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 2635 by chrisfen, Thu Mar 16 15:03:20 2006 UTC vs.
Revision 2636 by chrisfen, Sat Mar 18 05:31:17 2006 UTC

# Line 2 | Line 2
2   %\documentclass[aps,prb,preprint]{revtex4}
3   \documentclass[11pt]{article}
4   \usepackage{endfloat}
5 < \usepackage{amsmath}
5 > \usepackage{amsmath,bm}
6   \usepackage{amssymb}
7   \usepackage{epsf}
8   \usepackage{times}
# Line 81 | Line 81 | impractical task to perform these calculations.
81   impractical task to perform these calculations.
82  
83   \subsection{The Ewald Sum}
84 < blah blah blah Ewald Sum Important blah blah blah
84 > The complete accumulation electrostatic interactions in a system with periodic boundary conditions (PBC) requires the consideration of the effect of all charges within a simulation box, as well as those in the periodic replicas,
85 > \begin{equation}
86 > V_\textrm{elec} = \frac{1}{2} {\sum_{\mathbf{n}}}^\prime \left[ \sum_{i=1}^N\sum_{j=1}^N \phi\left( \mathbf{r}_{ij} + L\mathbf{n},\bm{\Omega}_i,\bm{\Omega}_j\right) \right],
87 > \label{eq:PBCSum}
88 > \end{equation}
89 > where the sum over $\mathbf{n}$ is a sum over all periodic box replicas
90 > with integer coordinates $\mathbf{n} = (l,m,n)$, and the prime indicates
91 > $i = j$ are neglected for $\mathbf{n} = 0$.\cite{deLeeuw80} Within the
92 > sum, $N$ is the number of electrostatic particles, $\mathbf{r}_{ij}$ is
93 > $\mathbf{r}_j - \mathbf{r}_i$, $L$ is the cell length, $\bm{\Omega}_{i,j}$ are
94 > the Euler angles for $i$ and $j$, and $\phi$ is Poisson's equation
95 > ($\phi(\mathbf{r}_{ij}) = q_i q_j |\mathbf{r}_{ij}|^{-1}$ for charge-charge
96 > interactions). In the case of monopole electrostatics,
97 > eq. (\ref{eq:PBCSum}) is conditionally convergent and is discontiuous
98 > for non-neutral systems.
99  
100 + This electrostatic summation problem was originally studied by Ewald
101 + for the case of an infinite crystal.\cite{Ewald21}. The approach he
102 + took was to convert this conditionally convergent sum into two
103 + absolutely convergent summations: a short-ranged real-space summation
104 + and a long-ranged reciprocal-space summation,
105 + \begin{equation}
106 + \begin{split}
107 + V_\textrm{elec} = \frac{1}{2}& \sum_{i=1}^N\sum_{j=1}^N \Biggr[ q_i q_j\Biggr( {\sum_{\mathbf{n}}}^\prime \frac{\textrm{erfc}\left( \alpha|\mathbf{r}_{ij}+\mathbf{n}|\right)}{|\mathbf{r}_{ij}+\mathbf{n}|} \\ &+ \frac{1}{\pi L^3}\sum_{\mathbf{k}\ne 0}\frac{4\pi^2}{|\mathbf{k}|^2} \exp{\left(-\frac{\pi^2|\mathbf{k}|^2}{\alpha^2}\right) \cos\left(\mathbf{k}\cdot\mathbf{r}_{ij}\right)}\Biggr)\Biggr] \\ &- \frac{\alpha}{\pi^{1/2}}\sum_{i=1}^N q_i^2 + \frac{2\pi}{3L^3}\left|\sum_{i=1}^N q_i\mathbf{r}_i\right|^2,
108 + \end{split}
109 + \label{eq:EwaldSum}
110 + \end{equation}
111 + where $\alpha$ is a damping parameter, or separation constant, with
112 + units of \AA$^{-1}$, and $\mathbf{k}$ are the reciprocal vectors and
113 + equal $2\pi\mathbf{n}/L^2$. The final two terms of
114 + eq. (\ref{eq:EwaldSum}) are a particle-self term and a dipolar term
115 + for interacting with a surrounding dielectric.\cite{Allen87} This
116 + dipolar term was neglected in early applications in molecular
117 + simulations,\cite{Brush66,Woodcock71} until it was introduced by de
118 + Leeuw {\it et al.} to address situations where the unit cell has a
119 + dipole moment and this dipole moment gets magnified through
120 + replication of the periodic images.\cite{deLeeuw80} This term is zero
121 + for systems where $\epsilon_{\rm S} = \infty$. Figure
122 + \ref{fig:ewaldTime} shows how the Ewald sum has been applied over
123 + time.  Initially, due to the small size of systems, the entire
124 + simulation box was replicated to convergence.  Currently, we balance a
125 + spherical real-space cutoff with the reciprocal sum and consider the
126 + surrounding dielectric.
