ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/multipole/dielectric_new.tex
Revision: 4420
Committed: Tue Apr 12 20:55:42 2016 UTC (8 years, 2 months ago) by gezelter
Content type: application/x-tex
File size: 77489 byte(s)
Log Message:
Added one new reference

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,Stenqvist:2015ph,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, Stern03, SalvchovI14, SalvchovII14} In
110 simulations, the net polarization of the system is also determined by
111 the interactions \textit{between} the molecules. Therefore the
112 macroscopic polarizablity obtained from a simulation depends on the
113 details of the electrostatic interaction methods that were employed in
114 the simulation. To determine the relevant physical properties of the
115 multipolar fluid from the system fluctuations, the interactions
116 between molecules must be incorporated into the formalism for the bulk
117 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 (see Fig. \ref{fig:schematic}):
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), or internal fields
174 and gradients generated by the molecules themselves (center), or
175 fields produced by embedded ions (right). The dielectric constant
176 ($\epsilon$) measures all three responses in dipolar fluids (top).
177 In quadrupolar liquids (bottom), the relevant bulk property is the
178 quadrupolar susceptibility ($\chi_Q$), and the geometry of the field
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 In standard electrostatics, the two radial functions, $v_{21}(r)$ and
239 $v_{22}(r)$ are proportional to $1/r^3$, but they have distinct radial
240 dependence in the TSF method. Careful choice of these functions makes
241 the forces and torques vanish smoothly as the molecules drift beyond
242 the cutoff radius (even when those molecules are in different
243 orientations).
244
245 A second and somewhat simpler approach involves shifting the gradient
246 of the Coulomb potential for each particular multipole order,
247 \begin{equation}
248 U^{\text{GSF}}_{ab} = \sum \left[ U(\mathbf{r}, \mathsf{A}, \mathsf{B}) -
249 U(r_c \hat{\mathbf{r}},\mathsf{A}, \mathsf{B}) - (r-r_c)
250 \hat{\mathbf{r}} \cdot \nabla U(r_c \hat{\mathbf{r}},\mathsf{A},
251 \mathsf{B}) \right]
252 \label{generic2}
253 \end{equation}
254 where the sum describes a separate force-shifting that is applied to
255 each orientational contribution to the energy, i.e. $v_{21}$ and
256 $v_{22}$ are shifted separately. In this expression,
257 $\hat{\mathbf{r}}$ is the unit vector connecting the two multipoles
258 ($a$ and $b$) in space, and $\mathsf{A}$ and $\mathsf{B}$ represent
259 the orientations the multipoles. Because this procedure is equivalent
260 to using the gradient of an image multipole placed at the cutoff
261 sphere for shifting the force, this method is called the gradient
262 shifted-force (GSF) approach.
263
264 Both the TSF and GSF approaches can be thought of as multipolar
265 extensions of the original damped shifted-force (DSF) approach that
266 was developed for point charges. There is also a multipolar extension
267 of the Wolf sum that is obtained by projecting an image multipole onto
268 the surface of the cutoff sphere, and including the interactions with
269 the central multipole and the image. This effectively shifts only the
270 total potential to zero at the cutoff radius,
271 \begin{equation}
272 U^{\text{SP}}_{ab} = \sum \left[ U(\mathbf{r}, \mathsf{A}, \mathsf{B}) -
273 U(r_c \hat{\mathbf{r}},\mathsf{A}, \mathsf{B}) \right]
274 \label{eq:SP}
275 \end{equation}
276 where the sum again describes separate potential shifting that is done
277 for each orientational contribution to the energy. The potential
278 energy between a central multipole and other multipolar sites goes
279 smoothly to zero as $r \rightarrow r_c$, but the forces and torques
280 obtained from this shifted potential (SP) approach are discontinuous
281 at $r_c$.
282
283 All three of the new real space methods share a common structure: the
284 various orientational contributions to multipolar interaction energies
285 require separate treatment of their radial functions, and these are
286 tabulated for both the raw Coulombic kernel ($1/r$) as well as the
287 damped kernel ($\mathrm{erfc}(\alpha r)/r$), in the first paper of this
288 series.\cite{PaperI} The second paper in this series evaluated the
289 fidelity with which the three new methods reproduced Ewald-based
290 results for a number of model systems.\cite{PaperII} One of the major
291 findings was that moderately-damped GSF simulations produced nearly
292 identical behavior with Ewald-based simulations, but the real-space
293 methods scale linearly with system size.
294
295 \section{Dipolar Fluids and the Dielectric Constant}
296
297 Dielectric properties of a fluid arise mainly from responses of the
298 fluid to either applied fields or transient fields internal to the
299 fluid. In response to an applied field, the molecules have electronic
300 polarizabilities, changes to internal bond lengths, and reorientations
301 towards the direction of the applied field. There is an added
302 complication that in the presence of external field, the perturbation
303 experienced by any single molecule is not only due to external field
304 but also to the fields produced by the all other molecules in the
305 system.
306
307 \subsection{Response to External Perturbations}
308
309 In the presence of uniform electric field $\mathbf{E}$, an individual
310 molecule with a permanent dipole moment $p_o$ will realign along the
311 direction of the field with an average polarization given by
312 \begin{equation}
313 \braket{\mathbf{p}} = \epsilon_0 \alpha_p \mathbf{E},
314 \end{equation}
315 where $\alpha_p = {p_o}^2 / 3 \epsilon_0 k_B T$ is the contribution to
316 molecular polarizability due solely to reorientation dynamics.
317 Because the applied field must overcome thermal motion, the
318 orientational polarization depends inversely on the temperature.
319
320 Likewise, a condensed phase system of permanent dipoles will also
321 polarize along the direction of an applied field. The polarization
322 density of the system is
323 \begin{equation}
324 \textbf{P} = \epsilon_o \alpha_{D} \mathbf{E}.
325 \label{pertDipole}
326 \end{equation}
327 The constant $\alpha_D$ is the macroscopic polarizability, which is an
328 emergent property of the dipolar fluid. Note that the polarizability,
329 $\alpha_D$ is distinct from the dipole susceptibility, $\chi_D$,
330 which is the quantity most directly related to the static dielectric
331 constant, $\epsilon = 1 + \chi_D$.
332
333 \subsection{Fluctuation Formula}
334 % Using the Ewald sum under tin-foil boundary conditions, $\epsilon$ can
335 % be calculated for systems of orientationally-polarized molecules,
336 % \begin{equation}
337 % \epsilon = 1 + \frac{\langle M^2\rangle}{3\epsilon_0k_\textrm{B}TV},
338 % \label{eq:staticDielectric}
339 % \end{equation}
340 % where $\epsilon_0$ is the permittivity of free space and
341 % $\langle M^2\rangle$ is the fluctuation of the system dipole
342 % moment.\cite{Allen89} The numerator in the fractional term in
343 % equation (\ref{eq:staticDielectric}) is identical to the quantity
344 % calculated in the finite-system Kirkwood $g$ factor ($G_k$):
345 % \begin{equation}
346 % G_k = \frac{\langle M^2\rangle}{N\mu^2},
347 % \label{eq:KirkwoodFactor}
348 % \end{equation}
349 % where $\mu$ is the dipole moment of a single molecule of the
350 % homogeneous system.\cite{NeumannI83,NeumannII83,Neumann84,Neumann85} The
351 % fluctuation term in both equation (\ref{eq:staticDielectric}) and
352 % (\ref{eq:KirkwoodFactor}) is calculated as follows,
353 % \begin{equation}
354 % \begin{split}
355 % \langle M^2\rangle &= \langle\bm{M}\cdot\bm{M}\rangle
356 % - \langle\bm{M}\rangle\cdot\langle\bm{M}\rangle \\
357 % &= \langle M_x^2+M_y^2+M_z^2\rangle
358 % - (\langle M_x\rangle^2 + \langle M_x\rangle^2
359 % + \langle M_x\rangle^2).
360 % \end{split}
361 % \label{eq:fluctBoxDipole}
362 % \end{equation}
363 % This fluctuation term can be accumulated during a simulation; however,
364 % it converges rather slowly, thus requiring multi-nanosecond simulation
365 % times.\cite{Horn04} In the case of tin-foil boundary conditions, the
366 % dielectric/surface term of the Ewald summation is equal to zero.
367
368 For a system of dipolar molecules at thermal equilibrium, we can
369 define both a system dipole moment, $\mathbf{M} = \sum_i \mathbf{p}_i$
370 as well as a dipole polarization density,
371 $\mathbf{P} = \braket{\mathbf{M}}/V$.
372
373 In the presence of applied field $\mathbf{E}$, the polarization
374 density can be expressed in terms of fluctuations in the net dipole
375 moment,
376 \begin{equation}
377 \mathbf{P} = \epsilon_o \frac{\braket{\mathbf{M}^2}-{\braket{\mathbf{M}}}^2}{3 \epsilon_o V k_B T}\bf{E}
378 \label{flucDipole}
379 \end{equation}
380 This has structural similarity with the Boltzmann average for the
381 polarization of a single molecule. Here
382 $ \braket{\mathbf{M}^2}-{\braket{\mathbf{M}}}^2$ measures fluctuations
383 in the net dipole moment,
384 \begin{equation}
385 \langle \mathbf{M}^2 \rangle - \langle \mathbf{M} \rangle^2 =
386 \langle M_x^2+M_y^2+M_z^2 \rangle - \left( \langle M_x \rangle^2 +
387 \langle M_y \rangle^2 +
388 \langle M_z \rangle^2 \right).
389 \label{eq:flucDip}
390 \end{equation}
391 For the limiting case $\textbf{E} \rightarrow 0 $, the ensemble
392 average of both the net dipole moment $\braket{\mathbf{M}}$ and
393 dipolar polarization $\bf{P}$ tends to vanish but
394 $\braket{\mathbf{M}^2}$ does not. The macroscopic dipole
395 polarizability can therefore be written as,
396 \begin{equation}
397 \alpha_D = \frac{\braket{\mathbf{M}^2}-{\braket{\mathbf{M}}}^2}{3 \epsilon_o V k_B T}
398 \end{equation}
399 This relationship between a macroscopic property of dipolar material
400 and microscopic fluctuations is true even when the applied field
401 $ \textbf{E} \rightarrow 0 $.
