ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/multipole/dielectric_new.tex
Revision: 4408
Committed: Wed Mar 30 20:12:46 2016 UTC (8 years, 3 months ago) by gezelter
Content type: application/x-tex
File size: 67567 byte(s)
Log Message:
New figures, some edits

File Contents

# Content
1 \documentclass[aip,jcp,amsmath,amssymb,preprint]{revtex4-1}
2
3 \usepackage{graphicx}% Include figure files
4 \usepackage{dcolumn}% Align table columns on decimal point
5 \usepackage{multirow}
6 \usepackage{bm}
7 \usepackage{natbib}
8 \usepackage{times}
9 \usepackage{mathptmx}
10 \usepackage[version=3]{mhchem} % this is a great package for formatting chemical reactions
11 \usepackage{url}
12 \usepackage{braket}
13 \usepackage{tabularx}
14 \usepackage{rotating}
15 \usepackage{multirow}
16 \usepackage{booktabs}
17
18 \newcolumntype{Y}{>{\centering\arraybackslash}X}
19
20 \begin{document}
21
22 \title{Real space electrostatics for multipoles. III. Dielectric Properties}
23
24 \author{Madan Lamichhane}
25 \affiliation{Department of Physics, University
26 of Notre Dame, Notre Dame, IN 46556}
27 \author{Thomas Parsons}
28 \affiliation{Department of Chemistry and Biochemistry, University
29 of Notre Dame, Notre Dame, IN 46556}
30 \author{Kathie E. Newman}
31 \affiliation{Department of Physics, University
32 of Notre Dame, Notre Dame, IN 46556}
33 \author{J. Daniel Gezelter}
34 \email{gezelter@nd.edu.}
35 \affiliation{Department of Chemistry and Biochemistry, University
36 of Notre Dame, Notre Dame, IN 46556}
37
38 \date{\today}% It is always \today, today,
39 % but any date may be explicitly specified
40
41 \begin{abstract}
42 In the first two papers in this series, we developed new shifted
43 potential (SP), gradient shifted force (GSF), and Taylor shifted
44 force (TSF) real-space methods for multipole interactions in
45 condensed phase simulations. Here, we discuss the dielectric
46 properties of fluids that emerge from simulations using these
47 methods. Most electrostatic methods (including the Ewald sum)
48 require correction to the conducting boundary fluctuation formula
49 for the static dielectric constants, and we discuss the derivation
50 of these corrections for the new real space methods. For quadrupolar
51 fluids, the analogous material property is the quadrupolar
52 susceptibility. As in the dipolar case, the fluctuation formula for
53 the quadrupolar susceptibility has corrections that depend on the
54 electrostatic method being utillized. One of the most important
55 effects measured by both the static dielectric and quadrupolar
56 susceptibility is the ability to screen charges embedded in the
57 fluid. We use potentials of mean force between solvated ions to
58 discuss how geometric factors can lead to distance-dependent
59 screening in both quadrupolar and dipolar fluids.
60 \end{abstract}
61
62 \maketitle
63
64 \section{Introduction}
65
66 Over the past several years, there has been increasing interest in
67 pairwise or ``real space'' methods for computing electrostatic
68 interactions in condensed phase
69 simulations.\cite{Wolf99,Zahn02,Kast03,Beckd05,Ma05,Wu:2005nr,Fennell06,Fukuda:2013uq,Wang:2016kx}
70 These techniques were initially developed by Wolf {\it et al.} in
71 their work towards an $\mathcal{O}(N)$ Coulombic sum.\cite{Wolf99}
72 Wolf's method of using cutoff neutralization and electrostatic damping
73 is able to obtain excellent agreement with Madelung energies in ionic
74 crystals.\cite{Wolf99}
75
76 Zahn \textit{et al.}\cite{Zahn02} and Fennell and Gezelter extended
77 this method using shifted force approximations at the cutoff distance
78 in order to conserve total energy in molecular dynamics
79 simulations.\cite{Fennell06} Other recent advances in real-space
80 methods for systems of point charges have included explicit
81 elimination of the net multipole moments inside the cutoff sphere
82 around each charge site.\cite{Fukuda:2013uq,Wang:2016kx}
83
84 In the previous two papers in this series, we developed three
85 generalized real space methods: shifted potential (SP), gradient
86 shifted force (GSF), and Taylor shifted force (TSF).\cite{PaperI,
87 PaperII} These methods evaluate electrostatic interactions for
88 charges and higher order multipoles using a finite-radius cutoff
89 sphere. The neutralization and damping of local moments within the
90 cutoff sphere is a multipolar generalization of Wolf's sum. In the
91 GSF and TSF methods, additional terms are added to the potential
92 energy so that forces and torques also vanish smoothly at the cutoff
93 radius. This ensures that the total energy is conserved in molecular
94 dynamics simulations.
95
96 One of the most stringent tests of any new electrostatic method is the
97 fidelity with which that method can reproduce the bulk-phase
98 polarizability or equivalently, the dielectric properties of a
99 fluid. Before the advent of computer simulations, Kirkwood and Onsager
100 developed fluctuation formulae for the dielectric properties of
101 dipolar fluids.\cite{Kirkwood39,Onsagar36} Along with projections of
102 the frequency-dependent dielectric to zero frequency, these
103 fluctuation formulae are now widely used to predict the static
104 dielectric constant of simulated materials.
105
106 If we consider a system of dipolar or quadrupolar molecules under the
107 influence of an external field or field gradient, the net polarization
108 of the system will largely be proportional to the applied
109 perturbation.\cite{Chitanvis96, Stern-Feller03, SalvchovI14,
110 SalvchovII14} In simulations, the net polarization of the system is
111 also determined by the interactions \textit{between} the
112 molecules. Therefore the macroscopic polarizablity obtained from a
113 simulation depends on the details of the electrostatic interaction
114 methods that were employed in the simulation. To determine the
115 relevant physical properties of the multipolar fluid from the system
116 fluctuations, the interactions between molecules must be incorporated
117 into the formalism for the bulk properties.
118
119 In most simulations, bulk materials are treated using periodic
120 replicas of small regions, and this level of approximation requires
121 corrections to the fluctuation formulae that were derived for the bulk
122 fluids. In 1983 Neumann proposed a general formula for evaluating
123 dielectric properties of dipolar fluids using both Ewald and
124 real-space cutoff methods.\cite{NeumannI83} Steinhauser and Neumann
125 used this formula to evaluate the corrected dielectric constant for
126 the Stockmayer fluid using two different methods: Ewald-Kornfeld (EK)
127 and reaction field (RF) methods.\cite{NeumannII83}
128
129 Zahn \textit{et al.}\cite{Zahn02} utilized this approach and evaluated
130 the correction factor for using damped shifted charge-charge
131 kernel. This was later generalized by Izvekov \textit{et
132 al.},\cite{Izvekov08} who noted that the expression for the
133 dielectric constant reduces to widely-used conducting boundary formula
134 for real-space methods that have first derivatives that vanish at the
135 cutoff radius.
136
137 One of the primary topics of this paper is the derivation of
138 correction factors for the three new real space methods. We find that
139 the correction formulae for dipolar molecules depends not only on the
140 methodology being used, but also on whether the molecular dipoles are
141 treated using point charges or point dipoles. We derive correction
142 factors for both cases.
143
144 In quadrupolar fluids, the relationship between quadrupolar
145 susceptibility and the dielectric constant is not as straightforward
146 as in the dipolar case. The effective dielectric constant depends on
147 the geometry of the external (or internal) field
148 perturbation.\cite{Ernst92} Significant efforts have been made to
149 increase our understanding the dielectric properties of these
150 fluids,\cite{JeonI03,JeonII03,Chitanvis96} although a general
151 correction formula has not yet been developed.
152
153 In this paper we derive general formulae for calculating the
154 quadrupolar susceptibility of quadrupolar fluids. We also evaluate the
155 correction factor for SP, GSF, and TSF methods for quadrupolar fluids
156 interacting via point charges, point dipoles or directly through
157 quadrupole-quadrupole interactions.
158
159 We also calculate the screening behavior for two ions immersed in
160 multipolar fluids to estimate the distance dependence of charge
161 screening in both dipolar and quadrupolar fluids. We use three
162 distinct methods to compare our analytical results with computer
163 simulations:
164 \begin{enumerate}
165 \item responses of the fluid to external perturbations,
166 \item fluctuations of system multipole moments, and
167 \item potentials of mean force between solvated ions,
168 \end{enumerate}
169
170 \begin{figure}
171 \includegraphics[width=\linewidth]{Schematic}
172 \caption{Dielectric properties of a fluid measure the response to
173 external electric fields and gradients (left panel), or internal
174 fields and gradients generated by the molecules themselves (central
175 panel), or fields produced by embedded ions (right panel). The
176 dielectric constant measures all three responses in dipolar fluids.
177 In quadrupolar liquids, the relevant bulk property is the
178 quadrupolar susceptibility, and the geometry of the perturbation
179 determines the effective dielectric screening.}
180 \label{fig:schematic}
181 \end{figure}
182
183 Under the influence of weak external fields, the bulk polarization of
184 the system is primarily a linear response to the perturbation, where
185 proportionality constant depends on the electrostatic interactions
186 between the multipoles. The fluctuation formulae connect bulk
187 properties of the fluid to equilibrium fluctuations in the system
188 multipolar moments during a simulation. These fluctuations also depend
189 on the form of the electrostatic interactions between molecules.
190 Therefore, the connections between the actual bulk properties and both
191 the computed fluctuation and external field responses must be modified
192 accordingly.
193
194 The potential of mean force (PMF) allows calculation of an effective
195 dielectric constant or screening factor from the potential energy
196 between ions before and after dielectric material is introduced.
197 Computing the PMF between embedded point charges is an additional
198 check on the bulk properties computed via the other two methods.
