r-R_\textrm{c})
193 > v_\textrm{SF}(r) =      \begin{cases}
194 > v(r)-v_\textrm{c}-\left(\frac{d v(r)}{d r}\right)_{r=R_\textrm{c}}(r-R_\textrm{c})
195   &\quad r\leqslant R_\textrm{c} \\ 0 &\quad r > R_\textrm{c}
196                                                  \end{cases},
197   \label{eq:shiftingForm}
# Line 202 | Line 203 | potential is smooth at the cutoff radius
203   potential is smooth at the cutoff radius
204   ($R_\textrm{c}$).\cite{Allen87}
205  
206 <
207 <
207 <
208 < If the derivative term is taken to be zero, we are left with the shifted Coulomb potential devised by Wolf \textit{et al.},\cite{Wolf99}
206 > The forces associated with the shifted potential are simply the forces
207 > of the unshifted potential itself (when inside the cutoff sphere),
208   \begin{equation}
209 < V^\textrm{SP}(r_{ij}) = q_iq_j\left(\frac{1}{r_{ij}}-\frac{1}{R_\textrm{c}}\right) \quad r_{ij}\leqslant R_\textrm{c}.                          \label{eq:WolfSP}
209 > F_{\textrm{SP}} = \left( \frac{d v(r)}{dr} \right),
210   \end{equation}
211 < The forces associated with this potential are obtained by taking the derivative, resulting in the following,
211 > and are zero outside.  Inside the cutoff sphere, the forces associated
212 > with the shifted force form can be written,
213   \begin{equation}
214 < F^\textrm{SP}(r_{ij}) = q_iq_j\left(-\frac{1}{r_{ij}^2}\right) \quad r_{ij}\leqslant R_\textrm{c}.
214 > F_{\textrm{SF}} = \left( \frac{d v(r)}{dr} \right) - \left(\frac{d
215 > v(r)}{dr} \right)_{r=R_\textrm{c}}.
216 > \end{equation}
217 >
218 > If the potential ($v(r)$) is taken to be the normal Coulomb potential,
219 > \begin{equation}
220 > v(r) = \frac{q_i q_j}{r},
221 > \label{eq:Coulomb}
222 > \end{equation}
223 > then the Shifted Potential ({\sc sp}) forms will give Wolf {\it et
224 > al.}'s undamped prescription:
225 > \begin{equation}
226 > V_\textrm{SP}(r) =
227 > q_iq_j\left(\frac{1}{r}-\frac{1}{R_\textrm{c}}\right) \quad
228 > r\leqslant R_\textrm{c},
229 > \label{eq:WolfSP}
230 > \end{equation}
231 > with associated forces,
232 > \begin{equation}
233 > F_\textrm{SP}(r) = q_iq_j\left(-\frac{1}{r^2}\right) \quad r\leqslant R_\textrm{c}.
234   \label{eq:FWolfSP}
235   \end{equation}
236 < These forces are identical to the forces of the standard electrostatic interaction, and this was addressed by Wolf \textit{et al.} as undesirable.  They pointed out that the effect of the image charges is neglected in the forces when they would expect there to be some pressure exerted due to their presence.\cite{Wolf99}  As a consequence the forces, though mathematically valid, may not be physically correct due to this missing component.  Additionally, there is a discontinuity in the forces.  This can be remedied with the use of a switching function to zero the potential and forces smoothly as particles near $R_\textrm{c}$.  
236 > These forces are identical to the forces of the standard Coulomb
237 > interaction, and cutting these off at $R_c$ was addressed by Wolf
238 > \textit{et al.} as undesirable.  They pointed out that the effect of
239 > the image charges is neglected in the forces when this form is
240 > used,\cite{Wolf99} thereby eliminating any benefit from the method in
241 > molecular dynamics.  Additionally, there is a discontinuity in the
242 > forces at the cutoff radius which results in energy drift during MD
243 > simulations.