402
403 \subsection{Correction Factors}
404 \label{sec:corrFactor}
405 In the presence of a uniform external field $ \mathbf{E}^\circ$, the
406 total electric field at $\mathbf{r}$ depends on the polarization
407 density at all other points in the system,\cite{NeumannI83}
408 \begin{equation}
409 \mathbf{E}(\mathbf{r}) = \mathbf{E}^\circ(\mathbf{r}) +
410 \frac{1}{4\pi\epsilon_o} \int d\mathbf{r}^\prime
411 \mathbf{T}(\mathbf{r}-\mathbf{r}^\prime)\cdot
412 {\mathbf{P}(\mathbf{r}^\prime)}.
413 \label{eq:localField}
414 \end{equation}
415 $\mathbf{T}$ is the dipole interaction tensor connecting dipoles at
416 $\mathbf{r}^\prime$ with the point of interest ($\mathbf{r}$).
417
418 In simulations of dipolar fluids, the molecular dipoles may be
419 represented either by closely-spaced point charges or by
420 point dipoles (see Fig. \ref{fig:tensor}).
421 \begin{figure}
422 \includegraphics[width=\linewidth]{Tensors}
423 \caption{In the real-space electrostatic methods, the molecular dipole
424 tensor, $\mathbf{T}_{\alpha\beta}(r)$, is not the same for
425 charge-charge interactions as for point dipoles (left panel). The
426 same holds true for the molecular quadrupole tensor (right panel),
427 $\mathbf{T}_{\alpha\beta\gamma\delta}(r)$, which can have distinct
428 forms if the molecule is represented by charges, dipoles, or point
429 quadrupoles.}
430 \label{fig:tensor}
431 \end{figure}
432 In the case where point charges are interacting via an electrostatic
433 kernel, $v(r)$, the effective {\it molecular} dipole tensor,
434 $\mathbf{T}$ is obtained from two successive applications of the
435 gradient operator to the electrostatic kernel,
436 \begin{eqnarray}
437 \mathbf{T}_{\alpha \beta}(r) &=& \nabla_\alpha \nabla_\beta
438 \left(v(r)\right) \\
439 &=& \delta_{\alpha \beta}
440 \left(\frac{1}{r} v^\prime(r) \right) + \frac{r_{\alpha}
441 r_{\beta}}{r^2} \left( v^{\prime \prime}(r) - \frac{1}{r}
442 v^{\prime}(r) \right)
443 \label{dipole-chargeTensor}
444 \end{eqnarray}
445 where $v(r)$ may be either the bare kernel ($1/r$) or one of the
446 modified (Wolf or DSF) kernels. This tensor describes the effective
447 interaction between molecular dipoles ($\mathbf{D}$) in Gaussian units
448 as $-\mathbf{D} \cdot \mathbf{T} \cdot \mathbf{D}$.
449
450 When utilizing any of the three new real-space methods for point
451 \textit{dipoles}, the tensor is explicitly constructed,
452 \begin{equation}
453 \mathbf{T}_{\alpha \beta}(r) = \delta_{\alpha \beta} v_{21}(r) +
454 \frac{r_{\alpha} r_{\beta}}{r^2} v_{22}(r)
455 \label{dipole-diopleTensor}
456 \end{equation}
457 where the functions $v_{21}(r)$ and $v_{22}(r)$ depend on the level of
458 the approximation.\cite{PaperI,PaperII} Although the Taylor-shifted
459 (TSF) and gradient-shifted (GSF) models produce to the same $v(r)$
460 function for point charges, they have distinct forms for the
461 dipole-dipole interaction.
462
463 Using a constitutive relation, the polarization density
464 $\mathbf{P}(\mathbf{r})$ is given by,
465 \begin{equation}
466 \mathbf{P}(\mathbf{r}) = \epsilon_o \chi_D
467 \left(\mathbf{E}^\circ(\mathbf{r}) + \frac{1}{4\pi\epsilon_o} \int
468 d\mathbf{r}^\prime \mathbf{T}(\mathbf{r}-\mathbf{r}^\prime ) \cdot \mathbf{P}(\mathbf{r}^\prime)\right).
469 \label{constDipole}
470 \end{equation}
471 Note that $\chi_D$ explicitly depends on the details of the dipole
472 interaction tensor. Neumann \textit{et al.}
473 \cite{NeumannI83,NeumannII83,Neumann84,Neumann85} derived an elegant
474 way to modify the fluctuation formula to correct for approximate
475 interaction tensors. This correction was derived using a Fourier
476 representation of the interaction tensor,
477 $\tilde{\mathbf{T}}(\mathbf{k})$, and involves the quantity,
478 \begin{equation}
479 A = \frac{3}{4\pi}\tilde{\mathbf{T}}(0) = \frac{3}{4\pi} \int_V
480 d\mathbf{r} \mathbf{T}(\mathbf{r})
481 \end{equation}
482 which is the $k \rightarrow 0$ limit of $\tilde{\mathbf{T}}$. Using
483 this quantity (originally called $Q$ in
484 refs. \onlinecite{NeumannI83,NeumannII83,Neumann84,Neumann85}), the
485 dielectric constant can be computed
486 \begin{equation}
487 \epsilon = \frac{3+(A + 2)(\epsilon_{CB}-1)}{3 + (A -1)(\epsilon_{CB}-1)}
488 \label{correctionFormula}
489 \end{equation}
490 where $\epsilon_{CB}$ is the widely-used conducting boundary
491 expression for the dielectric constant,
492 \begin{equation}
493 \epsilon_{CB} = 1 + \frac{\braket{\bf{M}^2}-{\braket{\bf{M}}}^2}{3
494 \epsilon_o V k_B T} = 1 + \alpha_{D}
495 \label{conductingBoundary}
496 \end{equation}
497 Eqs. (\ref{correctionFormula}) and (\ref{conductingBoundary}) allows
498 estimation of the static dielectric constant from fluctuations easily
499 computed directly from simulations.
500
501 % 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}
502 % \begin{equation}
503 % \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}
504 % \label{singular}
505 % \end{equation}
506 % Substituting equation (\ref{singular}) in the equation (\ref{constDipole}) we get,
507 % \begin{equation}
508 % \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).
509 % \end{equation}
510 % For both polarization and electric field homogeneous, this can be easily solved using Fourier transformation,
511 % \begin{equation}
512 % \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}).
513 % \end{equation}
514 % For homogeneous applied field Fourier component is non-zero only if $\kappa = 0$. Therefore,
515 % \begin{equation}
516 % \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}).
517 % \label{fourierDipole}
518 % \end{equation}
519 % 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,
520 % \begin{equation}
521 % \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}
522 % \end{equation}
523 % 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,
524 % \begin{equation}
525 % \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}
526 % \label{correctionFormula}
527 % \end{equation}
528
529 We have utilized the Neumann \textit{et al.} approach for the three
530 new real-space methods, and obtain method-dependent correction
531 factors. The expression for the correction factor also depends on
532 whether the simulation involves point charges or point dipoles to
533 represent the molecular dipoles. These corrections factors are
534 listed in Table \ref{tab:A}.
535 \begin{table}
536 \caption{Expressions for the dipolar correction factor ($A$) for the
537 real-space electrostatic methods in terms of the damping parameter
538 ($\alpha$) and the cutoff radius ($r_c$). The Ewald-Kornfeld result
539 derived in Refs. \onlinecite{Adams80,Adams81,NeumannI83} is shown
540 for comparison using the Ewald convergence parameter ($\kappa$)
541 and the real-space cutoff value ($r_c$). }
542 \label{tab:A}
543 \begin{tabular}{l|c|c}
544 \toprule
545 & \multicolumn{2}{c}{Molecular Representation} \\
546 Method & point charges & point dipoles \\
547 \colrule
548 Shifted Potental (SP) & $ \mathrm{erf}(r_c \alpha) - \frac{2
549 \alpha r_c}{\sqrt{\pi}} e^{-\alpha^2
550 r_c^2}$ & $\mathrm{erf}(r_c \alpha)
551 -\frac{2 \alpha
552 r_c}{\sqrt{\pi}}\left(1+\frac{2\alpha^2
553 {r_c}^2}{3}
554 \right)e^{-\alpha^2{r_c}^2}
555 $\\ \colrule
556 Gradient-shifted (GSF) & 1 & $\mathrm{erf}(\alpha r_c)-\frac{2
557 \alpha r_c}{\sqrt{\pi}} \left(1 +
558 \frac{2 \alpha^2 r_c^2}{3} +
559 \frac{\alpha^4
560 r_c^4}{3}\right)e^{-\alpha^2 r_c^2} $ \\ \colrule
561 Taylor-shifted (TSF) &\multicolumn{2}{c}{1} \\ \colrule
562 Spherical Cutoff (SC)& \multicolumn{2}{c}{$\mathrm{erf}(r_c \alpha) -
563 \frac{2 \alpha r_c}{\sqrt{\pi}} e^{-\alpha^2
564 r_c^2}$} \\
565 Ewald-Kornfeld (EK) & \multicolumn{2}{c}{$\mathrm{erf}(r_c \kappa) -
566 \frac{2 \kappa r_c}{\sqrt{\pi}} e^{-\kappa^2
567 r_c^2}$} \\
568 \botrule
569 \end{tabular}
570 \end{table}
571 Note that for point charges, the GSF and TSF methods produce estimates
572 of the dielectric that need no correction, and the TSF method likewise
573 needs no correction for point dipoles.
574
575 \section{Quadrupolar Fluids and the Quadrupolar Susceptibility}
576 \subsection{Response to External Perturbations}
577
578 A molecule with a permanent quadrupole, $\mathsf{q}$, will align in
579 the presence of an electric field gradient $\nabla\mathbf{E}$. The
580 anisotropic polarization of the quadrupole is given
581 by,\cite{AduGyamfi78,AduGyamfi81}
582 \begin{equation}
583 \braket{\mathsf{q}} - \frac{\mathbf{I}}{3}
584 \mathrm{Tr}(\mathsf{q}) = \epsilon_o \alpha_q \nabla\mathbf{E},
585 \end{equation}
586 where $\alpha_q = q_o^2 / 15 \epsilon_o k_B T $ is a molecular quadrupole
587 polarizablity and $q_o$ is an effective quadrupole moment for the molecule,
588 \begin{equation}
589 q_o^2 = 3 \mathsf{q}:\mathsf{q} - \mathrm{Tr}(\mathsf{q})^2.