199
200 \section{The Real-Space Methods}
201
202 In the first paper in this series, we derived interaction energies, as
203 well as expressions for the forces and torques for point multipoles
204 interacting via three new real-space methods.\cite{PaperI} The Taylor
205 shifted-force (TSF) method modifies the electrostatic kernel,
206 $f(r) = 1/r$, so that all forces and torques go smoothly to zero at
207 the cutoff radius,
208 \begin{equation}
209 U^{\text{TSF}}_{ab} = M_{a} M_{b} f_n(r).
210 \label{TSF}
211 \end{equation}
212 Here the multipole operator for site $a$, $M_{a}$, is expressed in
213 terms of the point charge, $C_{a}$, dipole, ${\bf D}_{a}$, and
214 quadrupole, $\mathsf{Q}_{a}$, for object $a$, etc. Because each of
215 the multipole operators includes a set of implied gradient operators,
216 an approximate electrostatic kernel, $f_n(r)$ is Taylor-expanded
217 around the cutoff radius, so that $n + 1$ derivatives vanish as
218 $r \rightarrow r_c$. This ensures smooth convergence of the energy,
219 forces, and torques as molecules leave and reenter each others cutoff
220 spheres. The order of the Taylor expansion is determined by the
221 multipolar order of the interaction. That is, smooth
222 quadrupole-quadrupole forces require the fifth derivative to vanish at
223 the cutoff radius, so the appropriate function Taylor expansion will
224 be of fifth order.
225
226 Following this procedure results in separate radial functions for each
227 of the distinct orientational contributions to the potential. For
228 example, in dipole-dipole interactions, the direct dipole dot product
229 ($\mathbf{D}_{a} \cdot \mathbf{D}_{b}$) is treated differently than
230 the dipole-distance dot products:
231 \begin{equation}
232 U_{\mathbf{D}_{a}\mathbf{D}_{b}}(r)= -\frac{1}{4\pi \epsilon_0} \left[ \left(
233 \mathbf{D}_{a} \cdot
234 \mathbf{D}_{b} \right) v_{21}(r) +
235 \left( \mathbf{D}_{a} \cdot \hat{\mathbf{r}} \right)
236 \left( \mathbf{D}_{b} \cdot \hat{\mathbf{r}} \right) v_{22}(r) \right]
237 \end{equation}
238 The two radial functions, $v_{21}(r)$ and $v_{22}(r)$ are both $1/r^3$
239 in standard electrostatics, but they become distinct in the TSF
240 method, the relative orientations for interacting molecules so that
241 forces and torques vanish smoothly as the molecules drift beyond the
242 cutoff radius.
243
244 A second and somewhat simpler approach involves shifting the gradient
245 of the Coulomb potential for each particular multipole order,
246 \begin{equation}
247 U^{\text{GSF}}_{ab} = \sum \left[ U(\mathbf{r}, \mathsf{A}, \mathsf{B}) -
248 U(r_c \hat{\mathbf{r}},\mathsf{A}, \mathsf{B}) - (r-r_c)
249 \hat{\mathbf{r}} \cdot \nabla U(r_c \hat{\mathbf{r}},\mathsf{A},
250 \mathsf{B}) \right]
251 \label{generic2}
252 \end{equation}
253 where the sum describes a separate force-shifting that is applied to
254 each orientational contribution to the energy, i.e. $v_{21}$ and
255 $v_{22}$ are shifted separately. In this expression,
256 $\hat{\mathbf{r}}$ is the unit vector connecting the two multipoles
257 ($a$ and $b$) in space, and $\mathsf{A}$ and $\mathsf{B}$ represent
258 the orientations the multipoles. Because this procedure is equivalent
259 to using the gradient of an image multipole placed at the cutoff
260 sphere for shifting the force, this method is called the gradient
261 shifted-force (GSF) approach.
262
263 Both the TSF and GSF approaches can be thought of as multipolar
264 extensions of the original damped shifted-force (DSF) approach that
265 was developed for point charges. There is also a multipolar extension
266 of the Wolf sum that is obtained by projecting an image multipole onto
267 the surface of the cutoff sphere, and including the interactions with
268 the central multipole and the image. This effectively shifts only the
269 total potential to zero at the cutoff radius,
270 \begin{equation}
271 U^{\text{SP}}_{ab} = \sum \left[ U(\mathbf{r}, \mathsf{A}, \mathsf{B}) -
272 U(r_c \hat{\mathbf{r}},\mathsf{A}, \mathsf{B}) \right]
273 \label{eq:SP}
274 \end{equation}
275 where the sum again describes separate potential shifting that is done
276 for each orientational contribution to the energy. The potential
277 energy between a central multipole and other multipolar sites goes
278 smoothly to zero as $r \rightarrow r_c$, but the forces and torques
279 obtained from this shifted potential (SP) approach are discontinuous
280 at $r_c$.
281
282 All three of the new real space methods share a common structure: the
283 various orientational contributions to multipolar interaction energies
284 require separate treatment of their radial functions, and these are
285 tabulated for both the raw Coulombic kernel ($1/r$) as well as the
286 damped kernel ($\mathrm{erfc}(\alpha r)/r$), in the first paper of this
287 series.\cite{PaperI} The second paper in this series evaluated the
288 fidelity with which the three new methods reproduced Ewald-based
289 results for a number of model systems.\cite{PaperII} One of the major
290 findings was that moderately-damped GSF simulations produced nearly
291 identical behavior with Ewald-based simulations, but the real-space
292 methods scale linearly with system size.
293
294 \section{Dipolar Fluids and the Dielectric Constant}
295
296 Dielectric properties of a fluid arise mainly from responses of the
297 fluid to either applied fields or transient fields internal to the
298 fluid. In response to an applied field, the molecules have electronic
299 polarizabilities, changes to internal bond lengths, and reorientations
300 towards the direction of the applied field. There is an added
301 complication that in the presence of external field, the perturbation
302 experienced by any single molecule is not only due to external field
303 but also to the fields produced by the all other molecules in the
304 system.
305
306 \subsection{Response to External Perturbations}
307
308 In the presence of uniform electric field $\mathbf{E}$, an individual
309 molecule with a permanent dipole moment $p_o$ will realign along the
310 direction of the field with an average polarization given by
311 \begin{equation}
312 \braket{\mathbf{p}} = \epsilon_0 \alpha_p \mathbf{E},
313 \end{equation}
314 where $\alpha_p = {p_o}^2 / 3 \epsilon_0 k_B T$ is the contribution to
315 molecular polarizability due solely to reorientation dynamics.
316 Because the applied field must overcome thermal motion, the
317 orientational polarization depends inversely on the temperature.
318
319 Likewise, a condensed phase system of permanent dipoles will also
320 polarize along the direction of an applied field. The polarization
321 density of the system is
322 \begin{equation}
323 \textbf{P} = \epsilon_o \alpha_{D} \mathbf{E}.
324 \label{pertDipole}
325 \end{equation}
326 The constant $\alpha_D$ is the macroscopic polarizability, which is an
327 emergent property of the dipolar fluid. Note that the polarizability,
328 $\alpha_D$ is distinct from the dipole susceptibility, $\chi_D$,
329 which is the quantity most directly related to the static dielectric
330 constant, $\epsilon = 1 + \chi_D$.
331
332 \subsection{Fluctuation Formula}
333 % Using the Ewald sum under tin-foil boundary conditions, $\epsilon$ can
334 % be calculated for systems of orientationally-polarized molecules,
335 % \begin{equation}
336 % \epsilon = 1 + \frac{\langle M^2\rangle}{3\epsilon_0k_\textrm{B}TV},
337 % \label{eq:staticDielectric}
338 % \end{equation}
339 % where $\epsilon_0$ is the permittivity of free space and
340 % $\langle M^2\rangle$ is the fluctuation of the system dipole
341 % moment.\cite{Allen89} The numerator in the fractional term in
342 % equation (\ref{eq:staticDielectric}) is identical to the quantity
343 % calculated in the finite-system Kirkwood $g$ factor ($G_k$):
344 % \begin{equation}
345 % G_k = \frac{\langle M^2\rangle}{N\mu^2},
346 % \label{eq:KirkwoodFactor}
347 % \end{equation}
348 % where $\mu$ is the dipole moment of a single molecule of the
349 % homogeneous system.\cite{NeumannI83,NeumannII83,Neumann84,Neumann85} The
350 % fluctuation term in both equation (\ref{eq:staticDielectric}) and
351 % (\ref{eq:KirkwoodFactor}) is calculated as follows,
352 % \begin{equation}
353 % \begin{split}
354 % \langle M^2\rangle &= \langle\bm{M}\cdot\bm{M}\rangle
355 % - \langle\bm{M}\rangle\cdot\langle\bm{M}\rangle \\
356 % &= \langle M_x^2+M_y^2+M_z^2\rangle
357 % - (\langle M_x\rangle^2 + \langle M_x\rangle^2
358 % + \langle M_x\rangle^2).
359 % \end{split}
360 % \label{eq:fluctBoxDipole}
361 % \end{equation}
362 % This fluctuation term can be accumulated during a simulation; however,
363 % it converges rather slowly, thus requiring multi-nanosecond simulation
364 % times.\cite{Horn04} In the case of tin-foil boundary conditions, the
365 % dielectric/surface term of the Ewald summation is equal to zero.
366
367 For a system of dipolar molecules at thermal equilibrium, we can
368 define both a system dipole moment, $\mathbf{M} = \sum_i \mathbf{p}_i$
369 as well as a dipole polarization density,
370 $\mathbf{P} = \braket{\mathbf{M}}/V$.
371
372 In the presence of applied field $\mathbf{E}$, the polarization
373 density can be expressed in terms of fluctuations in the net dipole
374 moment,
375 \begin{equation}
376 \mathbf{P} = \epsilon_o \frac{\braket{\mathbf{M}^2}-{\braket{\mathbf{M}}}^2}{3 \epsilon_o V k_B T}\bf{E}
377 \label{flucDipole}
378 \end{equation}
379 This has structural similarity with the Boltzmann average for the
380 polarization of a single molecule. Here
381 $ \braket{\mathbf{M}^2}-{\braket{\mathbf{M}}}^2$ measures fluctuations
382 in the net dipole moment,
383 \begin{equation}
384 \langle \mathbf{M}^2 \rangle - \langle \mathbf{M} \rangle^2 =
385 \langle M_x^2+M_y^2+M_z^2 \rangle - \left( \langle M_x \rangle^2 +
386 \langle M_y \rangle^2 +
387 \langle M_z \rangle^2 \right).
