ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/multipole/dielectric_new.tex
Revision: 4428
Committed: Fri Apr 15 12:41:57 2016 UTC (9 years, 5 months ago) by gezelter
Content type: application/x-tex
File size: 77851 byte(s)
Log Message:
Latest versions

File Contents

# User Rev Content
1 gezelter 4396 \documentclass[aip,jcp,amsmath,amssymb,preprint]{revtex4-1}
2 mlamichh 4353
3     \usepackage{graphicx}% Include figure files
4     \usepackage{dcolumn}% Align table columns on decimal point
5     \usepackage{multirow}
6 gezelter 4396 \usepackage{bm}
7 mlamichh 4353 \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 mlamichh 4389 \usepackage{tabularx}
14 gezelter 4400 \usepackage{rotating}
15 gezelter 4408 \usepackage{multirow}
16     \usepackage{booktabs}
17 gezelter 4396
18 mlamichh 4389 \newcolumntype{Y}{>{\centering\arraybackslash}X}
19 mlamichh 4353
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 mlamichh 4393 \begin{abstract}
42 gezelter 4399 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 mlamichh 4353 \end{abstract}
61    
62     \maketitle
63    
64     \section{Introduction}
65    
66     Over the past several years, there has been increasing interest in
67 gezelter 4396 pairwise or ``real space'' methods for computing electrostatic
68     interactions in condensed phase
69 gezelter 4420 simulations.\cite{Wolf99,Zahn02,Kast03,Beckd05,Ma05,Wu:2005nr,Fennell06,Fukuda:2013uq,Stenqvist:2015ph,Wang:2016kx}
70 gezelter 4396 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 gezelter 4401 crystals.\cite{Wolf99}
75 mlamichh 4353
76 gezelter 4396 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 gezelter 4399 simulations.\cite{Fennell06} Other recent advances in real-space
80 gezelter 4400 methods for systems of point charges have included explicit
81 gezelter 4401 elimination of the net multipole moments inside the cutoff sphere
82     around each charge site.\cite{Fukuda:2013uq,Wang:2016kx}
83 mlamichh 4353
84 gezelter 4399 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 gezelter 4420 dynamics simulations.
95 gezelter 4399
96 gezelter 4400 One of the most stringent tests of any new electrostatic method is the
97 gezelter 4390 fidelity with which that method can reproduce the bulk-phase
98     polarizability or equivalently, the dielectric properties of a
99 gezelter 4396 fluid. Before the advent of computer simulations, Kirkwood and Onsager
100     developed fluctuation formulae for the dielectric properties of
101 gezelter 4399 dipolar fluids.\cite{Kirkwood39,Onsagar36} Along with projections of
102     the frequency-dependent dielectric to zero frequency, these
103 gezelter 4400 fluctuation formulae are now widely used to predict the static
104     dielectric constant of simulated materials.
105 gezelter 4396
106 gezelter 4399 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 gezelter 4417 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 gezelter 4396
119 gezelter 4400 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 gezelter 4404 the Stockmayer fluid using two different methods: Ewald-Kornfeld (EK)
127 gezelter 4400 and reaction field (RF) methods.\cite{NeumannII83}
128 gezelter 4390
129 gezelter 4392 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 gezelter 4400 al.},\cite{Izvekov08} who noted that the expression for the
133 gezelter 4401 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 gezelter 4390
137 gezelter 4400 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 gezelter 4399 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 gezelter 4392 In quadrupolar fluids, the relationship between quadrupolar
145     susceptibility and the dielectric constant is not as straightforward
146 gezelter 4399 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 gezelter 4400 fluids,\cite{JeonI03,JeonII03,Chitanvis96} although a general
151     correction formula has not yet been developed.