127   \begin{figure}
128   \centering
129   \includegraphics[width = \linewidth]{./ewaldProgression.pdf}
# Line 96 | Line 137 | a surrounding dielectric term is included.}
137   \label{fig:ewaldTime}
138   \end{figure}
139  
140 + The Ewald summation in the straight-forward form is an
141 + $\mathscr{O}(N^2)$ algorithm.  The separation constant $(\alpha)$
142 + plays an important role in the computational cost balance between the
143 + direct and reciprocal-space portions of the summation.  The choice of
144 + the magnitude of this value allows one to whether the real-space or
145 + reciprocal space portion of the summation is an $\mathscr{O}(N^2)$
146 + calcualtion, with the other being $\mathscr{O}(N)$.\cite{Sagui99} With
147 + appropriate choice of $\alpha$ and thoughtful algorithm development,
148 + this cost can be brought down to
149 + $\mathscr{O}(N^{3/2})$.\cite{Perram88} The typical route taken to
150 + accelerate the Ewald summation is to se
151 +
152   \subsection{The Wolf and Zahn Methods}
153   In a recent paper by Wolf \textit{et al.}, a procedure was outlined
154   for the accurate accumulation of electrostatic interactions in an
# Line 126 | Line 179 | procedure gives an expression for the forces,
179   derivative of this potential prior to evaluation of the limit.  This
180   procedure gives an expression for the forces,
181   \begin{equation}
182 < 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\},
182 > F_{\textrm{Wolf}}(r_{ij}) = q_i q_j\left\{\left[\frac{\textrm{erfc}\left(\alpha r_{ij}\right)}{r^2_{ij}}+\frac{2\alpha}{\pi^{1/2}}\frac{\exp{\left(-\alpha^2r_{ij}^2\right)}}{r_{ij}}\right]-\left[\frac{\textrm{erfc}\left(\alpha R_{\textrm{c}}\right)}{R_{\textrm{c}}^2}+\frac{2\alpha}{\pi^{1/2}}\frac{\exp{\left(-\alpha^2R_{\textrm{c}}^2\right)}}{R_{\textrm{c}}}\right]\right\},
183   \label{eq:WolfForces}
184   \end{equation}
185   that incorporates both image charges and damping of the electrostatic
# Line 206 | Line 259 | of the unshifted potential itself (when inside the cut
259   The forces associated with the shifted potential are simply the forces
260   of the unshifted potential itself (when inside the cutoff sphere),
261   \begin{equation}
262 < F_{\textrm{SP}} = \left( \frac{d v(r)}{dr} \right),
262 > f_{\textrm{SP}} = -\left( \frac{d v(r)}{dr} \right),
263   \end{equation}
264   and are zero outside.  Inside the cutoff sphere, the forces associated
265   with the shifted force form can be written,
266   \begin{equation}
267 < F_{\textrm{SF}} = \left( \frac{d v(r)}{dr} \right) - \left(\frac{d
267 > f_{\textrm{SF}} = -\left( \frac{d v(r)}{dr} \right) + \left(\frac{d
268   v(r)}{dr} \right)_{r=R_\textrm{c}}.
269   \end{equation}
270  
# Line 223 | Line 276 | al.}'s undamped prescription:
276   then the Shifted Potential ({\sc sp}) forms will give Wolf {\it et
277   al.}'s undamped prescription:
278   \begin{equation}
279 < V_\textrm{SP}(r) =
279 > v_\textrm{SP}(r) =
280   q_iq_j\left(\frac{1}{r}-\frac{1}{R_\textrm{c}}\right) \quad
281   r\leqslant R_\textrm{c},
282 < \label{eq:WolfSP}
282 > \label{eq:SPPot}
283   \end{equation}
284   with associated forces,
285   \begin{equation}
286 < F_\textrm{SP}(r) = q_iq_j\left(-\frac{1}{r^2}\right) \quad r\leqslant R_\textrm{c}.
287 < \label{eq:FWolfSP}
286 > f_\textrm{SP}(r) = q_iq_j\left(\frac{1}{r^2}\right) \quad r\leqslant R_\textrm{c}.