388 \label{eq:flucDip}
389 \end{equation}
390 For the limiting case $\textbf{E} \rightarrow 0 $, the ensemble
391 average of both the net dipole moment $\braket{\mathbf{M}}$ and
392 dipolar polarization $\bf{P}$ tends to vanish but
393 $\braket{\mathbf{M}^2}$ does not. The macroscopic dipole
394 polarizability can therefore be written as,
395 \begin{equation}
396 \alpha_D = \frac{\braket{\mathbf{M}^2}-{\braket{\mathbf{M}}}^2}{3 \epsilon_o V k_B T}
397 \end{equation}
398 This relationship between a macroscopic property of dipolar material
399 and microscopic fluctuations is true even when the applied field
400 $ \textbf{E} \rightarrow 0 $.
401
402 \subsection{Correction Factors}
403 \label{sec:corrFactor}
404 In the presence of a uniform external field $ \mathbf{E}^\circ$, the
405 total electric field at $\mathbf{r}$ depends on the polarization
406 density at all other points in the system,\cite{NeumannI83}
407 \begin{equation}
408 \mathbf{E}(\mathbf{r}) = \mathbf{E}^\circ(\mathbf{r}) +
409 \frac{1}{4\pi\epsilon_o} \int d\mathbf{r}^\prime
410 \mathbf{T}(\mathbf{r}-\mathbf{r}^\prime)\cdot
411 {\mathbf{P}(\mathbf{r}^\prime)}.
412 \label{eq:localField}
413 \end{equation}
414 $\mathbf{T}$ is the dipole interaction tensor connecting dipoles at
415 $\mathbf{r}^\prime$ with the point of interest ($\mathbf{r}$).
416
417 In simulations of dipolar fluids, the molecular dipoles may be
418 represented either by closely-spaced point charges or by
419 point dipoles (see Fig. \ref{fig:tensor}).
420 \begin{figure}
421 \includegraphics[width=\linewidth]{Tensors}
422 \caption{In the real-space electrostatic methods, the molecular dipole
423 tensor, $\mathbf{T}_{\alpha\beta}(r)$, is not the same for
424 charge-charge interactions as for point dipoles (left panel). The
425 same holds true for the molecular quadrupole tensor (right panel),
426 $\mathbf{T}_{\alpha\beta\gamma\delta}(r)$, which can have distinct
427 forms if the molecule is represented by charges, dipoles, or point
428 quadrupoles.}
429 \label{fig:tensor}
430 \end{figure}
431 In the case where point charges are interacting via an electrostatic
432 kernel, $v(r)$, the effective {\it molecular} dipole tensor,
433 $\mathbf{T}$ is obtained from two successive applications of the
434 gradient operator to the electrostatic kernel,
435 \begin{align}
436 \mathbf{T}_{\alpha \beta}(r) =& \nabla_\alpha \nabla_\beta
437 \left(v(r)\right) \\
438 =& \delta_{\alpha \beta}
439 \left(\frac{1}{r} v^\prime(r) \right) + \frac{r_{\alpha}
440 r_{\beta}}{r^2} \left( v^{\prime \prime}(r) - \frac{1}{r}
441 v^{\prime}(r) \right)
442 \label{dipole-chargeTensor}
443 \end{align}
444 where $v(r)$ may be either the bare kernel ($1/r$) or one of the
445 modified (Wolf or DSF) kernels. This tensor describes the effective
446 interaction between molecular dipoles ($\mathbf{D}$) in Gaussian units
447 as $-\mathbf{D} \cdot \mathbf{T} \cdot \mathbf{D}$.
448
449 When utilizing any of the three new real-space methods for point
450 \textit{dipoles}, the tensor is explicitly constructed,
451 \begin{equation}
452 \mathbf{T}_{\alpha \beta}(r) = \delta_{\alpha \beta} v_{21}(r) +
453 \frac{r_{\alpha} r_{\beta}}{r^2} v_{22}(r)
454 \label{dipole-diopleTensor}
455 \end{equation}
456 where the functions $v_{21}(r)$ and $v_{22}(r)$ depend on the level of
457 the approximation.\cite{PaperI,PaperII} Although the Taylor-shifted
458 (TSF) and gradient-shifted (GSF) models produce to the same $v(r)$
459 function for point charges, they have distinct forms for the
460 dipole-dipole interaction.
461
462 Using a constitutive relation, the polarization density
463 $\mathbf{P}(\mathbf{r})$ is given by,
464 \begin{equation}
465 \mathbf{P}(\mathbf{r}) = \epsilon_o \chi_D
466 \left(\mathbf{E}^\circ(\mathbf{r}) + \frac{1}{4\pi\epsilon_o} \int
467 d\mathbf{r}^\prime \mathbf{T}(\mathbf{r}-\mathbf{r}^\prime ) \cdot \mathbf{P}(\mathbf{r}^\prime)\right).
468 \label{constDipole}
469 \end{equation}
470 Note that $\chi_D$ explicitly depends on the details of the dipole
471 interaction tensor. Neumann \textit{et al.}
472 \cite{NeumannI83,NeumannII83,Neumann84,Neumann85} derived an elegant
473 way to modify the fluctuation formula to correct for approximate
474 interaction tensors. This correction was derived using a Fourier
475 representation of the interaction tensor,
476 $\tilde{\mathbf{T}}(\mathbf{k})$, and involves the quantity,
477 \begin{equation}
478 A = \frac{3}{4\pi}\tilde{\mathbf{T}}(0) = \frac{3}{4\pi} \int_V
479 d\mathbf{r} \mathbf{T}(\mathbf{r})
480 \end{equation}
481 which is the $k \rightarrow 0$ limit of $\tilde{\mathbf{T}}$. Using
482 this quantity (originally called $Q$ in
483 refs. \onlinecite{NeumannI83,NeumannII83,Neumann84,Neumann85}), the
484 dielectric constant can be computed
485 \begin{equation}
486 \epsilon = \frac{3+(A + 2)(\epsilon_{CB}-1)}{3 + (A -1)(\epsilon_{CB}-1)}
487 \label{correctionFormula}
488 \end{equation}
489 where $\epsilon_{CB}$ is the widely-used conducting boundary
490 expression for the dielectric constant,
491 \begin{equation}
492 \epsilon_{CB} = 1 + \frac{\braket{\bf{M}^2}-{\braket{\bf{M}}}^2}{3
493 \epsilon_o V k_B T}
494 \label{conductingBoundary}
495 \end{equation}
496 Eqs. (\ref{correctionFormula}) and (\ref{conductingBoundary}) allows
497 estimation of the static dielectric constant from fluctuations easily
498 computed directly from simulations.
499
500 % Here $\chi^*_D$ is a dipolar susceptibility can be expressed in terms of dielectric constant as $ \chi^*_D = \epsilon - 1$ which different than macroscopic dipolar polarizability $\alpha_D$ in the sections \ref{subsec:perturbation} and \ref{subsec:fluctuation}. We can split integral into two parts: singular part i.e $|\textbf{r}-\textbf{r}'|\rightarrow 0 $ and non-singular part i.e $|\textbf{r}-\textbf{r}'| > 0 $ . The singular part of the integral can be written as,\cite{NeumannI83, Jackson98}
501 % \begin{equation}
502 % \frac{1}{4\pi\epsilon_o} \int_{|\textbf{r}-\textbf{r}'| \rightarrow 0} d^3r'\; \textbf{T}(\textbf{r}-\textbf{r}')\cdot {\textbf{P}(\textbf{r}')} = - \frac{\textbf{P}(\textbf{r})}{3\epsilon_o}
503 % \label{singular}
504 % \end{equation}
505 % Substituting equation (\ref{singular}) in the equation (\ref{constDipole}) we get,
506 % \begin{equation}
507 % \textbf{P}(\textbf{r}) = 3 \epsilon_o\; \frac{\chi^*_D}{\chi^*_D + 3} \left(\textbf{E}^o(\textbf{r}) + \frac{1}{4\pi\epsilon_o} \int_{|\textbf{r}-\textbf{r}'| > 0} d^3r'\; \textbf{T}(\textbf{r}-\textbf{r}')\cdot {\textbf{P}(\textbf{r}')}\right).
508 % \end{equation}
509 % For both polarization and electric field homogeneous, this can be easily solved using Fourier transformation,
510 % \begin{equation}
511 % \textbf{P}(\mathbf{k}) = 3 \epsilon_o\; \frac{\chi^*_D}{\chi^*_D + 3} \left(1- \frac{3}{4\pi}\;\frac{\chi^*_D}{\chi^*_D + 3}\; \textbf{T}(\mathbf{k})\right)^{-1}\textbf{E}^o(\mathbf{k}).
512 % \end{equation}
513 % For homogeneous applied field Fourier component is non-zero only if $\kappa = 0$. Therefore,
514 % \begin{equation}
515 % \textbf{P}(0) = 3 \epsilon_o\; \frac{\chi^*_D}{\chi^*_D + 3} \left(1- \frac{\chi^*_D}{\chi^*_D + 3}\; A_{dipole})\right)^{-1}\textbf{E}^o({0}).