590 \end{equation}
591
592 In the presence of an external field gradient, a system of quadrupolar
593 molecules also organizes with an anisotropic polarization,
594 \begin{equation}
595 \mathsf{Q} - \frac{\mathbf{I}}{3} \mathrm{Tr}(\mathsf{Q}) = \epsilon_o
596 \alpha_Q \nabla\mathbf{E}
597 \end{equation}
598 where $\mathsf{Q}$ is the traced quadrupole density of the system and
599 $\alpha_Q$ is a macroscopic quadrupole polarizability which has
600 dimensions of $\mathrm{length}^{-2}$. Equivalently, the traceless form
601 may be used,
602 \begin{equation}
603 \mathsf{\Theta} = 3 \epsilon_o \alpha_Q \nabla\mathbf{E},
604 \label{pertQuad}
605 \end{equation}
606 where
607 $\mathsf{\Theta} = 3\mathsf{Q} - \mathbf{I} \mathrm{Tr}(\mathsf{Q})$
608 is the traceless tensor that also describes the system quadrupole
609 density. It is this tensor that will be utilized to derive correction
610 factors below.
611
612 \subsection{Fluctuation Formula}
613 As in the dipolar case, we may define a system quadrupole moment,
614 $\mathsf{M}_Q = \sum_i \mathsf{q}_i$ and the traced quadrupolar
615 density, $\mathsf{Q} = \mathsf{M}_Q / V$. A fluctuation formula can
616 be written for a system comprising quadrupolar
617 molecules,\cite{LoganI81,LoganII82,LoganIII82}
618 \begin{equation}
619 \mathsf{Q} - \frac{\mathbf{I}}{3} \mathrm{Tr}(\mathsf{Q}) = \epsilon_o
620 \frac{\braket{\mathsf{M}_Q^2}-{\braket{\mathsf{M}_Q}}^2}{15 \epsilon_o
621 V k_B T} \nabla\mathbf{E}.
622 \label{flucQuad}
623 \end{equation}
624 Some care is needed in the definitions of the averaged quantities. These
625 refer to the effective quadrupole moment of the system, and they are
626 computed as follows,
627 \begin{align}
628 \braket{\mathsf{M}_Q^2} &= \braket{3 \mathsf{M}_Q:\mathsf{M}_Q -
629 \mathrm{Tr}(\mathsf{M}_Q)^2 }\\
630 \braket{\mathsf{M}_Q}^2 &= 3 \braket{\mathsf{M}_Q}:\braket{\mathsf{M}_Q} -
631 \mathrm{Tr}(\braket{\mathsf{M}_Q})^2
632 \label{eq:flucQuad}
633 \end{align}
634 The macroscopic quadrupolarizability is given by,
635 \begin{equation}
636 \alpha_Q = \frac{\braket{\mathsf{M}_Q^2}-{\braket{\mathsf{M}_Q}}^2}{15 \epsilon_o V k_B T}
637 \label{propConstQuad}
638 \end{equation}
639 This relationship between a macroscopic property of a quadrupolar
640 fluid and microscopic fluctuations should be true even when the
641 applied field gradient $\nabla \mathbf{E} \rightarrow 0$.
642
643 \subsection{Correction Factors}
644 In this section we generalize the treatment of Neumann \textit{et al.}
645 for quadrupolar fluids. Interactions involving multiple quadrupoles
646 are rank 4 tensors, and we therefore describe quantities in this
647 section using Einstein notation.
648
649 In the presence of a uniform external field gradient,
650 $\partial_\alpha {E}^\circ_\beta $, the total field gradient at
651 $\mathbf{r}$ depends on the quadrupole polarization density at all
652 other points in the system,
653 \begin{equation}
654 \partial_\alpha E_\beta(\mathbf{r}) = \partial_\alpha
655 {E}^\circ_\beta(\mathbf{r}) + \frac{1}{8\pi \epsilon_o}\int
656 T_{\alpha\beta\gamma\delta}({\mathbf{r}-\mathbf{r}^\prime})
657 Q_{\gamma\delta}(\mathbf{r}^\prime) d\mathbf{r}^\prime
658 \label{gradMaxwell}
659 \end{equation}
660 where and $T_{\alpha\beta\gamma\delta}$ is the quadrupole interaction
661 tensor connecting quadrupoles at $\mathbf{r}^\prime$ with the point of
662 interest ($\mathbf{r}$).
663
664 % \begin{figure}
665 % \includegraphics[width=3in]{QuadrupoleFigure}
666 % \caption{With the real-space electrostatic methods, the molecular
667 % quadrupole tensor, $\mathbf{T}_{\alpha\beta\gamma\delta}(r)$,
668 % governing interactions between molecules is not the same for
669 % quadrupoles represented via sets of charges, point dipoles, or by
670 % single point quadrupoles.}
671 % \label{fig:quadrupolarFluid}
672 % \end{figure}
673
674 In simulations of quadrupolar fluids, the molecular quadrupoles may be
675 represented by closely-spaced point charges, by multiple point
676 dipoles, or by a single point quadrupole (see
677 Fig. \ref{fig:tensor}). In the case where point charges are
678 interacting via an electrostatic kernel, $v(r)$, the effective
679 molecular quadrupole tensor can obtained from four successive
680 applications of the gradient operator to the electrostatic kernel,
681 \begin{eqnarray}
682 T_{\alpha\beta\gamma\delta}(\mathbf{r}) &=& \nabla_\alpha \nabla_\beta
683 \nabla_\gamma \nabla_\delta v(r) \\
684 &=& \left(\delta_{\alpha\beta}\delta_{\gamma\delta} +
685 \delta_{\alpha\gamma}\delta_{\beta\delta}+
686 \delta_{\alpha\delta}\delta_{\beta\gamma}\right)\left(-\frac{v^\prime(r)}{r^3}
687 + \frac{v^{\prime \prime}(r)}{r^2}\right) \nonumber \\
688 & &+ \left(\delta_{\alpha\beta} r_\gamma r_\delta + 5 \mathrm{~permutations}
689 \right) \left(\frac{3v^\prime(r)}{r^5}-\frac{3v^{\prime \prime}(r)}{r^4} +
690 \frac{v^{\prime \prime \prime}(r)}{r^3}\right) \nonumber \\
691 & &+ r_\alpha r_\beta r_\gamma r_\delta
692 \left(-\frac{15v^\prime(r)}{r^7}+\frac{15v^{\prime \prime}(r)}{r^6}-\frac{6v^{\prime
693 \prime \prime}(r)}{r^5} + \frac{v^{\prime \prime \prime \prime}(r)}{r^4}\right),
694 \label{quadCharge}
695 \end{eqnarray}
696 where $v(r)$ can either be the electrostatic kernel ($1/r$) or one of
697 the modified (Wolf or DSF) kernels.
698
699 Similarly, when representing quadrupolar molecules with multiple point
700 \textit{dipoles}, the molecular quadrupole interaction tensor can be
701 obtained using two successive applications of the gradient operator to
702 the dipole interaction tensor,
703 \begin{eqnarray}
704 T_{\alpha\beta\gamma\delta}(\mathbf{r}) &=& \nabla_\alpha \nabla_\beta
705 T_{\gamma\delta}(\mathbf{r}) \\
706 & = & \delta_{\alpha\beta}\delta_{\gamma\delta} \frac{v^\prime_{21}(r)}{r} +
707 \left(\delta_{\alpha\gamma}\delta_{\beta\delta}+
708 \delta_{\alpha\delta}\delta_{\beta\gamma}\right)\frac{v_{22}(r)}{r^2}
709 \nonumber\\
710 & &+ \delta_{\gamma\delta} r_\alpha r_\beta
711 \left(\frac{v^{\prime \prime}_{21}(r)}{r^2}-\frac{v^\prime_{21}(r)}{r^3} \right)\nonumber \\
712 & &+\left(\delta_{\alpha\beta} r_\gamma r_\delta +
713 \delta_{\alpha\gamma} r_\beta r_\delta +\delta_{\alpha\delta}
714 r_\gamma r_\beta + \delta_{\beta\gamma} r_\alpha r_\delta
715 +\delta_{\beta\delta} r_\alpha r_\gamma \right)
716 \left(\frac{v^\prime_{22}(r)}{r^3}-\frac{2v_{22}(r)}{r^4}\right)
717 \nonumber \\
718 & &+ r_\alpha r_\beta r_\gamma r_\delta
719 \left(\frac{v^{\prime
720 \prime}_{22}(r)}{r^4}-\frac{5v^\prime_{22}(r)}{r^5}+\frac{8v_{22}(r)}{r^6}\right),
721 \label{quadDip}
722 \end{eqnarray}
723 where $T_{\gamma\delta}(\mathbf{r})$ is a dipole-dipole interaction
724 tensor that depends on the level of the
725 approximation.\cite{PaperI,PaperII} Similarly $v_{21}(r)$ and
726 $v_{22}(r)$ are the radial function for different real space cutoff
727 methods defined in the first paper in this series.\cite{PaperI}
728
729 For quadrupolar liquids modeled using point quadrupoles, the
730 interaction tensor can be constructed as,
731 \begin{eqnarray}
732 T_{\alpha\beta\gamma\delta}(\mathbf{r}) &=&
733 \left(\delta_{\alpha\beta}\delta_{\gamma\delta}
734 +
735 \delta_{\alpha\gamma}\delta_{\beta\delta}+
736 \delta_{\alpha\delta}\delta_{\beta\gamma}\right)v_{41}(r)
737 + (\delta_{\gamma\delta} r_\alpha r_\beta + \mathrm{ 5\; permutations}) \frac{v_{42}(r)}{r^2} \nonumber \\
738 & & + r_\alpha r_\beta r_\gamma r_\delta \left(\frac{v_{43}(r)}{r^4}\right),
739 \label{quadRadial}
740 \end{eqnarray}
741 where again $v_{41}(r)$, $v_{42}(r)$, and $v_{43}(r)$ are radial
742 functions defined in Paper I of the series. \cite{PaperI} Note that
743 these radial functions have different functional forms depending on
744 the level of approximation being employed.