152 gezelter 4390
153 gezelter 4392 In this paper we derive general formulae for calculating the
154 gezelter 4400 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 gezelter 4390
159 gezelter 4400 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 gezelter 4419 simulations (see Fig. \ref{fig:schematic}):
164 gezelter 4390 \begin{enumerate}
165 gezelter 4396 \item responses of the fluid to external perturbations,
166 gezelter 4399 \item fluctuations of system multipole moments, and
167 gezelter 4396 \item potentials of mean force between solvated ions,
168 gezelter 4390 \end{enumerate}
169    
170 gezelter 4408 \begin{figure}
171 gezelter 4428 \includegraphics[width=\linewidth]{Schematic.eps}
172 gezelter 4408 \caption{Dielectric properties of a fluid measure the response to
173 gezelter 4411 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 gezelter 4408 determines the effective dielectric screening.}
180     \label{fig:schematic}
181     \end{figure}
182    
183 gezelter 4400 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 gezelter 4390
194 gezelter 4400 The potential of mean force (PMF) allows calculation of an effective
195 gezelter 4399 dielectric constant or screening factor from the potential energy
196     between ions before and after dielectric material is introduced.
197 gezelter 4400 Computing the PMF between embedded point charges is an additional
198     check on the bulk properties computed via the other two methods.
199 mlamichh 4363
200 gezelter 4400 \section{The Real-Space Methods}
201    
202 gezelter 4404 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 gezelter 4419 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 gezelter 4404
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 gezelter 4399 \section{Dipolar Fluids and the Dielectric Constant}
296 mlamichh 4353
297 gezelter 4399 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 mlamichh 4353
307 gezelter 4399 \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 mlamichh 4353 \begin{equation}
313 gezelter 4400 \braket{\mathbf{p}} = \epsilon_0 \alpha_p \mathbf{E},
314 mlamichh 4353 \end{equation}
315 gezelter 4400 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 mlamichh 4353
320 gezelter 4400 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 mlamichh 4353 \begin{equation}
324 gezelter 4399 \textbf{P} = \epsilon_o \alpha_{D} \mathbf{E}.
325 mlamichh 4353 \label{pertDipole}
326     \end{equation}
327 gezelter 4400 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 mlamichh 4363
333 gezelter 4399 \subsection{Fluctuation Formula}
334 gezelter 4400 % 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 mlamichh 4353
368 gezelter 4400 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 mlamichh 4353
373 gezelter 4400 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 mlamichh 4353 \begin{equation}
377 gezelter 4400 \mathbf{P} = \epsilon_o \frac{\braket{\mathbf{M}^2}-{\braket{\mathbf{M}}}^2}{3 \epsilon_o V k_B T}\bf{E}
378 gezelter 4399 \label{flucDipole}
379 mlamichh 4353 \end{equation}
380 gezelter 4400 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 gezelter 4401 in the net dipole moment,
384 mlamichh 4353 \begin{equation}
385 gezelter 4401 \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 gezelter 4400 \alpha_D = \frac{\braket{\mathbf{M}^2}-{\braket{\mathbf{M}}}^2}{3 \epsilon_o V k_B T}
398 mlamichh 4353 \end{equation}
399 gezelter 4400 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 gezelter 4399 \subsection{Correction Factors}
404 gezelter 4408 \label{sec:corrFactor}
405 gezelter 4400 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 mlamichh 4353 \begin{equation}
409 gezelter 4400 \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 mlamichh 4353 \end{equation}
415 gezelter 4400 $\mathbf{T}$ is the dipole interaction tensor connecting dipoles at
416     $\mathbf{r}^\prime$ with the point of interest ($\mathbf{r}$).
417 mlamichh 4353
418 gezelter 4400 In simulations of dipolar fluids, the molecular dipoles may be
419     represented either by closely-spaced point charges or by
420 gezelter 4408 point dipoles (see Fig. \ref{fig:tensor}).
421 mlamichh 4353 \begin{figure}
422 gezelter 4428 \includegraphics[width=\linewidth]{Tensors.eps}
423 gezelter 4408 \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 mlamichh 4353 \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 mlamichh 4412 \begin{eqnarray}
437 gezelter 4416 \mathbf{T}_{\alpha \beta}(r) &=& \nabla_\alpha \nabla_\beta
438 mlamichh 4412 \left(v(r)\right) \\
439 gezelter 4416 &=& \delta_{\alpha \beta}
440 mlamichh 4353 \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 mlamichh 4412 \end{eqnarray}
445 mlamichh 4353 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 gezelter 4400 interaction between molecular dipoles ($\mathbf{D}$) in Gaussian units
448     as $-\mathbf{D} \cdot \mathbf{T} \cdot \mathbf{D}$.