516 % \label{fourierDipole}
517 % \end{equation}
518 % where $A_{dipole}=\frac{3}{4\pi}T(0) = \frac{3}{4\pi} \int_V d^3r\;T(r)$. Now equation (\ref{fourierDipole}) can be compared with equation (\ref{flucDipole}). Therefore,
519 % \begin{equation}
520 % \frac{\braket{\bf{M}^2}-{\braket{\bf{M}}}^2}{3 \epsilon_o V k_B T} = \frac{3\;\chi^*_D}{\chi^*_D + 3} \left(1- \frac{\chi^*_D}{\chi^*_D + 3}\; A_{dipole})\right)^{-1}
521 % \end{equation}
522 % Substituting $\chi^*_D = \epsilon-1$ and $ \frac{\braket{\bf{M}^2}-{\braket{\bf{M}}}^2}{3 \epsilon_o V k_B T} = \epsilon_{CB}-1 = \alpha_D$ in above equation we get,
523 % \begin{equation}
524 % \epsilon = \frac{3+(A_{dipole} + 2)(\epsilon_{CB}-1)}{3+(A_{dipole} -1)(\epsilon_{CB}-1)} = \frac{3+(A_{dipole} + 2)\alpha_D}{3+(A_{dipole} -1)\alpha_D}
525 % \label{correctionFormula}
526 % \end{equation}
527
528 We have utilized the Neumann \textit{et al.} approach for the three
529 new real-space methods, and obtain method-dependent correction
530 factors. The expression for the correction factor also depends on
531 whether the simulation involves point charges or point dipoles to
532 represent the molecular dipoles. These corrections factors are
533 listed in Table \ref{tab:A}.
534 \begin{table}
535 \caption{Expressions for the dipolar correction factor ($A$) for the
536 real-space electrostatic methods in terms of the damping parameter
537 ($\alpha$) and the cutoff radius ($r_c$). The Ewald-Kornfeld result
538 derived in Refs. \onlinecite{Adams80,Adams81,NeumannI83} is shown
539 for comparison using the Ewald convergence parameter ($\kappa$)
540 and the real-space cutoff value ($r_c$). }
541 \label{tab:A}
542 \begin{tabular}{l|c|c}
543 \toprule
544 Method & point charges & point dipoles \\
545 \colrule
546 Spherical Cutoff (SC) & $\mathrm{erf}(r_c \alpha) - \frac{2 \alpha r_c}{\sqrt{\pi}} e^{-\alpha^2 r_c^2}$ & $\mathrm{erf}(r_c \alpha) - \frac{2 \alpha r_c}{\sqrt{\pi}} e^{-\alpha^2 r_c^2}$ \\
547 Shifted Potental (SP) & $ \mathrm{erf}(r_c \alpha) - \frac{2 \alpha r_c}{\sqrt{\pi}} e^{-\alpha^2 r_c^2}$ & $\mathrm{erf}(r_c \alpha) -\frac{2 \alpha r_c}{\sqrt{\pi}}\left(1+\frac{2\alpha^2 {r_c}^2}{3} \right)e^{-\alpha^2{r_c}^2} $\\
548 Gradient-shifted (GSF) & 1 & $\mathrm{erf}(\alpha r_c)-\frac{2 \alpha r_c}{\sqrt{\pi}} \left(1 + \frac{2 \alpha^2 r_c^2}{3} + \frac{\alpha^4 r_c^4}{3}\right)e^{-\alpha^2 r_c^2} $ \\
549 Taylor-shifted (TSF) & 1 & 1 \\ \colrule
550 Ewald-Kornfeld (EK) & \multicolumn{2}{c}{$\mathrm{erf}(r_c \kappa) -
551 \frac{2 \kappa r_c}{\sqrt{\pi}} e^{-\kappa^2
552 r_c^2}$} \\
553 \botrule
554 \end{tabular}
555 \end{table}
556 Note that for point charges, the GSF and TSF methods produce estimates
557 of the dielectric that need no correction, and the TSF method likewise
558 needs no correction for point dipoles.
559
560 \section{Quadrupolar Fluids and the Quadrupolar Susceptibility}
561 \subsection{Response to External Perturbations}
562
563 A molecule with a permanent quadrupole, $\mathsf{q}$, will align in
564 the presence of an electric field gradient $\nabla\mathbf{E}$. The
565 anisotropic polarization of the quadrupole is given
566 by,\cite{AduGyamfi78,AduGyamfi81}
567 \begin{equation}
568 \braket{\mathsf{q}} - \frac{\mathbf{I}}{3}
569 \mathrm{Tr}(\mathsf{q}) = \epsilon_o \alpha_q \nabla\mathbf{E},
570 \end{equation}
571 where $\alpha_q = q_o^2 / 15 \epsilon_o k_B T $ is a molecular quadrupole
572 polarizablity and $q_o$ is an effective quadrupole moment for the molecule,
573 \begin{equation}
574 q_o^2 = 3 \mathsf{q}:\mathsf{q} - \mathrm{Tr}(\mathsf{q})^2.
575 \end{equation}
576
577 In the presence of an external field gradient, a system of quadrupolar
578 molecules also organizes with an anisotropic polarization,
579 \begin{equation}
580 \mathsf{Q} - \frac{\mathbf{I}}{3} \mathrm{Tr}(\mathsf{Q}) = \epsilon_o
581 \alpha_Q \nabla\mathbf{E}
582 \end{equation}
583 where $\mathsf{Q}$ is the traced quadrupole density of the system and
584 $\alpha_Q$ is a macroscopic quadrupole polarizability which has
585 dimensions of $\mathrm{length}^{-2}$. Equivalently, the traceless form
586 may be used,
587 \begin{equation}
588 \mathsf{\Theta} = 3 \epsilon_o \alpha_Q \nabla\mathbf{E},
589 \label{pertQuad}
590 \end{equation}
591 where
592 $\mathsf{\Theta} = 3\mathsf{Q} - \mathbf{I} \mathrm{Tr}(\mathsf{Q})$
593 is the traceless tensor that also describes the system quadrupole
594 density. It is this tensor that will be utilized to derive correction
595 factors below.
596
597 \subsection{Fluctuation Formula}
598 As in the dipolar case, we may define a system quadrupole moment,
599 $\mathsf{M}_Q = \sum_i \mathsf{q}_i$ and the traced quadrupolar
600 density, $\mathsf{Q} = \mathsf{M}_Q / V$. A fluctuation formula can
601 be written for a system comprising quadrupolar
602 molecules,\cite{LoganI81,LoganII82,LoganIII82}
603 \begin{equation}
604 \mathsf{Q} - \frac{\mathbf{I}}{3} \mathrm{Tr}(\mathsf{Q}) = \epsilon_o
605 \frac{\braket{\mathsf{M}_Q^2}-{\braket{\mathsf{M}_Q}}^2}{15 \epsilon_o
606 V k_B T} \cdot \nabla\mathbf{E}.
607 \label{flucQuad}
608 \end{equation}
609 Some care is needed in the definitions of the averaged quantities. These
610 refer to the effective quadrupole moment of the system, and they are
611 computed as follows,
612 \begin{align}
613 \braket{\mathsf{M}_Q^2} &= \braket{3 \mathsf{M}_Q:\mathsf{M}_Q -
614 \mathrm{Tr}(\mathsf{M}_Q)^2 }\\
615 \braket{\mathsf{M}_Q}^2 &= 3 \braket{\mathsf{M}_Q}:\braket{\mathsf{M}_Q} -
616 \mathrm{Tr}(\braket{\mathsf{M}_Q})^2
617 \label{eq:flucQuad}
618 \end{align}
619 The macroscopic quadrupolarizability is given by,
620 \begin{equation}
621 \alpha_Q = \frac{\braket{\mathsf{M}_Q^2}-{\braket{\mathsf{M}_Q}}^2}{15 \epsilon_o V k_B T}
622 \label{propConstQuad}
623 \end{equation}
624 This relationship between a macroscopic property of a quadrupolar
625 fluid and microscopic fluctuations should be true even when the
626 applied field gradient $\nabla \mathbf{E} \rightarrow 0$.
627
628 \subsection{Correction Factors}
629 In this section we generalize the treatment of Neumann \textit{et al.}
630 for quadrupolar fluids. Interactions involving multiple quadrupoles
631 are rank 4 tensors, and we therefore describe quantities in this
632 section using Einstein notation.
633
634 In the presence of a uniform external field gradient,
635 $\partial_\alpha {E}^\circ_\beta $, the total field gradient at
636 $\mathbf{r}$ depends on the quadrupole polarization density at all
637 other points in the system,
638 \begin{equation}
639 \partial_\alpha E_\beta(\mathbf{r}) = \partial_\alpha
640 {E}^\circ_\beta(\mathbf{r}) + \frac{1}{4\pi \epsilon_o}\int
641 T_{\alpha\beta\gamma\delta}({\mathbf{r}-\mathbf{r}^\prime})
642 Q_{\gamma\delta}(\mathbf{r}^\prime) d\mathbf{r}^\prime
643 \label{gradMaxwell}
644 \end{equation}
645 where and $T_{\alpha\beta\gamma\delta}$ is the quadrupole interaction
646 tensor connecting quadrupoles at $\mathbf{r}^\prime$ with the point of
647 interest ($\mathbf{r}$).
648
649 % \begin{figure}
650 % \includegraphics[width=3in]{QuadrupoleFigure}
651 % \caption{With the real-space electrostatic methods, the molecular
652 % quadrupole tensor, $\mathbf{T}_{\alpha\beta\gamma\delta}(r)$,
653 % governing interactions between molecules is not the same for
654 % quadrupoles represented via sets of charges, point dipoles, or by
655 % single point quadrupoles.}
656 % \label{fig:quadrupolarFluid}
657 % \end{figure}
658
659 In simulations of quadrupolar fluids, the molecular quadrupoles may be
660 represented by closely-spaced point charges, by multiple point
661 dipoles, or by a single point quadrupole (see
662 Fig. \ref{fig:tensor}). In the case where point charges are
663 interacting via an electrostatic kernel, $v(r)$, the effective
664 molecular quadrupole tensor can obtained from four successive
665 applications of the gradient operator to the electrostatic kernel,
666 \begin{eqnarray}
667 T_{\alpha\beta\gamma\delta}(\mathbf{r}) &=& \nabla_\alpha \nabla_\beta
668 \nabla_\gamma \nabla_\delta v(r) \\
669 &=& \left(\delta_{\alpha\beta}\delta_{\gamma\delta} +
670 \delta_{\alpha\gamma}\delta_{\beta\delta}+
671 \delta_{\alpha\delta}\delta_{\beta\gamma}\right)\left(-\frac{v^\prime(r)}{r^3}
672 + \frac{v^{\prime \prime}(r)}{r^2}\right) \nonumber \\
673 & &+ \left(\delta_{\alpha\beta} r_\gamma r_\delta + 5 \mathrm{~permutations}
674 \right) \left(\frac{3v^\prime(r)}{r^5}-\frac{3v^{\prime \prime}(r)}{r^4} +
675 \frac{v^{\prime \prime \prime}(r)}{r^3}\right) \nonumber \\
676 & &+ r_\alpha r_\beta r_\gamma r_\delta
677 \left(-\frac{15v^\prime(r)}{r^7}+\frac{15v^{\prime \prime}(r)}{r^6}-\frac{6v^{\prime
678 \prime \prime}(r)}{r^5} + \frac{v^{\prime \prime \prime \prime}(r)}{r^4}\right),
679 \label{quadCharge}
680 \end{eqnarray}
681 where $v(r)$ can either be the electrostatic kernel ($1/r$) or one of
682 the modified (Wolf or DSF) kernels.