244  
245 < If the derivative term in equation \ref{eq:shiftingForm} is evaluated, we obtain an hitherto undiscussed shifted force Coulomb potential,
245 > The shifted force ({\sc sf}) form using the normal Coulomb potential
246 > will give,
247   \begin{equation}
248 < V^\textrm{SF}(r_{ij}) = q_iq_j\left[\frac{1}{r_{ij}}-\frac{1}{R_\textrm{c}}+\left(\frac{1}{R_\textrm{c}^2}\right)(r_{ij}-R_\textrm{c})\right] \quad r_{ij}\leqslant R_\textrm{c}.
248 > 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}.
249   \label{eq:SFPot}
250   \end{equation}
251 < Taking the derivative of this shifted force potential gives the
225 < following forces,
251 > with associated forces,
252   \begin{equation}
253 < F^\textrm{SF}(r_{ij} =  q_iq_j\left(-\frac{1}{r_{ij}^2}+\frac{1}{R_\textrm{c}^2}\right) \quad r_{ij}\leqslant R_\textrm{c}.
253 > 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}.
254   \label{eq:SFForces}
255   \end{equation}
256 < Using this formulation rather than the simple shifted potential
257 < (Eq. \ref{eq:WolfSP}) means that there are no discontinuities in the
258 < forces in addition to the potential.  This form also has the benefit
259 < that the image charges have a force presence, addressing the concerns
260 < about a missing physical component.  One side effect of this treatment
261 < is a slight alteration in the shape of the potential that comes about
262 < from the derivative term.  Thus, a degree of clarity about the
263 < original formulation of the potential is lost in order to gain
264 < functionality in dynamics simulations.
256 > This formulation has the benefits that there are no discontinuities at
257 > the cutoff distance, while the neutralizing image charges are present
258 > in both the energy and force expressions.  It would be simple to add
259 > the self-neutralizing term back when computing the total energy of the
260 > system, thereby maintaining the agreement with the Madelung energies.
261 > A side effect of this treatment is the alteration in the shape of the
262 > potential that comes from the derivative term.  Thus, a degree of
263 > clarity about agreement with the empirical potential is lost in order
264 > to gain functionality in dynamics simulations.
265  
266   Wolf \textit{et al.} originally discussed the energetics of the
267   shifted Coulomb potential (Eq. \ref{eq:WolfSP}), and they found that
268 < it was still insufficient for accurate determination of the energy.
269 < The energy would fluctuate around the expected value with increasing
270 < cutoff radius, but the oscillations appeared to be converging toward
271 < the correct value.\cite{Wolf99} A damping function was incorporated to
272 < accelerate convergence; and though alternative functional forms could
273 < be used,\cite{Jones56,Heyes81} the complimentary error function was
274 < chosen to draw parallels to the Ewald summation.  Incorporating
275 < damping into the simple Coulomb potential,
268 > it was still insufficient for accurate determination of the energy
269 > with reasonable cutoff distances.  The calculated Madelung energies
270 > fluctuate around the expected value with increasing cutoff radius, but
271 > the oscillations converge toward the correct value.\cite{Wolf99} A
272 > damping function was incorporated to accelerate the convergence; and
273 > though alternative functional forms could be
274 > used,\cite{Jones56,Heyes81} the complimentary error function was
275 > chosen to mirror the effective screening used in the Ewald summation.
276 > Incorporating this error function damping into the simple Coulomb
277 > potential,
278   \begin{equation}
279 < v(r_{ij}) = \frac{\mathrm{erfc}\left(\alpha r_{ij}\right)}{r_{ij}},
279 > v(r) = \frac{\mathrm{erfc}\left(\alpha r\right)}{r},
280   \label{eq:dampCoulomb}
281   \end{equation}
282 < the shifted potential (Eq. \ref{eq:WolfSP}) can be rederived
282 > the shifted potential (Eq. \ref{eq:WolfSP}) can be recovered
283   \textit{via} equation \ref{eq:shiftingForm},
284   \begin{equation}
285 < V^{\textrm{DSP}}(r_{ij}) = q_iq_j\left(\frac{\textrm{erfc}(\alpha r_{ij})}{r_{ij}}-\frac{\textrm{erfc}(\alpha R_\textrm{c})}{R_\textrm{c}}\right) \quad r_{ij}\leqslant R_\textrm{c}.
285 > v_{\textrm{DSP}}(r) = q_iq_j\left(\frac{\textrm{erfc}(\alpha r)}{r}-\frac{\textrm{erfc}(\alpha R_\textrm{c})}{R_\textrm{c}}\right) \quad r\leqslant R_\textrm{c}.