449 mlamichh 4353
450 gezelter 4400 When utilizing any of the three new real-space methods for point
451     \textit{dipoles}, the tensor is explicitly constructed,
452 mlamichh 4353 \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 gezelter 4400 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 mlamichh 4353
463 gezelter 4400 Using a constitutive relation, the polarization density
464     $\mathbf{P}(\mathbf{r})$ is given by,
465 mlamichh 4353 \begin{equation}
466 gezelter 4400 \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 mlamichh 4353 \label{constDipole}
470     \end{equation}
471 gezelter 4400 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 mlamichh 4353 \begin{equation}
479 gezelter 4400 A = \frac{3}{4\pi}\tilde{\mathbf{T}}(0) = \frac{3}{4\pi} \int_V
480     d\mathbf{r} \mathbf{T}(\mathbf{r})
481 mlamichh 4353 \end{equation}
482 gezelter 4400 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 mlamichh 4353 \begin{equation}
487 gezelter 4400 \epsilon = \frac{3+(A + 2)(\epsilon_{CB}-1)}{3 + (A -1)(\epsilon_{CB}-1)}
488     \label{correctionFormula}
489 mlamichh 4353 \end{equation}
490 gezelter 4400 where $\epsilon_{CB}$ is the widely-used conducting boundary
491     expression for the dielectric constant,
492 mlamichh 4353 \begin{equation}
493 gezelter 4400 \epsilon_{CB} = 1 + \frac{\braket{\bf{M}^2}-{\braket{\bf{M}}}^2}{3
494 mlamichh 4412 \epsilon_o V k_B T} = 1 + \alpha_{D}
495 gezelter 4408 \label{conductingBoundary}
496 mlamichh 4353 \end{equation}
497 gezelter 4408 Eqs. (\ref{correctionFormula}) and (\ref{conductingBoundary}) allows
498     estimation of the static dielectric constant from fluctuations easily
499     computed directly from simulations.
500 gezelter 4400
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 gezelter 4404 We have utilized the Neumann \textit{et al.} approach for the three
530 gezelter 4400 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 gezelter 4425 represent the molecular dipoles. These corrections factors are listed
534     in Table \ref{tab:A}. We note that the GSF correction factor for
535     point dipoles has been independently derived by Stenqvist \textit{et
536     al.}\cite{Stenqvist:2015ph}
537 mlamichh 4353 \begin{table}
538 gezelter 4400 \caption{Expressions for the dipolar correction factor ($A$) for the
539     real-space electrostatic methods in terms of the damping parameter
540 mlamichh 4353 ($\alpha$) and the cutoff radius ($r_c$). The Ewald-Kornfeld result
541 gezelter 4400 derived in Refs. \onlinecite{Adams80,Adams81,NeumannI83} is shown
542     for comparison using the Ewald convergence parameter ($\kappa$)
543     and the real-space cutoff value ($r_c$). }
544 mlamichh 4353 \label{tab:A}
545 gezelter 4416 \begin{tabular}{l|c|c}
546 gezelter 4408 \toprule
547 gezelter 4416 & \multicolumn{2}{c}{Molecular Representation} \\
548 gezelter 4400 Method & point charges & point dipoles \\
549 gezelter 4408 \colrule
550 gezelter 4411 Shifted Potental (SP) & $ \mathrm{erf}(r_c \alpha) - \frac{2
551     \alpha r_c}{\sqrt{\pi}} e^{-\alpha^2
552     r_c^2}$ & $\mathrm{erf}(r_c \alpha)
553     -\frac{2 \alpha
554     r_c}{\sqrt{\pi}}\left(1+\frac{2\alpha^2
555     {r_c}^2}{3}
556     \right)e^{-\alpha^2{r_c}^2}
557     $\\ \colrule
558     Gradient-shifted (GSF) & 1 & $\mathrm{erf}(\alpha r_c)-\frac{2
559     \alpha r_c}{\sqrt{\pi}} \left(1 +
560     \frac{2 \alpha^2 r_c^2}{3} +
561     \frac{\alpha^4
562     r_c^4}{3}\right)e^{-\alpha^2 r_c^2} $ \\ \colrule
563 gezelter 4416 Taylor-shifted (TSF) &\multicolumn{2}{c}{1} \\ \colrule
564     Spherical Cutoff (SC)& \multicolumn{2}{c}{$\mathrm{erf}(r_c \alpha) -
565     \frac{2 \alpha r_c}{\sqrt{\pi}} e^{-\alpha^2
566     r_c^2}$} \\
567     Ewald-Kornfeld (EK) & \multicolumn{2}{c}{$\mathrm{erf}(r_c \kappa) -
568 gezelter 4408 \frac{2 \kappa r_c}{\sqrt{\pi}} e^{-\kappa^2
569     r_c^2}$} \\
570     \botrule
571     \end{tabular}
572 mlamichh 4353 \end{table}
573 gezelter 4400 Note that for point charges, the GSF and TSF methods produce estimates
574     of the dielectric that need no correction, and the TSF method likewise
575     needs no correction for point dipoles.