745
746 The integral in equation (\ref{gradMaxwell}) can be divided into two
747 parts, $|\mathbf{r}-\mathbf{r}^\prime|\rightarrow 0 $ and
748 $|\mathbf{r}-\mathbf{r}^\prime|> 0$. Since the self-contribution to
749 the field gradient vanishes at the singularity (see the supporting
750 information), equation (\ref{gradMaxwell}) can be written as,
751 \begin{equation}
752 \partial_\alpha E_\beta(\mathbf{r}) = \partial_\alpha {E}^\circ_\beta(\mathbf{r}) +
753 \frac{1}{8\pi \epsilon_o}\int\limits_{|\mathbf{r}-\mathbf{r}^\prime|> 0 }
754 T_{\alpha\beta\gamma\delta}(\mathbf{r}-\mathbf{r}^\prime)
755 {Q}_{\gamma\delta}(\mathbf{r}^\prime) d\mathbf{r}^\prime.
756 \end{equation}
757 If $\mathbf{r} = \mathbf{r}^\prime$ is excluded from the integration,
758 the total gradient can be most easily expressed in terms of
759 traceless quadrupole density as below,\cite{LoganI81}
760 \begin{equation}
761 \partial_\alpha E_\beta(\mathbf{r}) = \partial_\alpha
762 {E}^\circ_\beta(\mathbf{r}) + \frac{1}{24\pi
763 \epsilon_o}\int\limits_{|\mathbf{r}-\mathbf{r}^\prime|> 0 }
764 T_{\alpha\beta\gamma\delta}(\mathbf{r}-\mathbf{r}^\prime) \Theta_{\gamma\delta}(\mathbf{r}') d\mathbf{r}^\prime,
765 \end{equation}
766 where
767 $\Theta_{\alpha\beta} = 3Q_{\alpha\beta} - \delta_{\alpha\beta}Tr(Q)$
768 is the traceless quadrupole density. In analogy to
769 Eq. (\ref{pertQuad}) above, the quadrupole polarization density may
770 now be related to the quadrupolar susceptibility, $\chi_Q$,
771 \begin{equation}
772 \frac{1}{3}{\Theta}_{\alpha\beta}(\mathbf{r}) = \epsilon_o {\chi}_Q
773 \left[\partial_\alpha {E}^\circ_\beta(\mathbf{r}) + \frac{1}{24\pi
774 \epsilon_o}\int\limits_{|\mathbf{r}-\mathbf{r}^\prime|> 0 }
775 T_{\alpha\beta\gamma\delta}(\mathbf{r}-\mathbf{r}^\prime)
776 \Theta_{\gamma\delta}(\mathbf{r}^\prime) d\mathbf{r}^\prime \right].
777 \end{equation}
778 For periodic boundaries and with a uniform imposed
779 $\partial_\alpha E^\circ_\beta$, the quadrupole density
780 ${\Theta}_{\alpha\beta}$ will be uniform over the entire space. After
781 performing a Fourier transform (see the Appendix in
782 ref. \onlinecite{NeumannI83}) we obtain,
783 \begin{equation}
784 \frac{1}{3}\tilde{\Theta}_{\alpha\beta}(\mathbf{k})=
785 \epsilon_o {\chi}_Q \left[{\partial_\alpha
786 \tilde{E}^\circ_\beta}(\mathbf{k})+ \frac{1}{24\pi
787 \epsilon_o} \tilde{T}_{\alpha\beta\gamma\delta}(\mathbf{k})
788 \tilde{\Theta}_{\gamma\delta}(\mathbf{k})\right].
789 \label{fourierQuad}
790 \end{equation}
791 If the applied field gradient is homogeneous over the entire volume,
792 ${\partial_ \alpha \tilde{E}^\circ_\beta}(\mathbf{k}) = 0 $ except at
793 $ \mathbf{k} = 0$. Similarly, the quadrupolar polarization density can
794 also considered uniform over entire space. As in the dipolar case,
795 \cite{NeumannI83} the only relevant contribution from the interaction
796 tensor will also be when $\mathbf{k} = 0$. Therefore equation
797 (\ref{fourierQuad}) can be written as,
798 \begin{equation}
799 \frac{1}{3}\tilde{\Theta}_{\alpha\beta}(\mathrm{0})=
800 \epsilon_o {\chi}_Q \left[{\partial_\alpha
801 \tilde{E}^\circ_\beta}(\mathrm{0})+ \frac{1}{24\pi
802 \epsilon_o} \tilde{T}_{\alpha\beta\gamma\delta}(\mathrm{0})
803 \tilde{\Theta}_{\gamma\delta}(\mathrm{0})\right].
804 \label{fourierQuadZeroK}
805 \end{equation}
806 The quadrupolar tensor
807 $\tilde{T}_{\alpha\beta\gamma\delta}(\mathrm{0})$ is a rank 4 tensor
808 with 81 elements. The only non-zero elements, however, are those with
809 two doubly-repeated indices, \textit{i.e.}
810 $\tilde{T}_{aabb}(\mathrm{0})$ and all permutations of these indices.
811 The special case of quadruply-repeated indices,
812 $\tilde{T}_{aaaa}(\mathrm{0})$ also survives (see appendix
813 \ref{ap:quadContraction}). Furthermore, for the both diagonal and
814 non-diagonal components of the quadrupolar polarization
815 $\tilde{\Theta}_{\alpha\beta}$, we can contract the second term in
816 equation \ref{fourierQuadZeroK} (see appendix
817 \ref{ap:quadContraction}):
818 \begin{equation}
819 \tilde{T}_{\alpha\beta\gamma\delta}(\mathrm{0})\tilde{\Theta}_{\gamma\delta}(\mathrm{0})=
820 8 \pi \mathrm{B} \tilde{\Theta}_{\alpha\beta}(\mathrm{0}).
821 \label{quadContraction}
822 \end{equation}
823 Here $\mathrm{B} = \tilde{T}_{abab}(\mathrm{0}) / 4 \pi$ for
824 $a \neq b$. Using this quadrupolar contraction we can solve equation
825 \ref{fourierQuadZeroK} as follows
826 \begin{eqnarray}
827 \frac{1}{3}\tilde{\Theta}_{\alpha\beta}(\mathrm{0}) &=& \epsilon_o
828 {\chi}_Q
829 \left[{\partial_\alpha
830 \tilde{E}^\circ_\beta}(\mathrm{0})+
831 \frac{\mathrm{B}}{3
832 \epsilon_o}
833 {\tilde{\Theta}}_{\alpha\beta}(\mathrm{0})\right]
834 \nonumber \\
835 &=& \left[\frac{\epsilon_o {\chi}_Q} {1-{\chi}_Q \mathrm{B}}\right]
836 {\partial_\alpha \tilde{E}^\circ_\beta}(\mathrm{0}).
837 \label{fourierQuad2}
838 \end{eqnarray}
839 In real space, the correction factor is found to be,
840 \begin{equation}
841 \mathrm{B} = \frac{1}{4 \pi} \tilde{T}_{abab}(0) = \frac{1}{4 \pi} \int_V {T}_{abab}(\mathbf{r}) d\mathbf{r},
842 \end{equation}
843 %If the applied field gradient is homogeneous over the
844 %entire volume, ${\partial_ \alpha \tilde{E}^\circ_\beta}(\mathbf{k}) = 0 $ except at
845 %$ \mathbf{k} = 0$. As in the dipolar case, the only relevant
846 %contribution of the interaction tensor will also be when $\mathbf{k} = 0$.
847 %Therefore equation (\ref{fourierQuad}) can be written as,
848 %\begin{equation}
849 %\frac{1}{3}{\tilde{\Theta}}_{\alpha\beta}({0}) = \epsilon_o {\chi}_Q
850 %\left(1-\frac{1}{4\pi} {\chi}_Q \tilde{T}_{\alpha\beta\alpha\beta}({0})\right)^{-1} \partial_\alpha \tilde{E}^\circ_\beta({0})
851 %\label{fourierQuad2}
852 %\end{equation}
853 %where $\tilde{T}_{\alpha\beta\alpha\beta}({0})$ can be evaluated as,
854 %\begin{equation}
855 %\tilde{T}_{\alpha\beta\alpha\beta}({0}) = \int_V {T}_{\alpha\beta\alpha\beta}(\mathbf{r})d\mathbf{r}
856 %\label{realTensorQaud}
857 %\end{equation}
858 %We may now define a quantity to represent the contribution from the
859 %$\mathbf{k} = 0$ portion of the interaction tensor,
860 %\begin{equation}
861 %B = \frac{1}{4 \pi} \int_V T_{\alpha \beta \alpha \beta}(\mathbf{r})
862 %d\mathbf{r}
863 %\end{equation}
864 which has been integrated over the interaction volume $V$ and has
865 units of $\mathrm{length}^{-2}$.
866
867 In terms of the traced quadrupole moment, equation (\ref{fourierQuad2})
868 can be written,
869 \begin{equation}
870 \mathsf{Q} - \frac{\mathbf{I}}{3} \mathrm{Tr}(\mathsf{Q})
871 = \frac{\epsilon_o {\chi}_Q}{1- {\chi}_Q \mathrm{B}} \nabla \mathbf{E}^\circ
872 \label{tracedConstQuad}
873 \end{equation}
874 Comparing (\ref{tracedConstQuad}) and (\ref{flucQuad}) we obtain,
875 \begin{equation}
876 \frac{\braket{\mathsf{M}_Q^2}-{\braket{\mathsf{M}_Q}}^2}{15 \epsilon_o
877 V k_B T} = \frac{\chi_Q} {1 - \chi_Q \mathrm{B}},
878 \end{equation}
879 or equivalently,
880 \begin{equation}
881 \chi_Q = \frac{\braket{\mathsf{M}_Q^2}-{\braket{\mathsf{M}_Q}}^2}{15 \epsilon_o
882 V k_B T} \left(1 + \mathrm{B} \frac{\braket{\mathsf{M}_Q^2}-{\braket{\mathsf{M}_Q}}^2}{15 \epsilon_o
883 V k_B T} \right)^{-1}
884 \label{eq:finalForm}
885 \end{equation}
886 Eq. (\ref{eq:finalForm}) now expresses a bulk property (the
887 quadrupolar susceptibility, $\chi_Q$) in terms of a fluctuation in the
888 system quadrupole moment and a quadrupolar correction factor
889 ($\mathrm{B}$). The correction factors depend on the cutoff method
890 being employed in the simulation, and these are listed in Table
891 \ref{tab:B}.