286   \label{eq:DSPPot}
287 < \end{equation}
288 < The derivative of this Shifted-Potential can be taken to obtain forces
261 < for use in MD,
287 > \end{equation},
288 > with associated forces,
289   \begin{equation}
290 < F^{\textrm{DSP}}(r_{ij}) = q_iq_j\left(-\frac{\textrm{erfc}(\alpha r_{ij})}{r^2_{ij}}-\frac{2\alpha}{\pi^{1/2}}\frac{\exp{(-\alpha^2r_{ij}^2)}}{r_{ij}}\right) \quad r_{ij}\leqslant R_\textrm{c}.
290 > f_{\textrm{DSP}}(r) = q_iq_j\left(-\frac{\textrm{erfc}(\alpha r)}{r^2}-\frac{2\alpha}{\pi^{1/2}}\frac{\exp{(-\alpha^2r^2)}}{r}\right) \quad r\leqslant R_\textrm{c}.
291   \label{eq:DSPForces}
292   \end{equation}
293 < Again, this Shifted-Potential suffers from a discontinuity in the
294 < forces, and a lack of an image-charge component in the forces.  To
295 < remedy these concerns, a Shifted-Force variant is obtained by
296 < inclusion of the derivative term in equation \ref{eq:shiftingForm} to
270 < give,
293 > Again, this damped shifted potential suffers from a discontinuity and
294 > a lack of the image charges in the forces.  To remedy these concerns,
295 > one may derive a Shifted-Force variant by including  the derivative
296 > term in equation \ref{eq:shiftingForm},
297   \begin{equation}
298   \begin{split}
299 < V^\mathrm{DSF}(r_{ij}) = q_iq_j\Biggr{[}&\frac{\mathrm{erfc}\left(\alpha r_{ij}\right)}{r_{ij}}-\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_{ij}-R_\mathrm{c}\right)\ \right] \quad r_{ij}\leqslant R_\textrm{c}.
299 > 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}.
300   \label{eq:DSFPot}
301   \end{split}
302   \end{equation}
303   The derivative of the above potential gives the following forces,
304   \begin{equation}
305   \begin{split}
306 < F^\mathrm{DSF}(r_{ij}) = q_iq_j\Biggr{[}&-\left(\frac{\textrm{erfc}(\alpha r_{ij})}{r^2_{ij}}+\frac{2\alpha}{\pi^{1/2}}\frac{\exp{(-\alpha^2r_{ij}^2)}}{r_{ij}}\right) \\ &\left.+\left(\frac{\textrm{erfc}(\alpha R_{\textrm{c}})}{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_{ij}\leqslant R_\textrm{c}.
306 > f_\mathrm{DSF}(r) = q_iq_j\Biggr{[}&-\left(\frac{\textrm{erfc}(\alpha r)}{r^2}+\frac{2\alpha}{\pi^{1/2}}\frac{\exp{(-\alpha^2r^2)}}{r}\right) \\ &\left.+\left(\frac{\textrm{erfc}(\alpha R_{\textrm{c}})}{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}.