683
684 Similarly, when representing quadrupolar molecules with multiple point
685 dipoles, the molecular quadrupole interaction tensor can be obtained
686 using two successive applications of the gradient operator to the
687 dipole interaction tensor,
688 \begin{eqnarray}
689 T_{\alpha\beta\gamma\delta}(\mathbf{r}) &=& \nabla_\alpha \nabla_\beta
690 v_{\gamma\delta}(\mathbf{r}) \\
691 & = & \delta_{\alpha\beta}\delta_{\gamma\delta} \frac{v^\prime_{21}(r)}{r} +
692 \left(\delta_{\alpha\gamma}\delta_{\beta\delta}+
693 \delta_{\alpha\delta}\delta_{\beta\gamma}\right)\frac{v_{22}(r)}{r^2}
694 \nonumber\\
695 & &+ \delta_{\gamma\delta} r_\alpha r_\beta
696 \left(\frac{v^{\prime \prime}_{21}(r)}{r^2}-\frac{v^\prime_{21}(r)}{r^3} \right)\nonumber \\
697 & &+\left(\delta_{\alpha\beta} r_\gamma r_\delta +
698 \delta_{\alpha\gamma} r_\beta r_\delta +\delta_{\alpha\delta}
699 r_\gamma r_\beta + \delta_{\beta\gamma} r_\alpha r_\delta
700 +\delta_{\beta\delta} r_\alpha r_\gamma \right)
701 \left(\frac{v^\prime_{22}(r)}{r^3}-\frac{2v_{22}(r)}{r^4}\right)
702 \nonumber \\
703 & &+ r_\alpha r_\beta r_\gamma r_\delta
704 \left(\frac{v^{\prime
705 \prime}_{22}(r)}{r^4}-\frac{5v^\prime_{22}(r)}{r^5}+\frac{8v_{22}(r)}{r^6}\right),
706 \label{quadDip}
707 \end{eqnarray}
708 where $v_{\gamma\delta}(\mathbf{r})$ is a dipole-dipole interaction
709 tensor that depends on the level of the
710 approximation.\cite{PaperI,PaperII} Similarly $v_{21}(r)$ and
711 $v_{22}(r)$ are the radial function for different real space cutoff
712 methods defined in the first paper in this series.\cite{PaperI}
713
714 For quadrupolar liquids modeled using point quadrupoles, the
715 interaction tensor can be constructed as,
716 \begin{eqnarray}
717 T_{\alpha\beta\gamma\delta}(\mathbf{r}) & = &
718 \left(\delta_{\alpha\beta}\delta_{\gamma\delta}
719 +
720 \delta_{\alpha\gamma}\delta_{\beta\delta}+
721 \delta_{\alpha\delta}\delta_{\beta\gamma}\right)v_{41}(r)
722 + \delta_{\gamma\delta}
723 r_\alpha r_\beta
724 \frac{v_{42}(r)}{r^2}
725 \nonumber \\
726 & & + r_\alpha r_\beta r_\gamma r_\delta \left(\frac{v_{43}(r)}{r^4}\right),
727 \label{quadRadial}
728 \end{eqnarray}
729 where again $v_{41}(r)$, $v_{42}(r)$, and $v_{43}(r)$ are radial
730 functions defined in Paper I of the series. \cite{PaperI} Note that
731 these radial functions have different functional forms depending on
732 the level of approximation being employed.
733
734 The integral in equation (\ref{gradMaxwell}) can be divided into two
735 parts, $|\mathbf{r}-\mathbf{r}^\prime|\rightarrow 0 $ and
736 $|\mathbf{r}-\mathbf{r}^\prime|> 0$. Since the self-contribution to
737 the field gradient vanishes at the singularity (see the supporting
738 information), equation (\ref{gradMaxwell}) can be written as,
739 \begin{equation}
740 \partial_\alpha E_\beta(\mathbf{r}) = \partial_\alpha {E}^\circ_\beta(\mathbf{r}) +
741 \frac{1}{4\pi \epsilon_o}\int\limits_{|\mathbf{r}-\mathbf{r}^\prime|> 0 }
742 T_{\alpha\beta\gamma\delta}(\mathbf{r}-\mathbf{r}^\prime)
743 {Q}_{\gamma\delta}(\mathbf{r}^\prime) d\mathbf{r}^\prime.
744 \end{equation}
745 If $\mathbf{r} = \mathbf{r}^\prime$ is excluded from the integration,
746 the total gradient can be most easily expressed in terms of
747 traceless quadrupole density as below,\cite{LoganI81}
748 \begin{equation}
749 \partial_\alpha E_\beta(\mathbf{r}) = \partial_\alpha
750 {E}^\circ_\beta(\mathbf{r}) + \frac{1}{12\pi
751 \epsilon_o}\int\limits_{|\mathbf{r}-\mathbf{r}^\prime|> 0 }
752 T_{\alpha\beta\gamma\delta}(\mathbf{r}-\mathbf{r}^\prime) \Theta_{\gamma\delta}(\mathbf{r}') d\mathbf{r}^\prime,
753 \end{equation}
754 where
755 $\Theta_{\alpha\beta} = 3Q_{\alpha\beta} - \delta_{\alpha\beta}Tr(Q)$
756 is the traceless quadrupole density. In analogy to
757 Eq. (\ref{pertQuad}) above, the quadrupolarization density may now be
758 related to the actual quadrupolar susceptibility, $\chi_Q$,
759 \begin{equation}
760 \frac{1}{3}{\Theta}_{\alpha\beta}(\mathbf{r}) = \epsilon_o {\chi}_Q
761 \left(\partial_\alpha {E}^\circ_\beta(\mathbf{r}) + \frac{1}{12\pi
762 \epsilon_o}\int\limits_{|\mathbf{r}-\mathbf{r}^\prime|> 0 }
763 T_{\alpha\beta\gamma\delta}(\mathbf{r}-\mathbf{r}^\prime)
764 \Theta_{\gamma\delta}(\mathbf{r}^\prime) d\mathbf{r}^\prime \right).
765 \end{equation}
766 For toroidal boundaries and with a uniform imposed
767 $\partial_\alpha E^\circ_\beta$, the quadrupole density
768 ${\Theta}_{\alpha\beta}$ will be uniform over the entire space. After
769 performing a Fourier transform (see the Appendix in
770 ref. \onlinecite{NeumannI83}) we obtain,
771 \begin{equation}
772 \frac{1}{3}\tilde{\Theta}_{\alpha\beta}(\mathbf{k})=
773 \epsilon_o {\chi}_Q \left[{\partial_\alpha
774 \tilde{E}^\circ_\beta}(\mathbf{k})+ \frac{1}{12\pi
775 \epsilon_o} \tilde{T}_{\alpha\beta\gamma\delta}(\mathbf{k})
776 \tilde{\Theta}_{\gamma\delta}(\mathbf{k})\right].
777 \end{equation}
778
779 The only surviving components of the quadrupole polarization will be
780 in the direction of the applied field gradient, so we may simplify
781 this expression considerably. That is, the only relevant polarization
782 term is $\tilde{\Theta}_{\alpha\beta}(\mathbf{k})$ and the only relevant
783 tensor contributions will have doubly-repeated indices, e.g.
784 $\tilde{T}_{\alpha\beta\alpha\beta}(\mathbf{k})$. Therefore we may
785 consider each component of the interaction tensor separately in a
786 scalar expression,
787 \begin{eqnarray}
788 \frac{1}{3}\tilde{\Theta}_{\alpha\beta}(\mathbf{k}) &=& \epsilon_o
789 {\chi}_Q
790 \left[{\partial_\alpha
791 \tilde{E}^\circ_\beta}(\mathbf{k})+
792 \frac{1}{12\pi
793 \epsilon_o}\tilde{T}_{\alpha\beta\alpha\beta}(\mathbf{k})
794 {\tilde{\Theta}}_{\alpha\beta}(\mathbf{k})\right]
795 \nonumber \\
796 &=& \epsilon_o {\chi}_Q \left(1-\frac{1}{4\pi} {\chi}_Q \tilde{T}_{\alpha\beta\alpha\beta}(\mathbf{k})\right)^{-1}
797 {\partial_\alpha \tilde{E}^\circ_\beta}(\mathbf{k})
798 \label{fourierQuad}
799 \end{eqnarray}
800 If the applied field gradient is homogeneous over the
801 entire volume, ${\partial_ \alpha \tilde{E}^\circ_\beta}(\mathbf{k}) = 0 $ except at
802 $ \mathbf{k} = 0$. As in the dipolar case, the only relevant
803 contribution of the interaction tensor will also be when $\mathbf{k} = 0$.