892
893 In terms of the macroscopic quadrupole polarizability, $\alpha_Q$,
894 which may be thought of as the ``conducting boundary'' version of the
895 susceptibility,
896 \begin{equation}
897 \chi_Q = \frac{\alpha_Q}{1 + \mathrm{B} \alpha_Q}.
898 \label{eq:quadrupolarSusceptiblity}
899 \end{equation}
900 If an electrostatic method produces $\mathrm{B} \rightarrow 0$, the computed
901 quadrupole polarizability and quadrupole susceptibility converge to
902 the same value.
903
904 \begin{sidewaystable}
905 \caption{Expressions for the quadrupolar correction factor
906 ($\mathrm{B}$) for the real-space electrostatic methods in terms
907 of the damping parameter ($\alpha$) and the cutoff radius
908 ($r_c$). The units of the correction factor are
909 $ \mathrm{length}^{-2}$ for quadrupolar fluids.}
910 \label{tab:B}
911 \begin{tabular}{l|c|c|c} \toprule
912 \multirow{2}{*}{Method} & \multicolumn{3}{c}{Molecular Representation} \\
913 & charges & dipoles & quadrupoles \\\colrule
914 Shifted Potental (SP) & $ -\frac{8 \alpha^5 {r_c}^3e^{-\alpha^2 r_c^2}}{15\sqrt{\pi}} $ & $- \frac{3 \mathrm{erfc(r_c\alpha)}}{5{r_c}^2}- \frac{2 \alpha e^{-\alpha^2 r_c^2}(9+6\alpha^2 r_c^2+4\alpha^4 r_c^4)}{15{\sqrt{\pi}r_c}}$& $ -\frac{16 \alpha^7 {r_c}^5 e^{-\alpha^2 r_c^2 }}{45\sqrt{\pi}}$ \\
915 Gradient-shifted (GSF) & $- \frac{8 \alpha^5 {r_c}^3e^{-\alpha^2 r_c^2}}{15\sqrt{\pi}} $ & 0 & $-\frac{4{\alpha}^7{r_c}^5 e^{-\alpha^2 r_c^2}(-1+2\alpha ^2 r_c^2)}{45\sqrt{\pi}}$\\
916 Taylor-shifted (TSF) & $ -\frac{8 \alpha^5 {r_c}^3e^{-\alpha^2 r_c^2}}{15\sqrt{\pi}} $ & $\frac{4\;\mathrm{erfc(\alpha r_c)}}{{5r_c}^2} + \frac{8\alpha e^{-\alpha^2{r_c}^2}\left(3+ 2\alpha^2 {r_c}^2 +\alpha^4{r_c}^4 \right)}{15\sqrt{\pi}r_c}$ & $\frac{2\;\mathrm{erfc}(\alpha r_c )}{{r_c}^2} + \frac{4{\alpha}e^{-\alpha^2 r_c^2}\left(45 + 30\alpha ^2 {r_c}^2 + 12\alpha^4 {r_c}^4 + 3\alpha^6 {r_c}^6 + 2 \alpha^8 {r_c}^8\right)}{45\sqrt{\pi}{r_c}}$ \\
917 \colrule
918 Spherical Cutoff (SC) & \multicolumn{3}{c}{$ -\frac{8 \alpha^5
919 {r_c}^3e^{-\alpha^2 r_c^2}}{15\sqrt{\pi}} $}\\
920 Ewald-Kornfeld (EK) & \multicolumn{3}{c}{$ -\frac{8 \kappa^5 {r_c}^3e^{-\kappa^2 r_c^2}}{15\sqrt{\pi}}$} \\
921 \botrule
922 \end{tabular}
923 \end{sidewaystable}
924
925 \section{Screening of Charges by Multipolar Fluids}
926 \label{sec:PMF}
927 In a dipolar fluid, the static dielectric constant is also a measure
928 of the ability of the fluid to screen charges from one another. A set
929 of point charges creates an inhomogeneous field in the fluid, and the
930 fluid responds to this field as if it was created externally or via
931 local polarization fluctuations. For this reason, the dielectric
932 constant can be used to estimate an effective potential between two
933 point charges ($C_i$ and $C_j$) embedded in the fluid,
934 \begin{equation}
935 U_\mathrm{effective} = \frac{C_i C_j}{4 \pi \epsilon_0 \epsilon
936 r_{ij}}.
937 \label{eq:effectivePot}
938 \end{equation}
939
940 The same set of point charges can also create an inhomogeneous field
941 \textit{gradient}, and this will cause a response in a quadrupolar
942 fluid that will also cause an effective screening. As discussed above,
943 however, the relevant phyiscal property in quadrupolar fluids is the
944 susceptibility, $\chi_Q$. The screening dielectric associated with
945 the quadrupolar susceptibility is defined as,\cite{Ernst92}
946 \begin{equation}
947 \epsilon = 1 + \chi_Q G = 1 + G \frac{\alpha_Q}{1 + \alpha_Q B}
948 \label{eq:dielectricFromQuadrupoles}
949 \end{equation}
950 where $G$ is a geometrical factor that depends on the geometry of the
951 field perturbation,
952 \begin{equation}
953 G = \frac{\int_V \left| \nabla \mathbf{E}^\circ \right|^2 d\mathbf{r}}
954 {\int_V \left|\mathbf{E}^\circ\right|^2 d\mathbf{r}}
955 \end{equation}
956 integrated over the interaction volume. Note that this geometrical
957 factor is also required to compute effective dielectric constants even
958 when the field gradient is homogeneous over the entire sample.
959
960 To measure effective screening in a multipolar fluid, we compute an
961 effective interaction potential, the potential of mean force (PMF),
962 between positively and negatively charged ions when they screened by
963 the intervening fluid. The PMF is obtained from a sequence of
964 simulations in which two ions are constrained to a fixed distance, and
965 the average constraint force to hold them at a fixed distance $r$ is
966 collected during a long simulation,\cite{Wilfred07}
967 \begin{equation}
968 w(r) = \int_{r_o}^{r}\left\langle \frac{\partial f}{\partial r^\prime}
969 \right\rangle_{r^\prime} dr^\prime + 2k_BT \ln\left(\frac{r}{r_o}\right) + w(r_o),
970 \label{eq:pmf}
971 \end{equation}
972 where $\braket{\partial f/\partial r^\prime}_{r^\prime}$ is the mean
973 constraint force required to hold the ions at distance $r^\prime$,
974 $2k_BT \log(r/r_o)$ is the Fixman factor,\cite{Fixman:1974fk} and
975 $r_o$ is a reference position (usually taken as a large separation
976 betwen the ions). If the dielectric constant is a good measure of the
977 screening at all inter-ion separations, we would expect $w(r)$ to have
978 the form in Eq. (\ref{eq:effectivePot}). Because real fluids are not
979 continuum dielectrics, the effective dielectric constant is a function
980 of the interionic separation,
981 \begin{equation}
982 \epsilon(r) = \frac{u_\mathrm{raw}(r) }{w(r)}
983 \end{equation}
984 where $u_\mathrm{raw}(r)$ is the direct charge-charge interaction
985 potential that is in use during the simulation. $\epsilon(r)$ may
986 vary considerably from the bulk estimates at short distances, although
987 it should converge to the bulk value as the separation between the
988 ions increases.
989
990 \section{Simulation Methodology}
991 To test the formalism developed in the preceding sections, we have
992 carried out computer simulations using three different techniques: i)
993 simulations in the presence of external fields, ii) equilibrium
994 calculations of box moment fluctuations, and iii) potentials of mean
995 force (PMF) between embedded ions. In all cases, the fluids were
996 composed of point multipoles protected by a Lennard-Jones potential.
997 The parameters used in the test systems are given in table
998 \ref{Tab:C}.
999 \begin{table}
1000 \caption{\label{Tab:C}The parameters used in simulations to evaluate
1001 the dielectric response of the new real-space methods.}
1002 \begin{tabularx}{\textwidth}{r|cc|YYccc|Yccc} \hline
1003 & \multicolumn{2}{c|}{LJ parameters} &
1004 \multicolumn{5}{c|}{Electrostatic moments} & & & & \\
1005 Test system & $\sigma$& $\epsilon$ & $C$ & $D$ &
1006 $Q_{xx}$ & $Q_{yy}$ & $Q_{zz}$ & mass & $I_{xx}$ & $I_{yy}$ &
1007 $I_{zz}$ \\ \cline{6-8}\cline{10-12}
1008 & (\AA) & (kcal/mol) & (e) & (debye) & \multicolumn{3}{c|}{(debye \AA)} & (amu) & \multicolumn{3}{c}{(amu
1009 \AA\textsuperscript{2})} \\ \hline
1010 Dipolar fluid & 3.41 & 0.2381 & - & 1.4026 &-&-&-& 39.948 & 11.613 & 11.613 & 0.0 \\
1011 Quadrupolar fluid & 2.985 & 0.265 & - & - & 0.0 & 0.0 &-2.139 & 18.0153 & 43.0565 & 43.0565 & 0.0 \\
1012 \ce{q+} & 1.0 & 0.1 & +1 & - & - & - & - & 22.98 & - & - & - \\
1013 \ce{q-} & 1.0 & 0.1 & -1 & - & - & - & - & 22.98 & - & - & - \\ \hline
1014 \end{tabularx}
1015 \end{table}
1016
1017 The first of the test systems consists entirely of fluids of point
1018 dipolar or quadrupolar molecules in the presence of constant field or
1019 field gradients. Since there are no isolated charges within the
1020 system, the divergence of the field will be zero, \textit{i.e.}
1021 $\vec{\nabla} \cdot \mathbf{E} = 0$. This condition can be satisfied
1022 by using the relatively simple applied potential as described in the
1023 supporting information.
1024
1025 When a constant electric field or field gradient is applied to the
1026 system, the molecules align along the direction of the applied field,
1027 and polarize to a degree determined both by the strength of the field
1028 and the fluid's polarizibility. We have calculated ensemble averages
1029 of the box dipole and quadrupole moments as a function of the strength
1030 of the applied fields. If the fields are sufficiently weak, the
1031 response is linear in the field strength, and one can easily compute
1032 the polarizability directly from the simulations.