307   \label{eq:DSFForces}
308   \end{split}
309   \end{equation}
# Line 285 | Line 311 | from equation \ref{eq:shiftingForm} is equal to equati
311   This new Shifted-Force potential is similar to equation
312   \ref{eq:ZahnPot} derived by Zahn \textit{et al.}; however, there are
313   two important differences.\cite{Zahn02} First, the $v_\textrm{c}$ term
314 < from equation \ref{eq:shiftingForm} is equal to equation
315 < \ref{eq:dampCoulomb} with $R_\textrm{c}$ supplied for $r_{ij}$.  This
316 < term is not present in the Zahn potential, resulting in a
317 < discontinuity as particles cross $R_\textrm{c}$.  Second, the sign of
318 < the derivative portion is different.  The constant $v_\textrm{c}$ term
319 < is not going to have a presence in the forces after performing the
320 < derivative, but the negative sign does effect the derivative.  In
321 < fact, it introduces a discontinuity in the forces at the cutoff,
322 < because the force function is shifted in the wrong direction and
323 < doesn't cross zero at $R_\textrm{c}$.  Thus, these alterations make
324 < for an electrostatic summation method that is continuous in both the
325 < potential and forces and incorporates the pairwise sum considerations
300 < stressed by Wolf \textit{et al.}\cite{Wolf99}
314 > from eq. (\ref{eq:shiftingForm}) is equal to
315 > eq. (\ref{eq:dampCoulomb}) with $r$ replaced by $R_\textrm{c}$.  This
316 > term is {\it not} present in the Zahn potential, resulting in a
317 > potential discontinuity as particles cross $R_\textrm{c}$.  Second,
318 > the sign of the derivative portion is different.  The missing
319 > $v_\textrm{c}$ term would not affect molecular dynamics simulations
320 > (although the computed energy would be expected to have sudden jumps
321 > as particle distances crossed $R_c$).  The sign problem would be a
322 > potential source of errors, however.  In fact, it introduces a
323 > discontinuity in the forces at the cutoff, because the force function
324 > is shifted in the wrong direction and doesn't cross zero at
325 > $R_\textrm{c}$.  
326  
327 + Eqs. (\ref{eq:DSFPot}) and (\ref{eq:DSFForces}) result in an
328 + electrostatic summation method that is continuous in both the
329 + potential and forces and which incorporates the damping function
330 + proposed by Wolf \textit{et al.}\cite{Wolf99} In the rest of this
331 + paper, we will evaluate exactly how good these methods ({\sc sp}, {\sc
332 + sf}, damping) are at reproducing the correct electrostatic summation
333 + performed by the Ewald sum.
334 +
335 + \subsection{Other alternatives}
336 +
337 + Reaction Field blah
338 +
339 + Group-based cutoff blah
340 +
341 +
342   \section{Methods}
343  
304 \subsection{What Qualities are Important?}\label{sec:Qualities}
344   In classical molecular mechanics simulations, there are two primary
345   techniques utilized to obtain information about the system of
346   interest: Monte Carlo (MC) and Molecular Dynamics (MD).  Both of these
# Line 310 | Line 349 | the origins of this method, the Canonical ensemble acc
349  
350   In MC, the potential energy difference between two subsequent
351   configurations dictates the progression of MC sampling.  Going back to
352 < the origins of this method, the Canonical ensemble acceptance criteria
353 < laid out by Metropolis \textit{et al.} states that a subsequent
354 < configuration is accepted if $\Delta E < 0$ or if $\xi < \exp(-\Delta
355 < E/kT)$, where $\xi$ is a random number between 0 and
356 < 1.\cite{Metropolis53} Maintaining a consistent $\Delta E$ when using
357 < an alternate method for handling the long-range electrostatics ensures
358 < proper sampling within the ensemble.
352 > the origins of this method, the acceptance criterion for the canonical
353 > ensemble laid out by Metropolis \textit{et al.} states that a
354 > subsequent configuration is accepted if $\Delta E < 0$ or if $\xi <
355 > \exp(-\Delta E/kT)$, where $\xi$ is a random number between 0 and
356 > 1.\cite{Metropolis53} Maintaining the correct $\Delta E$ when using an
357 > alternate method for handling the long-range electrostatics will
358 > ensure proper sampling from the ensemble.
359  
360 < In MD, the derivative of the potential directs how the system will
360 > In MD, the derivative of the potential governs how the system will
361   progress in time.  Consequently, the force and torque vectors on each
362 < body in the system dictate how it develops as a whole.  If the
363 < magnitude and direction of these vectors are similar when using
364 < alternate electrostatic summation techniques, the dynamics in the near
365 < term will be indistinguishable.  Because error in MD calculations is
366 < cumulative, one should expect greater deviation in the long term
367 < trajectories with greater differences in these vectors between
368 < configurations using different long-range electrostatics.
362 > body in the system dictate how the system evolves.  If the magnitude
363 > and direction of these vectors are similar when using alternate
364 > electrostatic summation techniques, the dynamics in the short term
365 > will be indistinguishable.  Because error in MD calculations is
366 > cumulative, one should expect greater deviation at longer times,
367 > although methods which have large differences in the force and torque
368 > vectors will diverge from each other more rapidly.