804 Therefore equation (\ref{fourierQuad}) can be written as,
805 \begin{equation}
806 \frac{1}{3}{\tilde{\Theta}}_{\alpha\beta}({0}) = \epsilon_o {\chi}_Q
807 \left(1-\frac{1}{4\pi} {\chi}_Q \tilde{T}_{\alpha\beta\alpha\beta}({0})\right)^{-1} \partial_\alpha \tilde{E}^\circ_\beta({0})
808 \label{fourierQuad2}
809 \end{equation}
810 where $\tilde{T}_{\alpha\beta\alpha\beta}({0})$ can be evaluated as,
811 \begin{equation}
812 \tilde{T}_{\alpha\beta\alpha\beta}({0}) = \int_V {T}_{\alpha\beta\alpha\beta}(\mathbf{r})d\mathbf{r}
813 \label{realTensorQaud}
814 \end{equation}
815 We may now define a quantity to represent the contribution from the
816 $\mathbf{k} = 0$ portion of the interaction tensor,
817 \begin{equation}
818 B = \frac{1}{4 \pi} \int_V T_{\alpha \beta \alpha \beta}(\mathbf{r})
819 d\mathbf{r}
820 \end{equation}
821 which has been integrated over the interaction volume $V$ and has
822 units of $\mathrm{length}^{-2}$.
823
824 In terms of traced quadrupole moment, equation (\ref{fourierQuad2})
825 can be written as,
826 \begin{equation}
827 \mathsf{Q} - \frac{\mathbf{I}}{3} \mathrm{Tr}(\mathsf{Q})
828 = \epsilon_o {\chi}_Q\left(1- {\chi}_Q B \right)^{-1} \nabla \mathbf{E}^\circ
829 \label{tracedConstQuad}
830 \end{equation}
831 Comparing (\ref{tracedConstQuad}) and (\ref{flucQuad}) we obtain,
832 \begin{equation}
833 \frac{\braket{\mathsf{M}_Q^2}-{\braket{\mathsf{M}_Q}}^2}{15 \epsilon_o
834 V k_B T} = \chi_Q \left(1 - \chi_Q B \right)^{-1},
835 \end{equation}
836 or equivalently,
837 \begin{equation}
838 \chi_Q = \frac{\braket{\mathsf{M}_Q^2}-{\braket{\mathsf{M}_Q}}^2}{15 \epsilon_o
839 V k_B T} \left(1 + B \frac{\braket{\mathsf{M}_Q^2}-{\braket{\mathsf{M}_Q}}^2}{15 \epsilon_o
840 V k_B T} \right)^{-1}
841 \label{eq:finalForm}
842 \end{equation}
843 Eq. (\ref{eq:finalForm}) now expresses a bulk property (the
844 quadrupolar susceptibility, $\chi_Q$) in terms of a straightforward
845 fluctuation in the system quadrupole moment and a quadrupolar
846 correction factor ($B$). The correction factors depend on the cutoff
847 method being employed in the simulation, and these are listed in Table
848 \ref{tab:B}.
849
850 In terms of the macroscopic quadrupole polarizability, $\alpha_Q$,
851 which may be thought of as the ``conducting boundary'' version of the
852 susceptibility,
853 \begin{equation}
854 \chi_Q = \frac{\alpha_Q}{1 + B \alpha_Q}.
855 \label{eq:quadrupolarSusceptiblity}
856 \end{equation}
857 If an electrostatic method produces $B \rightarrow 0$, the computed
858 quadrupole polarizability and quadrupole susceptibility converge to
859 the same value.
860
861 \begin{sidewaystable}
862 \caption{Expressions for the quadrupolar correction factor ($B$) for
863 the real-space electrostatic methods in terms of the damping
864 parameter ($\alpha$) and the cutoff radius ($r_c$). The units
865 of the correction factor are $ \mathrm{length}^{-2}$ for quadrupolar
866 fluids.}
867 \label{tab:B}
868 \begin{tabular}{l|c|c|c}
869 \toprule
870 Method & charges & dipoles & quadrupoles \\\colrule
871 Spherical Cutoff (SC) & $ -\frac{8 \alpha^5 {r_c}^3}{3\sqrt{\pi}} e^{-\alpha^2 r_c^2}$ & $ -\frac{8 \alpha^5 {r_c}^3}{3\sqrt{\pi}} e^{-\alpha^2 r_c^2}$ & $ -\frac{8 {\alpha}^5 {r_c}^3}{3\sqrt{\pi}} e^{-\alpha^2 r_c^2}$ \\
872 Shifted Potental (SP) & $ -\frac{8 \alpha^5 {r_c}^3}{3\sqrt{\pi}} e^{-\alpha^2 r_c^2}$ & $- \frac{8 \alpha^5 {r_c}^3}{3\sqrt{\pi}} e^{-\alpha^2 r_c^2}$& $ -\frac{16 \alpha^7 {r_c}^5}{9\sqrt{\pi}} e^{-\alpha^2 r_c^2}$ \\
873 Gradient-shifted (GSF) & $- \frac{8 \alpha^5 {r_c}^3}{3\sqrt{\pi}} e^{-\alpha^2 r_c^2}$ & 0 & $-\frac{4{\alpha}^7{r_c}^5 }{9\sqrt{\pi}}e^{-\alpha^2 r_c^2}(-1+2\alpha ^2 r_c^2)$\\
874 Taylor-shifted (TSF) & $ -\frac{8 \alpha^5 {r_c}^3}{3\sqrt{\pi}}
875 e^{-\alpha^2 r_c^2}$ &
876 $\frac{4\;\mathrm{erfc(\alpha
877 r_c)}}{{r_c}^2} +
878 \frac{8
879 \alpha}{3\sqrt{\pi}r_c}e^{-\alpha^2
880 {r_c}^2}\left(3+ 2
881 \alpha^2 {r_c}^2 +
882 \alpha^4 {r_c}^4\right)
883 $ &
884 $\frac{10\;\mathrm{erfc}(\alpha r_c )}{{r_c}^2} + \frac{4{\alpha}}{9\sqrt{\pi}{r_c}}e^{-\alpha^2 r_c^2}\left(45 + 30\alpha ^2 {r_c}^2 + 12\alpha^4 {r_c}^4 + 3\alpha^6 {r_c}^6 + 2 \alpha^8 {r_c}^8\right)$ \\
885 \colrule
886 Ewald-Kornfeld (EK) & \multicolumn{3}{c}{ Nothing Yet} \\
887 \botrule
888 \end{tabular}
889 \end{sidewaystable}
890
891 \section{Screening of Charges by Multipolar Fluids}
892 \label{sec:PMF}
893 In a dipolar fluid, the static dielectric constant is also a measure
894 of the ability of the fluid to screen charges from one another. A set
895 of point charges creates an inhomogeneous field in the fluid, and the
896 fluid responds to this field as if it was created externally or via
897 local polarization fluctuations. For this reason, the dielectric
898 constant can be used to estimate an effective potential between two
899 point charges ($C_i$ and $C_j$) embedded in the fluid,
900 \begin{equation}
901 U_\mathrm{effective} = \frac{C_i C_j}{4 \pi \epsilon_0 \epsilon
902 r_{ij}}.
903 \label{eq:effectivePot}
904 \end{equation}
905
906 The same set of point charges can also create an inhomogeneous field
907 \textit{gradient}, and this will cause a response in a quadrupolar
908 fluid that will also cause an effective screening. As discussed above,
909 however, the relevant phyiscal property in quadrupolar fluids is the
910 susceptibility, $\chi_Q$. The screening dielectric associated with
911 the quadrupolar susceptibility is defined as,\cite{Ernst92}
912 \begin{equation}
913 \epsilon = 1 + \chi_Q G = 1 + G \frac{\alpha_Q}{1 + \alpha_Q B}
914 \label{eq:dielectricFromQuadrupoles}
915 \end{equation}
916 where $G$ is a geometrical factor that depends on the geometry of the
917 field perturbation,
918 \begin{equation}
919 G = \frac{\int_V \left| \nabla \mathbf{E}^\circ \right|^2 d\mathbf{r}}
920 {\int_V \left|\mathbf{E}^\circ\right|^2 d\mathbf{r}}
921 \end{equation}
922 integrated over the interaction volume. Note that this geometrical
923 factor is also required to compute effective dielectric constants even
924 when the field gradient is homogeneous over the entire sample.
925
926 To measure effective screening in a multipolar fluid, we compute an
927 effective interaction potential, the potential of mean force (PMF),
928 between positively and negatively charged ions when they screened by
929 the intervening fluid. The PMF is obtained from a sequence of
930 simulations in which two ions are constrained to a fixed distance, and
931 the average constraint force to hold them at a fixed distance $r$ is
932 collected during a long simulation,\cite{Wilfred07}
933 \begin{equation}
934 w(r) = \int_{r_o}^{r}\left\langle \frac{\partial f}{\partial r^\prime}
935 \right\rangle_{r^\prime} dr^\prime + 2k_BT \ln\left(\frac{r}{r_o}\right) + w(r_o),
936 \label{eq:pmf}
937 \end{equation}
938 where $\braket{\partial f/\partial r^\prime}_{r^\prime}$ is the mean
939 constraint force required to hold the ions at distance $r^\prime$,
940 $2k_BT \log(r/r_o)$ is the Fixman factor,\cite{Fixman:1974fk} and
941 $r_o$ is a reference position (usually taken as a large separation
942 betwen the ions). If the dielectric constant is a good measure of the
943 screening at all inter-ion separations, we would expect $w(r)$ to have
944 the form in Eq. (\ref{eq:effectivePot}). Because real fluids are not
945 continuum dielectrics, the effective dielectric constant is a function
946 of the interionic separation,
947 \begin{equation}
948 \epsilon(r) = \frac{u_\mathrm{raw}(r) }{w(r)}
949 \end{equation}
950 where $u_\mathrm{raw}(r)$ is the direct charge-charge interaction
951 potential that is in use during the simulation. $\epsilon(r)$ may
952 vary considerably from the bulk estimates at short distances, although
953 it should converge to the bulk value as the separation between the
954 ions increases.
955
956 \section{Simulation Methodology}
957 To test the formalism developed in the preceding sections, we have
958 carried out computer simulations using three different techniques: i)
959 simulations in the presence of external fields, ii) equilibrium
960 calculations of box moment fluctuations, and iii) potentials of mean
961 force (PMF) between embedded ions. In all cases, the fluids were
962 composed of point multipoles protected by a Lennard-Jones potential.
963 The parameters used in the test systems are given in table
964 \ref{Tab:C}.