1033
1034 The second set of test systems consists of equilibrium simulations of
1035 fluids of point dipolar or quadrupolar molecules simulated in the
1036 absence of any external perturbation. The fluctuation of the ensemble
1037 averages of the box multipolar moment was calculated for each of the
1038 multipolar fluids. The box multipolar moments were computed as simple
1039 sums over the instantaneous molecular moments, and fluctuations in
1040 these quantities were obtained from Eqs. (\ref{eq:flucDip}) and
1041 (\ref{eq:flucQuad}). The macroscopic polarizabilities of the system at
1042 a were derived using Eqs.(\ref{flucDipole}) and (\ref{flucQuad}).
1043
1044 The final system consists of dipolar or quadrupolar fluids with two
1045 oppositely charged ions embedded within the fluid. These ions are
1046 constrained to be at fixed distance throughout a simulation, although
1047 they are allowed to move freely throughout the fluid while satisfying
1048 that constraint. Separate simulations were run at a range of
1049 constraint distances. A dielectric screening factor was computed using
1050 the ratio between the potential between the two ions in the absence of
1051 the fluid medium and the PMF obtained from the simulations.
1052
1053 We carried out these simulations for all three of the new real-space
1054 electrostatic methods (SP, GSF, and TSF) that were developed in the
1055 first paper (Ref. \onlinecite{PaperI}) in the series. The radius of
1056 the cutoff sphere was taken to be 12~\AA. Each of the real space
1057 methods also depends on an adjustable damping parameter $\alpha$ (in
1058 units of $\mathrm{length}^{-1}$). We have selected ten different
1059 values of damping parameter: 0.0, 0.05, 0.1, 0.15, 0.175, 0.2, 0.225,
1060 0.25, 0.3, and 0.35~\AA$^{-1}$ in our simulations of the dipolar
1061 liquids, while four values were chosen for the quadrupolar fluids:
1062 0.0, 0.1, 0.2, and 0.3~\AA$^{-1}$.
1063
1064 For each of the methods and systems listed above, a reference
1065 simulation was carried out using a multipolar implementation of the
1066 Ewald sum.\cite{Smith82,Smith98} A default tolerance of
1067 $1 \times 10^{-8}$~kcal/mol was used in all Ewald calculations,
1068 resulting in Ewald coefficient 0.3119~\AA$^{-1}$ for a cutoff radius
1069 of 12~\AA. All of the electrostatics and constraint methods were
1070 implemented in our group's open source molecular simulation program,
1071 OpenMD,\cite{Meineke05,openmd} which was used for all calculations in
1072 this work.
1073
1074 Dipolar systems contained 2048 Lennard-Jones-protected point dipolar
1075 (Stockmayer) molecules with reduced density $\rho^* = 0.822$,
1076 temperature $T^* = 1.15$, moment of inertia $I^* = 0.025$, and dipole
1077 moment $\mu^* = \sqrt{3.0}$. These systems were equilibrated for
1078 0.5~ns in the canonical (NVT) ensemble. Data collection was carried
1079 out over a 1~ns simulation in the microcanonical (NVE) ensemble. Box
1080 dipole moments were sampled every fs. For simulations with external
1081 perturbations, field strengths ranging from $0 - 10 \times
1082 10^{-4}$~V/\AA\ with increments of $ 10^{-4}$~V/\AA\ were carried out
1083 for each system.
1084
1085 Quadrupolar systems contained 4000 linear point quadrupoles with a
1086 density $2.338 \mathrm{~g/cm}^3$ at a temperature of 500~K. These
1087 systems were equilibrated for 200~ps in a canonical (NVT) ensemble.
1088 Data collection was carried out over a 500~ps simulation in the
1089 microcanonical (NVE) ensemble. Components of box quadrupole moments
1090 were sampled every 100 fs. For quadrupolar simulations with external
1091 field gradients, field strengths ranging from
1092 $0 - 9 \times 10^{-2}$~V/\AA$^2$ with increments of
1093 $10^{-2}$~V/\AA$^2$ were carried out for each system.
1094
1095 To carry out the PMF simulations, two of the multipolar molecules in
1096 the test system were converted into \ce{q+} and \ce{q-} ions and
1097 constrained to remain at a fixed distance for the duration of the
1098 simulation. The constrained distance was then varied from 5--12~\AA.
1099 In the PMF calculations, all simulations were equilibrated for 500 ps
1100 in the NVT ensemble and run for 5 ns in the microcanonical (NVE)
1101 ensemble. Constraint forces were sampled every 20~fs.
1102
1103 \section{Results}
1104 \subsection{Dipolar fluid}
1105 \begin{figure}
1106 \includegraphics[width=5in]{dielectricFinal_Dipole.pdf}
1107 \caption{The polarizability ($\alpha_D$), correction factor ($A$), and
1108 dielectric constant ($\epsilon$) for the test dipolar fluid. The
1109 left panels were computed using external fields, and those on the
1110 right are the result of equilibrium fluctuations. In the GSF and SP
1111 methods, the corrections are large in with small values of $\alpha$,
1112 and a optimal damping coefficient is evident around 0.25 \AA$^{-1}$.
1113 The dashed lines in the upper panel indicate back-calculation of the
1114 polarizability using the Ewald estimate (Refs. \onlinecite{Adams81}
1115 and \onlinecite{NeumannI83}) for the dielectric constant.}
1116 \label{fig:dielectricDipole}
1117 \end{figure}
1118 The macroscopic polarizability ($\alpha_D$) for the dipolar fluid is
1119 shown in the upper panels in Fig. \ref{fig:dielectricDipole}. The
1120 polarizability obtained from the both perturbation and fluctuation
1121 approaches are in excellent agreement with each other. The data also
1122 show a stong dependence on the damping parameter for both the Shifted
1123 Potential (SP) and Gradient Shifted force (GSF) methods, while Taylor
1124 shifted force (TSF) is largely independent of the damping
1125 parameter.
1126
1127 The calculated correction factors discussed in section
1128 \ref{sec:corrFactor} are shown in the middle panels. Because the TSF
1129 method has $A = 1$ for all values of the damping parameter, the
1130 computed polarizabilities need no correction for the dielectric
1131 calculation. The value of $A$ varies with the damping parameter in
1132 both the SP and GSF methods, and inclusion of the correction yields
1133 dielectric estimates (shown in the lower panel) that are generally too
1134 large until the damping reaches $\sim$~0.25~\AA$^{-1}$. Above this
1135 value, the dielectric constants are generally in good agreement with
1136 previous simulation results.\cite{NeumannI83}
1137
1138 Figure \ref{fig:dielectricDipole} also contains back-calculations of
1139 the polarizability using the reference (Ewald) simulation
1140 results.\cite{NeumannI83} These are indicated with dashed lines in the
1141 upper panels. It is clear that the expected polarizability for the SP
1142 and GSF methods are quite close to results obtained from the
1143 simulations. This indicates that the correction formula for the
1144 dipolar fluid (Eq. \ref{correctionFormula}) is quite sensitive when
1145 the value of $\mathrm{A}$ deviates significantly from unity.
1146
1147 These results also suggest an optimal value for the damping parameter
1148 of ($\alpha \sim 0.2-0.3$~\AA$^{-1}$) when evaluating dielectric
1149 constants of point dipolar fluids using either the perturbation and
1150 fluctuation approaches for the new real-space methods.
1151
1152 \begin{figure}
1153 \includegraphics[width=4in]{ScreeningFactor_Dipole.pdf}
1154 \caption{The distance-dependent screening factor, $\epsilon(r)$,
1155 between two ions immersed in the dipolar fluid. The new methods are
1156 shown in separate panels, and different values of the damping
1157 parameter ($\alpha$) are indicated with different symbols. All of
1158 the methods appear to be converging to the bulk dielectric constant
1159 ($\sim 65$) at large ion separations.}
1160 \label{fig:ScreeningFactor_Dipole}
1161 \end{figure}
1162 We have also evaluated the distance-dependent screening factor,
1163 $\epsilon(r)$, between two oppositely charged ions when they are
1164 placed in the dipolar fluid. These results were computed using
1165 Eq. \ref{eq:pmf} and are shown in Fig.
1166 \ref{fig:ScreeningFactor_Dipole}.
1167
1168 The screening factor is similar to the dielectric constant, but
1169 measures a local property of the ions in the fluid and depends on both
1170 ion-dipole and dipole-dipole interactions. These interactions depend
1171 on the distance between ions as well as the electrostatic interaction
1172 methods utilized in the simulations. The screening should converge to
1173 the dielectric constant when the field due to ions is small. This
1174 occurs when the ions are separated (or when the damping parameter is
1175 large). In Fig. \ref{fig:ScreeningFactor_Dipole} we observe that for
1176 the higher value of damping alpha \textit{i.e.}
1177 $\alpha = 0.2$~\AA$^{-1}$ and $0.3$~\AA$^{-1}$ and large separation
1178 between ions, the screening factor does indeed approach the correct
1179 dielectric constant.
1180
1181 REVISIT AFTER EWALD RESULTS:
1182 It is also notable that the TSF method again displays smaller
1183 perturbations away from the correct dielectric screening behavior. We
1184 also observe that for TSF method yields high dielectric screening even
1185 for lower values of $\alpha$.
1186
1187 At short distances, the presence of the ions creates a strong local
1188 field that acts to align nearby dipoles nearly perfectly in opposition
1189 to the field from the ions. This has the effect of increasing the
1190 effective screening when the ions are brought close to one another.
1191
1192 This is due to the absence of corrections for dipole-dipole
1193 interactions as discussed above. However, the TSF remains quite
1194 perturbative for ion-dipole interactions, an effect that is most
1195 apparent at small ion separations.
1196
1197 \subsection{Quadrupolar fluid}
1198 \begin{figure}
1199 \includegraphics[width=4in]{polarizabilityFinal_Quad.pdf}
1200 \caption{The quadrupole polarizability ($\alpha_Q$), correction factor
1201 ($B$), and susceptibility ($\chi_Q$) for the test quadrupolar
1202 fluid. The left panels were computed using external field gradients,
1203 and those on the right are the result of equilibrium fluctuations.