369  
370   \subsection{Monte Carlo and the Energy Gap}\label{sec:MCMethods}
371 < Evaluation of the pairwise summation techniques (outlined in section
372 < \ref{sec:ESMethods}) for use in MC simulations was performed through
373 < study of the energy differences between conformations.  Considering
374 < the SPME results to be the correct or desired behavior, ideal
375 < performance of a tested method was taken to be agreement between the
376 < energy differences calculated.  Linear least squares regression of the
377 < $\Delta E$ values between configurations using SPME against $\Delta E$
378 < values using tested methods provides a quantitative comparison of this
379 < agreement.  Unitary results for both the correlation and correlation
380 < coefficient for these regressions indicate equivalent energetic
381 < results between the methods.  The correlation is the slope of the
382 < plotted data while the correlation coefficient ($R^2$) is a measure of
383 < the of the data scatter around the fitted line and tells about the
384 < quality of the fit (Fig. \ref{fig:linearFit}).
371 > The pairwise summation techniques (outlined in section
372 > \ref{sec:ESMethods}) were evaluated for use in MC simulations by
373 > studying the energy differences between conformations.  We took the
374 > SPME-computed energy difference between two conformations to be the
375 > correct behavior. An ideal performance by an alternative method would
376 > reproduce these energy differences exactly.  Since none of the methods
377 > provide exact energy differences, we used linear least squares
378 > regressions of the $\Delta E$ values between configurations using SPME
379 > against $\Delta E$ values using tested methods provides a quantitative
380 > comparison of this agreement.  Unitary results for both the
381 > correlation and correlation coefficient for these regressions indicate
382 > equivalent energetic results between the method under consideration
383 > and electrostatics handled using SPME.  Sample correlation plots for
384 > two alternate methods are shown in Fig. \ref{fig:linearFit}.
385  
386   \begin{figure}
387   \centering
# Line 351 | Line 390 | Each system type (detailed in section \ref{sec:RepSims
390   \label{fig:linearFit}
391   \end{figure}
392  
393 < Each system type (detailed in section \ref{sec:RepSims}) studied
394 < consisted of 500 independent configurations, each equilibrated from
395 < higher temperature trajectories. Thus, 124,750 $\Delta E$ data points
396 < are used in a regression of a single system type.  Results and
397 < discussion for the individual analysis of each of the system types
359 < appear in the supporting information, while the cumulative results
360 < over all the investigated systems appears below in section
361 < \ref{sec:EnergyResults}.
393 > Each system type (detailed in section \ref{sec:RepSims}) was
394 > represented using 500 independent configurations.  Additionally, we
395 > used seven different system types, so each of the alternate
396 > (non-Ewald) electrostatic summation methods was evaluated using
397 > 873,250 configurational energy differences.
398  
399 < \subsection{Molecular Dynamics and the Force and Torque Vectors}\label{sec:MDMethods}
400 < Evaluation of the pairwise methods (outlined in section
401 < \ref{sec:ESMethods}) for use in MD simulations was performed through
402 < comparison of the force and torque vectors obtained with those from
367 < SPME.  Both the magnitude and the direction of these vectors on each
368 < of the bodies in the system were analyzed.  For the magnitude of these
369 < vectors, linear least squares regression analysis can be performed as
370 < described previously for comparing $\Delta E$ values. Instead of a
371 < single value between two system configurations, there is a value for
372 < each particle in each configuration.  For a system of 1000 water
373 < molecules and 40 ions, there are 1040 force vectors and 1000 torque
374 < vectors.  With 500 configurations, this results in 520,000 force and
375 < 500,000 torque vector comparisons samples for each system type.
399 > Results and discussion for the individual analysis of each of the
400 > system types appear in the supporting information, while the
401 > cumulative results over all the investigated systems appears below in
402 > section \ref{sec:EnergyResults}.