287 > \label{eq:SPForces}
288   \end{equation}
289   These forces are identical to the forces of the standard Coulomb
290   interaction, and cutting these off at $R_c$ was addressed by Wolf
# Line 245 | Line 298 | will give,
298   The shifted force ({\sc sf}) form using the normal Coulomb potential
299   will give,
300   \begin{equation}
301 < 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}.
301 > 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}.
302   \label{eq:SFPot}
303   \end{equation}
304   with associated forces,
305   \begin{equation}
306 < 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}.
306 > 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}.
307   \label{eq:SFForces}
308   \end{equation}
309   This formulation has the benefits that there are no discontinuities at
# Line 264 | Line 317 | Wolf \textit{et al.} originally discussed the energeti
317   to gain functionality in dynamics simulations.
318  
319   Wolf \textit{et al.} originally discussed the energetics of the
320 < shifted Coulomb potential (Eq. \ref{eq:WolfSP}), and they found that
320 > shifted Coulomb potential (Eq. \ref{eq:SPPot}), and they found that
321   it was still insufficient for accurate determination of the energy
322   with reasonable cutoff distances.  The calculated Madelung energies
323   fluctuate around the expected value with increasing cutoff radius, but
# Line 279 | Line 332 | v(r) = \frac{\mathrm{erfc}\left(\alpha r\right)}{r},
332   v(r) = \frac{\mathrm{erfc}\left(\alpha r\right)}{r},
333   \label{eq:dampCoulomb}
334   \end{equation}
335 < the shifted potential (Eq. (\ref{eq:WolfSP})) can be recovered
336 < using eq. (\ref{eq:shiftingForm}),
335 > the shifted potential (Eq. (\ref{eq:SPPot})) can be reacquired using
336 > eq. (\ref{eq:shiftingForm}),
337   \begin{equation}
338 < 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},
338 > v_{\textrm{DSP}}(r) = q_iq_j\left(\frac{\textrm{erfc}\left(\alpha r\right)}{r}-\frac{\textrm{erfc}\left(\alpha R_\textrm{c}\right)}{R_\textrm{c}}\right) \quad r\leqslant R_\textrm{c},
339   \label{eq:DSPPot}
340   \end{equation}
341   with associated forces,
342   \begin{equation}
343 < 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}.
343 > f_{\textrm{DSP}}(r) = q_iq_j\left(\frac{\textrm{erfc}\left(\alpha r\right)}{r^2}+\frac{2\alpha}{\pi^{1/2}}\frac{\exp{\left(-\alpha^2r^2\right)}}{r}\right) \quad r\leqslant R_\textrm{c}.
344   \label{eq:DSPForces}
345   \end{equation}
346   Again, this damped shifted potential suffers from a discontinuity and
# Line 300 | Line 353 | v_\mathrm{DSF}(r) = q_iq_j\Biggr{[}&\frac{\mathrm{erfc
353   \label{eq:DSFPot}
354   \end{split}
355   \end{equation}
356 < The derivative of the above potential gives the following forces,
356 > The derivative of the above potential will lead to the following forces,
357   \begin{equation}
358   \begin{split}
359 < 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}.
359 > f_\mathrm{DSF}(r) = q_iq_j\Biggr{[}&\left(\frac{\textrm{erfc}\left(\alpha r\right)}{r^2}+\frac{2\alpha}{\pi^{1/2}}\frac{\exp{\left(-\alpha^2r^2\right)}}{r}\right) \\ &\left.-\left(\frac{\textrm{erfc}\left(\alpha R_{\textrm{c}}\right)}{R_{\textrm{c}}^2}+\frac{2\alpha}{\pi^{1/2}}\frac{\exp{\left(-\alpha^2R_{\textrm{c}}^2\right)}}{R_{\textrm{c}}}\right)\right] \quad r\leqslant R_\textrm{c}.
360   \label{eq:DSFForces}
361   \end{split}
362   \end{equation}
363 + If the damping parameter $(\alpha)$ is chosen to be zero, the undamped
364 + case, eqs. (\ref{eq:SPPot}-\ref{eq:SFForces}) are correctly recovered
365 + from eqs. (\ref{eq:DSPPot}-\ref{eq:DSFForces}).