1204 The GSF and SP methods allow nearly unmodified use of the
1205 ``conducting boundary'' or polarizability results in place of the
1206 bulk susceptibility.}
1207 \label{fig:dielectricQuad}
1208 \end{figure}
1209 The polarizability ($\alpha_Q$), correction factor ($B$), and
1210 susceptibility ($\chi_Q$) for the quadrupolar fluid is plotted against
1211 damping parameter Fig. \ref{fig:dielectricQuad}. In quadrupolar
1212 fluids, both the polarizability and susceptibility have units of
1213 $\mathrm{length}^2$. Although the susceptibility has dimensionality, it
1214 is the relevant measure of macroscopic quadrupolar
1215 properties.\cite{JeonI03, JeonII03} The left panel in
1216 Fig. \ref{fig:dielectricQuad} shows results obtained from the applied
1217 field gradient simulations whereas the results from the equilibrium
1218 fluctuation formula are plotted in the right panels.
1219
1220 The susceptibility for the quadrupolar fluid is obtained from
1221 quadrupolarizability and a correction factor using
1222 Eq. (\ref{eq:quadrupolarSusceptiblity}). The susceptibilities are
1223 shown in the bottom panels of Fig. \ref{fig:dielectricQuad}. All three
1224 methods: (SP, GSF, and TSF) produce similar susceptibilities over the
1225 range of damping parameters. This shows that susceptibility derived
1226 using the quadrupolarizability and the correction factors are
1227 essentially independent of the electrostatic method utilized in the
1228 simulation.
1229
1230 \begin{figure}
1231 \includegraphics[width=5in]{ScreeningFactor_Quad.pdf}
1232 \caption{\label{fig:screenQuad} The distance-dependent screening
1233 factor, $\epsilon(r)$, between two ions immersed in the quadrupolar
1234 fluid. Results from the perturbation and fluctuation methods are
1235 shown in right and central panels. Here the susceptibility is
1236 calculated from the simulation and the geometrical factor is
1237 evaluated analytically, using the field and field-gradient produced
1238 by ions. The right hand panel shows the screening factor obtained
1239 from the PMF calculations.}
1240 \end{figure}
1241 A more difficult test of the quadrupolar susceptibility is made by
1242 comparing with direct calculation of the electrostatic screening using
1243 the potential of mean force (PMF). Since the effective dielectric
1244 constant for a quadrupolar fluid depends on the geometry of the field
1245 and field gradient, this is not a physical property of the quadrupolar
1246 fluid.
1247
1248 The geometrical factor for embedded ions changes with the ion
1249 separation distance. It is therefore reasonable to treat the
1250 dielectric constant as a distance-dependent screening factor. Since
1251 the quadrupolar molecules couple with the gradient of the field, the
1252 distribution of the quadrupoles will be inhomogeneously distributed
1253 around the point charges. Hence the distribution of quadrupolar
1254 molecules should be taken into account when computing the geometrical
1255 factors in the presence of this perturbation,
1256 \begin{eqnarray}
1257 G &=& \frac{\int_V g(\mathbf{r}) \left|\nabla \mathbf{E}^\circ
1258 \right|^2 d\mathbf{r}}{\int_V\left|\mathbf{E}^\circ\right|^2
1259 d\mathbf{r}} \nonumber \\ \nonumber \\
1260 &=& \frac{ 2 \pi \int_{-1}^{1} \int_{0}^{R} r^2 g(r,
1261 \cos\theta) \left|\nabla \mathbf{E}^\circ \right|^2 dr d(\cos\theta) }{\int_V\left|\mathbf{E}^\circ\right|^2 d\mathbf{r}}
1262 \label{eq:geometricalFactor}
1263 \end{eqnarray}
1264 where $g(r,\cos\theta)$ is a distribution function for the quadrupoles
1265 with respect to an origin at midpoint of a line joining the two probe
1266 charges.
1267
1268 The effective screening factor is plotted against ion separation
1269 distance in Fig. \ref{fig:screenQuad}. The screening evaluated from
1270 the perturbation and fluctuation methods are shown in right and
1271 central panels. Here the susceptibility is calculated from the
1272 simulation and the geometrical factor is evaluated analytically, using
1273 the field and field-gradient produced by ions. The right hand panel
1274 shows the screening factor obtained from the PMF calculations.
1275
1276 We note that the screening factor obtained from both the perturbation
1277 and fluctuation formula show good agreement with the screening factor
1278 calculated using PMF method. As there are no large differences in
1279 quadrupole-quadruople interactions for various real-space
1280 methods,\cite{PaperI,PaperII} we generally good agreement for the
1281 screening factors using any of the real space methods.
1282
1283 \section{Conclusions}
1284 We have used both perturbation and fluctuation approaches to evaluate
1285 dielectric properties for dipolar and quadrupolar fluids. The static
1286 dielectric constant is the relevant bulk property for dipolar fluids,
1287 while the quadrupolar susceptibility plays a similar role for
1288 quadrupoles. Corrections to both the static dielectric constant and
1289 the quadrupolar susceptibility were derived for three new real space
1290 electrostatic methods, and these corrections were tested against a
1291 third measure of dielectric screening, the potential of mean force
1292 between two ions immersed in the fluids.
1293
1294 For the dipolar fluids, we find that the polarizability evaluated
1295 using the perturbation and fluctuation methods show excellent
1296 agreement, indicating that equilibrium calculations of the dipole
1297 fluctuations are good measures of bulk polarizability. One of the
1298 findings of the second paper in this series is that the moderately
1299 damped GSF and SP methods were most suitable for molecular dynamics
1300 and Monte Carlo simulations, respectively.\cite{PaperII}
1301
1302 The dielectric constant evaluated using the computed polarizability
1303 and correction factors agrees well with the previous Ewald-based
1304 simulation results \cite{Adams81,NeumannI83} for moderate damping
1305 parameters in the range 0.25--0.3~\AA~$^{-1}$.
1306
1307 Although the TSF method alters many dynamic and structural features in
1308 multipolar liquids,\cite{PaperII} it is surprisingly good at computing
1309 bulk dielectric properties at nearly all ranges of the damping
1310 parameter. In fact, the correction factor, $A = 1$, for the TSF
1311 method so the conducting boundary formula is essentially correct when
1312 using this method for point dipolar fluids.
1313
1314 The dielectric correction formula (equation ~\ref{correctionFormula})
1315 is extremely sensitive when the correction factor ($A$) deviates from
1316 1, and this behavior is also seen in the effective screening of ions
1317 embedded in the fluid.
1318
1319 As in the dipolar case, the quadpole polarizability evaluated from
1320 both perturbation and fluctuation simulations show excellent
1321 agreement, again confirming that equilibrium fluctuation calculations
1322 are sufficient to reproduce bulk dielectric properties in these
1323 fluids. The quadrupolar susceptibility calculated via our derived
1324 correction factors tends to produce the same result for all three real
1325 space methods. Similarly, the screening factor calculated using the
1326 susceptibility and a weighted geometric factor provides good agreement
1327 with results obtained directly via potentials of mean force. For
1328 quadrupolar fluids, the distance dependence of the electrostatic
1329 interaction is significantly reduced and the correction factors are
1330 all small. These points suggest that how an electrostatic method
1331 treats the cutoff radius become less consequential for higher order
1332 multipoles.
1333
1334 For this reason, we renew our recommendation that the
1335 moderately-damped GSF method is an excellent choice for molecular
1336 dynamics simulation where point-multipole interactions are being
1337 utilized to compute bulk properties of fluids.
1338
1339 \appendix
1340 \section{Contraction of the quadrupolar tensor with the traceless
1341 quadrupole moment }
1342 \label{ap:quadContraction}
1343 For quadrupolar liquids modeled using point quadrupoles, the
1344 interaction tensor is shown in Eq. (\ref{quadRadial}). The Fourier
1345 transformation of this tensor for $ \mathbf{k} = 0$ is,
1346 \begin{equation}
1347 \tilde{T}_{\alpha\beta\gamma\delta}(0) = \int_V T_{\alpha\beta\gamma\delta}(\mathbf{r}) d \mathbf{r}
1348 \end{equation}
1349 On the basis of symmetry, the 81 elements can be placed in four
1350 different groups: $\tilde{T}_{aaaa}$, $\tilde{T}_{aaab}$,
1351 $\tilde{T}_{aabb}$, and $\tilde{T}_{aabc}$, where $a$, $b$, and $c$,
1352 and can take on distinct values from the set $\left\{x, y, z\right\}$.
1353 The elements belonging to each of these groups can be obtained using
1354 permutations of the indices. Integration of all of the elements shows
1355 that only the groups with indices ${aaaa}$ and ${aabb}$ are non-zero.
1356
1357 We can derive values of the components of $\tilde{T}_{aaaa}$ and
1358 $\tilde{T}_{aabb}$ as follows;
1359 \begin{eqnarray}
1360 \tilde{T}_{xxxx}(0) &=&
1361 \int_{\textrm{V}}
1362 \left[ 3v_{41}(R)+6x^2v_{42}(r)/r^2 + x^4\,v_{43}(r)/r^4 \right] d\mathbf{r} \nonumber \\
1363 &=&12\pi \int_0^{r_c}
1364 \left[ v_{41}(r)+\frac{2}{3} v_{42}(r) + \frac{1}{15}v_{43}(r) \right] r^2\,dr =
1365 \mathrm{12 \pi B}
1366 \end{eqnarray}
1367 and
1368 \begin{eqnarray}
1369 \tilde{T}_{xxyy}(0)&=&
1370 \int_{\textrm{V}}
1371 \left[ v_{41}(R)+(x^2+y^2) v_{42}(r)/r^2 + x^2 y^2\,v_{43}(r)/r^4 \right] d\mathbf{r} \nonumber \\
1372 &=&4\pi \int_0^{r_c}
1373 \left[ v_{41}(r)+\frac{2}{3} v_{42}(r) + \frac{1}{15}v_{43}(r) \right] r^2\,dr =
1374 \mathrm{4 \pi B}.