403  
404 < The force and torque vector directions were investigated through
405 < measurement of the angle ($\theta$) formed between those from the
406 < particular method and those from SPME
404 > \subsection{Molecular Dynamics and the Force and Torque Vectors}\label{sec:MDMethods}
405 > We evaluated the pairwise methods (outlined in section
406 > \ref{sec:ESMethods}) for use in MD simulations by
407 > comparing the force and torque vectors with those obtained using the
408 > reference Ewald summation (SPME).  Both the magnitude and the
409 > direction of these vectors on each of the bodies in the system were
410 > analyzed.  For the magnitude of these vectors, linear least squares
411 > regression analyses were performed as described previously for
412 > comparing $\Delta E$ values.  Instead of a single energy difference
413 > between two system configurations, we compared the magnitudes of the
414 > forces (and torques) on each molecule in each configuration.  For a
415 > system of 1000 water molecules and 40 ions, there are 1040 force
416 > vectors and 1000 torque vectors.  With 500 configurations, this
417 > results in 520,000 force and 500,000 torque vector comparisons.
418 > Additionally, data from seven different system types was aggregated
419 > before the comparison was made.
420 >
421 > The {\it directionality} of the force and torque vectors was
422 > investigated through measurement of the angle ($\theta$) formed
423 > between those computed from the particular method and those from SPME,
424   \begin{equation}
425 < \theta_F = \frac{\vec{F}_\textrm{SPME}}{|\vec{F}_\textrm{SPME}|}\cdot\frac{\vec{F}_\textrm{Method}}{|\vec{F}_\textrm{Method}|}.
425 > \theta_f = \cos^{-1} \hat{f}_\textrm{SPME} \cdot \hat{f}_\textrm{Method},
426   \end{equation}
427 + where $\hat{f}_\textrm{M}$ is the unit vector pointing along the
428 + force vector computed using method $M$.  
429 +
430   Each of these $\theta$ values was accumulated in a distribution
431 < function, weighted by the area on the unit sphere.  Non-linear fits
432 < were used to measure the shape of the resulting distributions.
431 > function, weighted by the area on the unit sphere.  Non-linear
432 > Gaussian fits were used to measure the width of the resulting
433 > distributions.
434  
435   \begin{figure}
436   \centering
# Line 395 | Line 443 | between two different electrostatic summation methods,
443   non-linear fits.  The solid line is a Gaussian profile, while the
444   dotted line is a Voigt profile, a convolution of a Gaussian and a
445   Lorentzian.  Since this distribution is a measure of angular error
446 < between two different electrostatic summation methods, there is
447 < particular reason for the profile to adhere to a specific shape.
448 < Because of this and the Gaussian profile's more statistically
449 < meaningful properties, Gaussian fits was used to compare all the
450 < tested methods.  The variance ($\sigma^2$) was extracted from each of
451 < these fits and was used to compare distribution widths.  Values of
452 < $\sigma^2$ near zero indicate vector directions indistinguishable from
453 < those calculated when using SPME.
446 > between two different electrostatic summation methods, there is no
447 > {\it a priori} reason for the profile to adhere to any specific shape.
448 > Gaussian fits was used to compare all the tested methods.  The
449 > variance ($\sigma^2$) was extracted from each of these fits and was
450 > used to compare distribution widths.  Values of $\sigma^2$ near zero
451 > indicate vector directions indistinguishable from those calculated
452 > when using the reference method (SPME).
453 >
454 > \subsection{Short-time Dynamics}
455  
456   \subsection{Long-Time and Collective Motion}\label{sec:LongTimeMethods}
457   Evaluation of the long-time dynamics of charged systems was performed
# Line 752 | Line 801 | V_\textrm{damped}=\sum^N_i\sum^N_{j\ne i}q_iq_j\left[\
801   clearly, we show how damping strength affects a simple real-space
802   electrostatic potential,
803   \begin{equation}
804 < V_\textrm{damped}=\sum^N_i\sum^N_{j\ne i}q_iq_j\left[\frac{\textrm{erfc}({\alpha r_{ij}})}{r_{ij}}\right]S(r),
804 > V_\textrm{damped}=\sum^N_i\sum^N_{j\ne i}q_iq_j\left[\frac{\textrm{erfc}({\alpha r})}{r}\right]S(r),
805   \end{equation}
806   where $S(r)$ is a switching function that smoothly zeroes the
807   potential at the cutoff radius.  Figure \ref{fig:dampInc} shows how

Diff Legend

Removed lines
+ Added lines
< Changed lines
> Changed lines