576 gezelter 4399
577     \section{Quadrupolar Fluids and the Quadrupolar Susceptibility}
578     \subsection{Response to External Perturbations}
579    
580 gezelter 4400 A molecule with a permanent quadrupole, $\mathsf{q}$, will align in
581     the presence of an electric field gradient $\nabla\mathbf{E}$. The
582     anisotropic polarization of the quadrupole is given
583     by,\cite{AduGyamfi78,AduGyamfi81}
584 gezelter 4399 \begin{equation}
585 gezelter 4400 \braket{\mathsf{q}} - \frac{\mathbf{I}}{3}
586     \mathrm{Tr}(\mathsf{q}) = \epsilon_o \alpha_q \nabla\mathbf{E},
587 gezelter 4399 \end{equation}
588 gezelter 4400 where $\alpha_q = q_o^2 / 15 \epsilon_o k_B T $ is a molecular quadrupole
589     polarizablity and $q_o$ is an effective quadrupole moment for the molecule,
590     \begin{equation}
591     q_o^2 = 3 \mathsf{q}:\mathsf{q} - \mathrm{Tr}(\mathsf{q})^2.
592     \end{equation}
593 gezelter 4399
594 gezelter 4400 In the presence of an external field gradient, a system of quadrupolar
595     molecules also organizes with an anisotropic polarization,
596 gezelter 4399 \begin{equation}
597 gezelter 4400 \mathsf{Q} - \frac{\mathbf{I}}{3} \mathrm{Tr}(\mathsf{Q}) = \epsilon_o
598     \alpha_Q \nabla\mathbf{E}
599     \end{equation}
600     where $\mathsf{Q}$ is the traced quadrupole density of the system and
601     $\alpha_Q$ is a macroscopic quadrupole polarizability which has
602     dimensions of $\mathrm{length}^{-2}$. Equivalently, the traceless form
603     may be used,
604     \begin{equation}
605     \mathsf{\Theta} = 3 \epsilon_o \alpha_Q \nabla\mathbf{E},
606 gezelter 4399 \label{pertQuad}
607     \end{equation}
608 gezelter 4400 where
609     $\mathsf{\Theta} = 3\mathsf{Q} - \mathbf{I} \mathrm{Tr}(\mathsf{Q})$
610     is the traceless tensor that also describes the system quadrupole
611     density. It is this tensor that will be utilized to derive correction
612     factors below.
613    
614 gezelter 4399 \subsection{Fluctuation Formula}
615 gezelter 4400 As in the dipolar case, we may define a system quadrupole moment,
616     $\mathsf{M}_Q = \sum_i \mathsf{q}_i$ and the traced quadrupolar
617     density, $\mathsf{Q} = \mathsf{M}_Q / V$. A fluctuation formula can
618     be written for a system comprising quadrupolar
619     molecules,\cite{LoganI81,LoganII82,LoganIII82}
620 gezelter 4399 \begin{equation}
621 gezelter 4400 \mathsf{Q} - \frac{\mathbf{I}}{3} \mathrm{Tr}(\mathsf{Q}) = \epsilon_o
622     \frac{\braket{\mathsf{M}_Q^2}-{\braket{\mathsf{M}_Q}}^2}{15 \epsilon_o
623 gezelter 4419 V k_B T} \nabla\mathbf{E}.