965
966 \begin{table}
967 \caption{\label{Tab:C}The parameters used in simulations to evaluate
968 the dielectric response of the new real-space methods.}
969 \begin{tabularx}{\textwidth}{r|cc|YYccc|Yccc} \hline
970 & \multicolumn{2}{c|}{LJ parameters} &
971 \multicolumn{5}{c|}{Electrostatic moments} & & & & \\
972 Test system & $\sigma$& $\epsilon$ & $C$ & $D$ &
973 $Q_{xx}$ & $Q_{yy}$ & $Q_{zz}$ & mass & $I_{xx}$ & $I_{yy}$ &
974 $I_{zz}$ \\ \cline{6-8}\cline{10-12}
975 & (\AA) & (kcal/mol) & (e) & (debye) & \multicolumn{3}{c|}{(debye \AA)} & (amu) & \multicolumn{3}{c}{(amu
976 \AA\textsuperscript{2})} \\ \hline
977 Dipolar fluid & 3.41 & 0.2381 & - & 1.4026 &-&-&-& 39.948 & 11.613 & 11.613 & 0.0 \\
978 Quadrupolar fluid & 2.985 & 0.265 & - & - & 0.0 & 0.0 &-2.139 & 18.0153 & 43.0565 & 43.0565 & 0.0 \\
979 \ce{q+} & 1.0 & 0.1 & +1 & - & - & - & - & 22.98 & - & - & - \\
980 \ce{q-} & 1.0 & 0.1 & -1 & - & - & - & - & 22.98 & - & - & - \\ \hline
981 \end{tabularx}
982 \end{table}
983
984 The first of the test systems consists entirely of fluids of point
985 dipolar or quadrupolar molecules in the presence of constant field or
986 field gradients. Since there are no isolated charges within the
987 system, the divergence of the field will be zero, \textit{i.e.}
988 $\vec{\nabla} \cdot \mathbf{E} = 0$. This condition can be satisfied
989 by using the relatively simple applied potential as described in the
990 supporting information.
991
992 When a constant electric field or field gradient is applied to the
993 system, the molecules align along the direction of the applied field,
994 and polarize to a degree determined both by the strength of the field
995 and the fluid's polarizibility. We have calculated ensemble averages
996 of the box dipole and quadrupole moments as a function of the strength
997 of the applied fields. If the fields are sufficiently weak, the
998 response is linear in the field strength, and one can easily compute
999 the polarizability directly from the simulations.
1000
1001 The second set of test systems consists of equilibrium simulations of
1002 fluids of point dipolar or quadrupolar molecules simulated in the
1003 absence of any external perturbation. The fluctuation of the ensemble
1004 averages of the box multipolar moment was calculated for each of the
1005 multipolar fluids. The box multipolar moments were computed as simple
1006 sums over the instantaneous molecular moments, and fluctuations in
1007 these quantities were obtained from Eqs. (\ref{eq:flucDip}) and
1008 (\ref{eq:flucQuad}). The macroscopic polarizabilities of the system at
1009 a were derived using Eqs.(\ref{flucDipole}) and (\ref{flucQuad}).
1010
1011 The final system consists of dipolar or quadrupolar fluids with two
1012 oppositely charged ions embedded within the fluid. These ions are
1013 constrained to be at fixed distance throughout a simulation, although
1014 they are allowed to move freely throughout the fluid while satisfying
1015 that constraint. Separate simulations were run at a range of
1016 constraint distances. A dielectric screening factor was computed using
1017 the ratio between the potential between the two ions in the absence of
1018 the fluid medium and the PMF obtained from the simulations.
1019
1020 We carried out these simulations for all three of the new real-space
1021 electrostatic methods (SP, GSF, and TSF) that were developed in the
1022 first paper (Ref. \onlinecite{PaperI}) in the series. The radius of
1023 the cutoff sphere was taken to be 12~\AA. Each of the real space
1024 methods also depends on an adjustable damping parameter $\alpha$ (in
1025 units of $\mathrm{length}^{-1}$). We have selected ten different
1026 values of damping parameter: 0.0, 0.05, 0.1, 0.15, 0.175, 0.2, 0.225,
1027 0.25, 0.3, and 0.35~\AA$^{-1}$ in our simulations of the dipolar
1028 liquids, while four values were chosen for the quadrupolar fluids:
1029 0.0, 0.1, 0.2, and 0.3~\AA$^{-1}$.
1030
1031 For each of the methods and systems listed above, a reference
1032 simulation was carried out using a multipolar implementation of the
1033 Ewald sum.\cite{Smith82,Smith98} A default tolerance of
1034 $1 \times 10^{-8}$~kcal/mol was used in all Ewald calculations,
1035 resulting in Ewald coefficient 0.3119~\AA$^{-1}$ for a cutoff radius
1036 of 12~\AA. All of the electrostatics and constraint methods were
1037 implemented in our group's open source molecular simulation program,
1038 OpenMD,\cite{Meineke05,openmd} which was used for all calculations in
1039 this work.
1040
1041 Dipolar systems contained 2048 Lennard-Jones-protected point dipolar
1042 (Stockmayer) molecules with reduced density $\rho^* = 0.822$,
1043 temperature $T^* = 1.15$, moment of inertia $I^* = 0.025$, and dipole
1044 moment $\mu^* = \sqrt{3.0}$. These systems were equilibrated for
1045 0.5~ns in the canonical (NVT) ensemble. Data collection was carried
1046 out over a 1~ns simulation in the microcanonical (NVE) ensemble. Box
1047 dipole moments were sampled every fs. For simulations with external
1048 perturbations, field strengths ranging from $0 - 10 \times
1049 10^{-4}$~V/\AA with increments of $ 10^{-4}$~V/\AA were carried out
1050 for each system.
1051
1052 Quadrupolar systems contained 4000 linear point quadrupoles with a
1053 density $2.338 \mathrm{~g/cm}^3$ at a temperature of 500~K. These
1054 systems were equilibrated for 200~ps in a canonical (NVT) ensemble.
1055 Data collection was carried out over a 500~ps simulation in the
1056 microcanonical (NVE) ensemble. Components of box quadrupole moments
1057 were sampled every 100 fs. For quadrupolar simulations with external
1058 field gradients, field strengths ranging from
1059 $0 - 9 \times 10^{-2}$~V/\AA$^2$ with increments of
1060 $10^{-2}$~V/\AA$^2$ were carried out for each system.
1061
1062 To carry out the PMF simulations, two of the multipolar molecules in
1063 the test system were converted into \ce{q+} and \ce{q-} ions and
1064 constrained to remain at a fixed distance for the duration of the
1065 simulation. The constrained distance was then varied from 5--12~\AA.
1066 In the PMF calculations, all simulations were equilibrated for 500 ps
1067 in the NVT ensemble and run for 5 ns in the microcanonical (NVE)
1068 ensemble. Constraint forces were sampled every 20~fs.
1069
1070 \section{Results}
1071 \subsection{Dipolar fluid}
1072 \begin{figure}
1073 \includegraphics[width=5in]{dielectricFinal_Dipole.pdf}
1074 \caption{The polarizability ($\alpha_D$), correction factor ($A$), and
1075 dielectric constant ($\epsilon$) for the test dipolar fluid. The
1076 left panels were computed using external fields, and those on the
1077 right are the result of equilibrium fluctuations. In the GSF and SP
1078 methods, the corrections are large in with small values of $\alpha$,
1079 and a optimal damping coefficient is evident around 0.25 \AA$^{-1}$.
1080 The dashed lines in the upper panel indicate back-calculation of the
1081 polarizability using the Ewald estimate (Refs. \onlinecite{Adams81}
1082 and \onlinecite{NeumannI83}) for the dielectric constant.}
1083 \label{fig:dielectricDipole}
1084 \end{figure}
1085 The macroscopic polarizability ($\alpha_D$) for the dipolar fluid is
1086 shown in the upper panels in Fig. \ref{fig:dielectricDipole}. The
1087 polarizability obtained from the both perturbation and fluctuation
1088 approaches are in excellent agreement with each other. The data also
1089 show a stong dependence on the damping parameter for both the Shifted
1090 Potential (SP) and Gradient Shifted force (GSF) methods, while Taylor
1091 shifted force (TSF) is largely independent of the damping
1092 parameter.
1093
1094 The calculated correction factors discussed in section
1095 \ref{sec:corrFactor} are shown in the middle panels. Because the TSF
1096 method has $A = 1$ for all values of the damping parameter, the
1097 computed polarizabilities need no correction for the dielectric
1098 calculation. The value of $A$ varies with the damping parameter in
1099 both the SP and GSF methods, and inclusion of the correction yields
1100 dielectric estimates (shown in the lower panel) that are generally too
1101 large until the damping reaches $\sim$~0.25~\AA$^{-1}$. Above this
1102 value, the dielectric constants are generally in good agreement with
1103 previous simulation results.\cite{NeumannI83}
1104
1105 Figure \ref{fig:dielectricDipole} also contains back-calculations of
1106 the polarizability using the reference (Ewald) simulation
1107 results.\cite{NeumannI83} These are indicated with dashed lines in the
1108 upper panels. It is clear that the expected polarizability for the SP
1109 and GSF methods are quite close to results obtained from the
1110 simulations. This indicates that the correction formula for the
1111 dipolar fluid (Eq. \ref{correctionFormula}) is quite sensitive when
1112 the value of $\mathrm{A}$ deviates significantly from unity.
1113
1114 These results also suggest an optimal value for the damping parameter
1115 of ($\alpha \sim 0.2-0.3$~\AA$^{-1}$ when evaluating dielectric
1116 constants of point dipolar fluids using either the perturbation and
1117 fluctuation approaches for the new real-space methods.