1375 \end{eqnarray}
1376 These integrals yield the same values for all permutations of the
1377 indices in both tensor element groups. In equation
1378 \ref{fourierQuadZeroK}, for a particular value of the quadrupolar
1379 polarization $\tilde{\Theta}_{aa}$ we can contract
1380 $\tilde{T}_{aa\gamma\delta}(0)$ with $\tilde{\Theta}_{\gamma\delta}$,
1381 using the traceless properties of the quadrupolar moment,
1382 \begin{eqnarray}
1383 \tilde{T}_{xx\gamma\delta}(0)\tilde{\Theta}_{\gamma\delta}(0) &=& \tilde{T}_{xxxx}(0)\tilde{\Theta}_{xx}(0) + \tilde{T}_{xxyy}(0)\tilde{\Theta}_{yy}(0) + \tilde{T}_{xxzz}(0)\tilde{\Theta}_{zz}(0) \nonumber \\
1384 &=& 12 \pi \mathrm{B}\tilde{\Theta}_{xx}(0) +
1385 4 \pi \mathrm{B}\tilde{\Theta}_{yy}(0) +
1386 4 \pi \mathrm{B}\tilde{\Theta}_{zz}(0) \nonumber \\
1387 &=& 8 \pi \mathrm{B}\tilde{\Theta}_{xx}(0) + 4 \pi
1388 \mathrm{B}\left(\tilde{\Theta}_{xx}(0)+\tilde{\Theta}_{yy}(0) +
1389 \tilde{\Theta}_{zz}(0)\right) \nonumber \\
1390 &=& 8 \pi \mathrm{B}\tilde{\Theta}_{xx}(0)
1391 \end{eqnarray}
1392 Similarly for a quadrupolar polarization $\tilde{\Theta}_{xy}$ in
1393 equation \ref{fourierQuadZeroK}, we can contract
1394 $\tilde{T}_{xy\gamma\delta}(0)$ with $\tilde{\Theta}_{\gamma\delta}$,
1395 using the only surviving terms of the tensor,
1396 \begin{eqnarray}
1397 \tilde{T}_{xy\gamma\delta}(0)\tilde{\Theta}_{\gamma\delta}(0) &=& \tilde{T}_{xyxy}(0)\tilde{\Theta}_{xy}(0) + \tilde{T}_{xyyx}(0)\tilde{\Theta}_{yx}(0) \nonumber \\
1398 &=& 4 \pi \mathrm{B}\tilde{\Theta}_{xy}(0) +
1399 4 \pi \mathrm{B}\tilde{\Theta}_{yx}(0) \nonumber \\
1400 &=& 8 \pi \mathrm{B}\tilde{\Theta}_{xy}(0)
1401 \end{eqnarray}
1402 Here, we have used the symmetry of the quadrupole tensor to combine
1403 the symmetric terms. Therefore we can write matrix contraction for
1404 $\tilde{T}_{\alpha\beta\gamma\delta}(\mathrm{0})$ and
1405 $ \tilde{\Theta}_{\gamma\delta}(\mathrm{0})$ in a general form,
1406 \begin{equation}
1407 \tilde{T}_{\alpha\beta\gamma\delta}(\mathrm{0})\tilde{\Theta}_{\gamma\delta}(\mathrm{0})
1408 = 8 \pi \mathrm{B} \tilde{\Theta}_{\alpha\beta}(\mathrm{0}),
1409 \label{contract}
1410 \end{equation}
1411 which is the same as Eq. (\ref{quadContraction}).
1412
1413 When the molecular quadrupoles are represented by point charges, the
1414 symmetry of the quadrupolar tensor is same as for point quadrupoles
1415 (see Eqs.~\ref{quadCharge} and \ref{quadRadial}). However, for
1416 molecular quadrupoles represented by point dipoles, the symmetry of
1417 the quadrupolar tensor must be handled separately (compare
1418 equations~\ref{quadDip} and~\ref{quadRadial}). Although there is a
1419 difference in symmetry, the final result (Eq.~\ref{contract}) also holds
1420 true for dipolar representations.
1421
1422 \section{Quadrupolar correction factor for the Ewald-Kornfeld (EK)
1423 method}
1424 The interaction tensor between two point quadrupoles in the Ewald
1425 method may be expressed,\cite{Smith98,NeumannII83}
1426 \begin{align}
1427 {T}_{\alpha\beta\gamma\delta}(\mathbf{r}) = &\frac{4\pi}{V
1428 }\sum_{k\neq0}^{\infty}
1429 e^{-k^2 / 4
1430 \kappa^2} e^{-i\mathbf{k}\cdot
1431 \mathbf{r}} \left(\frac{r_\alpha r_\beta k_\delta k_\gamma}{k^2}\right) \nonumber \\
1432 &+ \left(\delta_{\alpha\beta}\delta_{\gamma\delta}+\delta_{\alpha\gamma}\delta_{\beta\delta}+\delta_{\alpha\delta}\delta_{\beta\gamma}\right)
1433 B_2(r) \nonumber \\
1434 &- \left(\delta_{\gamma\delta} r_\alpha r_\beta + \mathrm{ 5\; permutations}\right) B_3(r) \nonumber \\
1435 &+ \left(r_\alpha r_\beta r_\gamma r_\delta \right) B_4(r)
1436 \label{ewaldTensor}
1437 \end{align}
1438 where $B_n(r)$ are radial functions defined in reference
1439 \onlinecite{Smith98},
1440 \begin{align}
1441 B_2(r) =& \frac{3}{r^5} \left(\frac{2r\kappa e^{-r^2 \kappa^2}}{\sqrt{\pi}}+\frac{4r^3\kappa^3 e^{-r^2 \kappa^2}}{3\sqrt{\pi}}+\mathrm{erfc(\kappa r)} \right) \\
1442 B_3(r) =& - \frac{15}{r^7}\left(\frac{2r\kappa e^{-r^2 \kappa^2}}{\sqrt{\pi}}+\frac{4r^3\kappa^3 e^{-r^2 \kappa^2}}{3\sqrt{\pi}}+\frac{8r^5\kappa^5 e^{-r^2 \kappa^2}}{15\sqrt{\pi}}+\mathrm{erfc(\kappa r)} \right) \\
1443 B_4(r) =& \frac{105}{r^9}\left(\frac{2r\kappa e^{-r^2
1444 \kappa^2}}{\sqrt{\pi}}+\frac{4r^3\kappa^3 e^{-r^2 \kappa^2}}{3\sqrt{\pi}}+\frac{8r^5\kappa^5 e^{-r^2 \kappa^2}}{15\sqrt{\pi}}
1445 + \frac{16r^7\kappa^7 e^{-r^2 \kappa^2}}{105\sqrt{\pi}} + \mathrm{erfc(\kappa r)} \right)
1446 \end{align}
1447
1448 We can divide ${T}_{\alpha\beta\gamma\delta}(\mathbf{r})$ into three
1449 parts:
1450 \begin{eqnarray}
1451 & & \mathbf{T}(\mathbf{r}) =
1452 \mathbf{T}^\mathrm{K}(\mathbf{r}) +
1453 \mathbf{T}^\mathrm{R1}(\mathbf{r}) +
1454 \mathbf{T}^\mathrm{R2}(\mathbf{r})
1455 \end{eqnarray}
1456 where the first term is the reciprocal space portion. Since the
1457 quadrupolar correction factor $B = \tilde{T}_{abab}(0) / 4\pi$ and
1458 $\mathbf{k} = 0 $ is excluded from the reciprocal space sum,
1459 $\mathbf{T}^\mathrm{K}$ will not contribute.\cite{NeumannII83} The
1460 remaining terms,
1461 \begin{equation}
1462 \mathbf{T}^\mathrm{R1}(\mathbf{r}) = \mathbf{T}^\mathrm{bare}(\mathbf{r}) \left(\frac{2r\kappa e^{-r^2
1463 \kappa^2}}{\sqrt{\pi}}+\frac{4r^3\kappa^3 e^{-r^2 \kappa^2}}{3\sqrt{\pi}}+\frac{8r^5\kappa^5 e^{-r^2 \kappa^2}}{15\sqrt{\pi}}
1464 + \frac{16r^7\kappa^7 e^{-r^2 \kappa^2}}{105\sqrt{\pi}} + \mathrm{erfc(\kappa r)} \right)
1465 \end{equation}
1466 and
1467 \begin{eqnarray}
1468 T^\mathrm{R2}_{\alpha\beta\gamma\delta}(\mathbf{r}) = &+& \left(\delta_{\gamma\delta} r_\alpha r_\beta + \mathrm{ 5\; permutations}\right) \frac{16 \kappa^7 e^{-r^2 \kappa^2}}{7\sqrt{\pi}} \nonumber \\
1469 &-&\left(\delta_{\alpha\beta}\delta_{\gamma\delta}+\delta_{\alpha\gamma}\delta_{\beta\delta}+\delta_{\alpha\delta}\delta_{\beta\gamma}\right) \left(\frac{8 \kappa^5 e^{-r^2 \kappa^2}}{5\sqrt{\pi}}+ \frac{16 r^2\kappa^7 e^{-r^2 \kappa^2}}{35\sqrt{\pi} }\right)
1470 \end{eqnarray}
1471 are contributions from the real space
1472 sum.\cite{Adams76,Adams80,Adams81} Here
1473 $\mathbf{T}^\mathrm{bare}(\mathbf{r})$ is the unmodified quadrupolar
1474 tensor (for undamped quadrupoles). Due to the angular symmetry of the
1475 unmodified tensor, the integral of
1476 $\mathbf{T}^\mathrm{R1}(\mathbf{r})$ will vanish when integrated over
1477 a spherical region. The only term contributing to the correction
1478 factor (B) is therefore
1479 $T^\mathrm{R2}_{\alpha\beta\gamma\delta}(\mathbf{r})$, which allows us
1480 to derive the correction factor for the Ewald-Kornfeld (EK) method,
1481 \begin{eqnarray}
1482 \mathrm{B} &=& \frac{1}{4\pi} \int_V T^\mathrm{R2}_{abab}(\mathbf{r}) \nonumber \\
1483 &=& -\frac{8r_c^3 \kappa^5 e^{-\kappa^2 r_c^2}}{15\sqrt{\pi}}
1484 \end{eqnarray}
1485 which is essentially identical with the correction factor from the
1486 direct spherical cutoff (SC) method.
1487 \newpage
1488 \bibliography{dielectric_new}
1489
1490 \end{document}