366  
367 < This new {\sc sf} potential is similar to equation
368 < \ref{eq:ZahnPot} derived by Zahn \textit{et al.}; however, there are
369 < two important differences.\cite{Zahn02} First, the $v_\textrm{c}$ term
370 < from eq. (\ref{eq:shiftingForm}) is equal to
371 < eq. (\ref{eq:dampCoulomb}) with $r$ replaced by $R_\textrm{c}$.  This
372 < term is {\it not} present in the Zahn potential, resulting in a
373 < potential discontinuity as particles cross $R_\textrm{c}$.  Second,
374 < the sign of the derivative portion is different.  The missing
375 < $v_\textrm{c}$ term would not affect molecular dynamics simulations
376 < (although the computed energy would be expected to have sudden jumps
377 < as particle distances crossed $R_c$).  The sign problem would be a
378 < potential source of errors, however.  In fact, it introduces a
379 < discontinuity in the forces at the cutoff, because the force function
380 < is shifted in the wrong direction and doesn't cross zero at
325 < $R_\textrm{c}$.  
367 > This new {\sc sf} potential is similar to equation \ref{eq:ZahnPot}
368 > derived by Zahn \textit{et al.}; however, there are two important
369 > differences.\cite{Zahn02} First, the $v_\textrm{c}$ term from
370 > eq. (\ref{eq:shiftingForm}) is equal to eq. (\ref{eq:dampCoulomb})
371 > with $r$ replaced by $R_\textrm{c}$.  This term is {\it not} present
372 > in the Zahn potential, resulting in a potential discontinuity as
373 > particles cross $R_\textrm{c}$.  Second, the sign of the derivative
374 > portion is different.  The missing $v_\textrm{c}$ term would not
375 > affect molecular dynamics simulations (although the computed energy
376 > would be expected to have sudden jumps as particle distances crossed
377 > $R_c$).  The sign problem would be a potential source of errors,
378 > however.  In fact, it introduces a discontinuity in the forces at the
379 > cutoff, because the force function is shifted in the wrong direction
380 > and doesn't cross zero at $R_\textrm{c}$.
381  
382   Eqs. (\ref{eq:DSFPot}) and (\ref{eq:DSFForces}) result in an
383   electrostatic summation method that is continuous in both the
# Line 570 | Line 625 | tolerance (typically less than $1 \times 10^{-4}$ kcal
625   the energies and forces calculated.  Typical molecular mechanics
626   packages default this to a value dependent on the cutoff radius and a
627   tolerance (typically less than $1 \times 10^{-4}$ kcal/mol).  Smaller
628 < tolerances are typically associated with increased accuracy in the
629 < real-space portion of the summation.\cite{Essmann95} The default
630 < TINKER tolerance of $1 \times 10^{-8}$ kcal/mol was used in all SPME
628 > tolerances are typically associated with increased accuracy, but this
629 > usually means more time spent calculating the reciprocal-space portion
630 > of the summation.\cite{Perram88,Essmann95} The default TINKER
631 > tolerance of $1 \times 10^{-8}$ kcal/mol was used in all SPME
632   calculations, resulting in Ewald Coefficients of 0.4200, 0.3119, and
633   0.2476 \AA$^{-1}$ for cutoff radii of 9, 12, and 15 \AA\ respectively.
634  
# Line 851 | Line 907 | required for accurate reproduction of crystal dynamics
907   \begin{figure}
908   \centering
909   \includegraphics[width = \linewidth]{./comboSquare.pdf}
910 < \caption{Upper: Zoomed inset from figure \ref{fig:methodPS}.  As the damping value for the {\sc sf} potential increases, the low-frequency peaks red-shift.  Lower: Low-frequency correlated motions for NaCl crystals at 1000 K when using SPME and a simple damped Coulombic sum with damping coefficients ($\alpha$) ranging from 0.15 to 0.3 \AA$^{-1}$.  As $\alpha$ increases, the peaks are red-shifted toward and eventually beyond the values given by SPME.  The larger $\alpha$ values weaken the real-space electrostatics, explaining this shift towards less strongly correlated motions in the crystal.}
910 > \caption{Regions of spectra showing the low-frequency correlated motions for NaCl crystals at 1000 K using various electrostatic summation methods.  The upper plot is a zoomed inset from figure \ref{fig:methodPS}.  As the damping value for the {\sc sf} potential increases, the low-frequency peaks red-shift.  The lower plot is of spectra when using SPME and a simple damped Coulombic sum with damping coefficients ($\alpha$) ranging from 0.15 to 0.3 \AA$^{-1}$.  As $\alpha$ increases, the peaks are red-shifted toward and eventually beyond the values given by SPME.  The larger $\alpha$ values weaken the real-space electrostatics, explaining this shift towards less strongly correlated motions in the crystal.}
911   \label{fig:dampInc}
912   \end{figure}
913  

Diff Legend

Removed lines
+ Added lines
< Changed lines
> Changed lines