1118
1119 \begin{figure}
1120 \includegraphics[width=4in]{ScreeningFactor_Dipole.pdf}
1121 \caption{The distance-dependent screening factor, $\epsilon(r)$,
1122 between two ions immersed in the dipolar fluid. The new methods are
1123 shown in separate panels, and different values of the damping
1124 parameter ($\alpha$) are indicated with different symbols. All of
1125 the methods appear to be converging to the bulk dielectric constant
1126 ($\sim 65$) at large ion separations. CHECK PLOT}
1127 \label{fig:ScreeningFactor_Dipole}
1128 \end{figure}
1129 We have also evaluated the distance-dependent screening factor,
1130 $\epsilon(r)$, between two oppositely charged ions when they are
1131 placed in the dipolar fluid. These results were computed using
1132 Eq. \ref{eq:pmf} and are shown in Fig.
1133 \ref{fig:ScreeningFactor_Dipole}.
1134
1135 The screening factor is similar to the dielectric constant, but
1136 measures a local property of the ions in the fluid and depends on both
1137 ion-dipole and dipole-dipole interactions. These interactions depend
1138 on the distance between ions as well as the electrostatic interaction
1139 methods utilized in the simulations. The screening should converge to
1140 the dielectric constant when the field due to ions is small. This
1141 occurs when the ions are separated (or when the damping parameter is
1142 large). In Fig. \ref{fig:ScreeningFactor_Dipole} we observe that for
1143 the higher value of damping alpha \textit{i.e.}
1144 $\alpha = 0.2$~\AA$^{-1}$ and $0.3$~\AA$^{-1}$ and large separation
1145 between ions, the screening factor does indeed approach the correct
1146 dielectric constant.
1147
1148 REVISIT THESE THREE PARAGRAPHS:
1149
1150 It is also notable that the TSF method again displays smaller
1151 perturbations away from the correct dielectric screening behavior. We
1152 also observe that for TSF method yields high dielectric screening even
1153 for lower values of $\alpha$.
1154
1155 At short distances, the presence of the ions creates a strong local
1156 field that acts to align nearby dipoles nearly perfectly in opposition
1157 to the field from the ions. This has the effect of increasing the
1158 effective screening when the ions are brought close to one another.
1159
1160 This is due to the absence of corrections for dipole-dipole
1161 interactions as discussed above. However, the TSF remains quite
1162 perturbative for ion-dipole interactions, an effect that is most
1163 apparent at small ion separations.
1164
1165 \subsection{Quadrupolar fluid}
1166 \begin{figure}
1167 \includegraphics[width=4in]{polarizabilityFinal_Quad.pdf}
1168 \caption{The quadrupole polarizability ($\alpha_Q$), correction factor
1169 ($B$), and susceptibility ($\chi_Q$) for the test quadrupolar
1170 fluid. The left panels were computed using external field gradients,
1171 and those on the right are the result of equilibrium fluctuations.
1172 The GSF and SP methods allow nearly unmodified use of the
1173 ``conducting boundary'' or polarizability results in place of the
1174 bulk susceptibility.}
1175 \label{fig:dielectricQuad}
1176 \end{figure}
1177 The polarizability ($\alpha_Q$), correction factor ($B$), and
1178 susceptibility ($\chi_Q$) for the quadrupolar fluid is plotted against
1179 damping parameter Fig. \ref{fig:dielectricQuad}. In quadrupolar
1180 fluids, both the polarizability and susceptibility have units of
1181 $\mathrm{length}^2$ Although the susceptibility has dimensionality, it
1182 is the relevant measure of macroscopic quadrupolar
1183 properties.\cite{JeonI03, JeonII03}. The left panel in
1184 Fig. \ref{fig:dielectricQuad} shows results obtained from the applied
1185 field gradient simulations whereas the results from the equilibrium
1186 fluctuation formula are plotted in the right panels.
1187
1188 The susceptibility for the quadrupolar fluid is obtained from
1189 quadrupolarizability and a correction factor using
1190 Eq. (\ref{eq:quadrupolarSusceptiblity}). The susceptibilities are
1191 shown in the bottom panels of Fig. \ref{fig:dielectricQuad}. All three
1192 methods: (SP, GSF, and TSF) produce similar susceptibilities over the
1193 range of damping parameters. This shows that susceptibility derived
1194 using the quadrupolarizability and the correction factors are
1195 essentially independent of the electrostatic method utilized in the
1196 simulation.
1197
1198 \begin{figure}
1199 \includegraphics[width=5in]{ScreeningFactor_Quad.pdf}
1200 \caption{\label{fig:screenQuad} The distance-dependent screening
1201 factor, $\epsilon(r)$, between two ions immersed in the quadrupolar
1202 fluid. Results from the perturbation and fluctuation methods are
1203 shown in right and central panels. Here the susceptibility is
1204 calculated from the simulation and the geometrical factor is
1205 evaluated analytically, using the field and field-gradient produced
1206 by ions. The right hand panel shows the screening factor obtained
1207 from the PMF calculations.}
1208 \end{figure}
1209 A more difficult test of the quadrupolar susceptibility is made by
1210 comparing with direct calculation of the electrostatic screening using
1211 the potential of mean force (PMF). Since the effective dielectric
1212 constant for a quadrupolar fluid depends on the geometry of the field
1213 and field gradient, this is not a physical property of the quadrupolar
1214 fluid.
1215
1216 The geometrical factor for embedded ions changes with the ion
1217 separation distance. It is therefore reasonable to treat the
1218 dielectric constant as a distance-dependent screening factor. Since
1219 the quadrupolar molecules couple with the gradient of the field, the
1220 distribution of the quadrupoles will be inhomogeneously distributed
1221 around the point charges. Hence the distribution of quadrupolar
1222 molecules should be taken into account when computing the geometrical
1223 factors in the presence of this perturbation,
1224 \begin{eqnarray}
1225 G &=& \frac{\int_V g(\mathbf{r}) \left|\nabla \mathbf{E}^\circ
1226 \right|^2 d\mathbf{r}}{\int_V\left|\mathbf{E}^\circ\right|^2
1227 d\mathbf{r}} \nonumber \\ \nonumber \\
1228 &=& \frac{ 2 \pi \int_{-1}^{1} \int_{0}^{R} r^2 g(r,
1229 \cos\theta) \left|\nabla \mathbf{E}^\circ \right|^2 dr d\cos\theta }{\int_V\left|\mathbf{E}^\circ\right|^2 d\mathbf{r}}
1230 \label{eq:geometricalFactor}
1231 \end{eqnarray}
1232 where $g(r,\cos\theta)$ is a distribution function for the quadrupoles
1233 with respect to an origin at midpoint of a line joining the two probe
1234 charges.
1235
1236 The effective screening factor is plotted against ion separation
1237 distance in Fig. \ref{fig:screenQuad}. The screening evaluated from
1238 the perturbation and fluctuation methods are shown in right and
1239 central panels. Here the susceptibility is calculated from the
1240 simulation and the geometrical factor is evaluated analytically, using
1241 the field and field-gradient produced by ions. The right hand panel
1242 shows the screening factor obtained from the PMF calculations.
1243
1244 We note that the screening factor obtained from both the perturbation
1245 and fluctuation formula show good agreement with the screening factor
1246 calculated using PMF method. As there are no large differences in
1247 quadrupole-quadruople interactions for various real-space
1248 methods,\cite{PaperI,PaperII} we generally good agreement for the
1249 screening factors using any of the real space methods.
1250
1251 \section{Conclusions}
1252 We have used both perturbation and fluctuation approaches to evaluate
1253 dielectric properties for dipolar and quadrupolar fluids. The static
1254 dielectric constant is the relevant bulk property for dipolar fluids,
1255 while the quadrupolar susceptibility plays a similar role for
1256 quadrupoles. Corrections to both the static dielectric constant and
1257 the quadrupolar susceptibility were derived for three new real space
1258 electrostatic methods, and these corrections were tested against a
1259 third measure of dielectric screening, the potential of mean force
1260 between two ions immersed in the fluids.
1261
1262 For the dipolar fluids, we find that the polarizability evaluated
1263 using the perturbation and fluctuation methods show excellent
1264 agreement, indicating that equilibrium calculations of the dipole
1265 fluctuations are good measures of bulk polarizability. One of the
1266 findings of the second paper in this series is that the moderately
1267 damped GSF and SP methods were most suitable for molecular dynamics
1268 and Monte Carlo simulations, respectively.\cite{PaperII}
1269
1270 The dielectric constant evaluated using the computed polarizability
1271 and correction factors agrees well with the previous Ewald-based
1272 simulation results \cite{Adams81,NeumannI83} for moderate damping
1273 parameters in the range 0.25--0.3~\AA~$^{-1}$.
1274
1275 Although the TSF method alters many dynamic and structural features in
1276 multipolar liquids,\cite{PaperII} it is surprisingly good at computing
1277 bulk dielectric properties at nearly all ranges of the damping
1278 parameter. In fact, the correction factor, $A = 1$, for the TSF
1279 method so the conducting boundary formula is essentially correct when
1280 using this method for point dipolar fluids.
1281
1282 The dielectric correction formula (equation ~\ref{correctionFormula})
1283 is extremely sensitive when the correction factor ($A$) deviates from
1284 1, and this behavior is also seen in the effective screening of ions
1285 embedded in the fluid.
1286
1287 As in the dipolar case, the quadpole polarizability evaluated from
1288 both perturbation and fluctuation simulations show excellent
1289 agreement, again confirming that equilibrium fluctuation calculations
1290 are sufficient to reproduce bulk dielectric properties in these
1291 fluids. The quadrupolar susceptibility calculated via our derived
1292 correction factors tends to produce the same result for all three real
1293 space methods. Similarly, the screening factor calculated using the
1294 susceptibility and a weighted geometric factor provides good agreement
1295 with results obtained directly via potentials of mean force. For
1296 quadrupolar fluids, the distance dependence of the electrostatic
1297 interaction is significantly reduced and the correction factors are
1298 all small. These points suggest that how an electrostatic method
1299 treats the cutoff radius become less consequential for higher order
1300 multipoles.
1301
1302 For this reason, we renew our recommendation that the
1303 moderately-damped GSF method is an excellent choice for molecular
1304 dynamics simulation where point-multipole interactions are being
1305 utilized to compute bulk properties of fluids.
1306
1307 \newpage
1308
1309 \bibliography{dielectric_new}
1310
1311 \end{document}