624 gezelter 4399 \label{flucQuad}
625     \end{equation}
626 gezelter 4400 Some care is needed in the definitions of the averaged quantities. These
627     refer to the effective quadrupole moment of the system, and they are
628     computed as follows,
629     \begin{align}
630     \braket{\mathsf{M}_Q^2} &= \braket{3 \mathsf{M}_Q:\mathsf{M}_Q -
631     \mathrm{Tr}(\mathsf{M}_Q)^2 }\\
632     \braket{\mathsf{M}_Q}^2 &= 3 \braket{\mathsf{M}_Q}:\braket{\mathsf{M}_Q} -
633     \mathrm{Tr}(\braket{\mathsf{M}_Q})^2
634 gezelter 4401 \label{eq:flucQuad}
635 gezelter 4400 \end{align}
636     The macroscopic quadrupolarizability is given by,
637 gezelter 4399 \begin{equation}
638 gezelter 4400 \alpha_Q = \frac{\braket{\mathsf{M}_Q^2}-{\braket{\mathsf{M}_Q}}^2}{15 \epsilon_o V k_B T}
639 gezelter 4399 \label{propConstQuad}
640     \end{equation}
641 gezelter 4400 This relationship between a macroscopic property of a quadrupolar
642     fluid and microscopic fluctuations should be true even when the
643     applied field gradient $\nabla \mathbf{E} \rightarrow 0$.
644 gezelter 4399
645     \subsection{Correction Factors}
646 gezelter 4400 In this section we generalize the treatment of Neumann \textit{et al.}
647     for quadrupolar fluids. Interactions involving multiple quadrupoles
648     are rank 4 tensors, and we therefore describe quantities in this
649     section using Einstein notation.
650    
651     In the presence of a uniform external field gradient,
652     $\partial_\alpha {E}^\circ_\beta $, the total field gradient at
653     $\mathbf{r}$ depends on the quadrupole polarization density at all
654     other points in the system,
655 mlamichh 4353 \begin{equation}
656 gezelter 4400 \partial_\alpha E_\beta(\mathbf{r}) = \partial_\alpha
657 mlamichh 4412 {E}^\circ_\beta(\mathbf{r}) + \frac{1}{8\pi \epsilon_o}\int
658 gezelter 4400 T_{\alpha\beta\gamma\delta}({\mathbf{r}-\mathbf{r}^\prime})
659     Q_{\gamma\delta}(\mathbf{r}^\prime) d\mathbf{r}^\prime
660 mlamichh 4353 \label{gradMaxwell}
661     \end{equation}
662 gezelter 4400 where and $T_{\alpha\beta\gamma\delta}$ is the quadrupole interaction
663     tensor connecting quadrupoles at $\mathbf{r}^\prime$ with the point of
664     interest ($\mathbf{r}$).
665 mlamichh 4353
666 gezelter 4408 % \begin{figure}
667     % \includegraphics[width=3in]{QuadrupoleFigure}
668     % \caption{With the real-space electrostatic methods, the molecular
669     % quadrupole tensor, $\mathbf{T}_{\alpha\beta\gamma\delta}(r)$,
670     % governing interactions between molecules is not the same for
671     % quadrupoles represented via sets of charges, point dipoles, or by
672     % single point quadrupoles.}
673     % \label{fig:quadrupolarFluid}
674     % \end{figure}
675 gezelter 4400
676     In simulations of quadrupolar fluids, the molecular quadrupoles may be
677     represented by closely-spaced point charges, by multiple point
678     dipoles, or by a single point quadrupole (see
679 gezelter 4408 Fig. \ref{fig:tensor}). In the case where point charges are
680 gezelter 4400 interacting via an electrostatic kernel, $v(r)$, the effective
681     molecular quadrupole tensor can obtained from four successive
682     applications of the gradient operator to the electrostatic kernel,
683     \begin{eqnarray}
684     T_{\alpha\beta\gamma\delta}(\mathbf{r}) &=& \nabla_\alpha \nabla_\beta
685     \nabla_\gamma \nabla_\delta v(r) \\
686     &=& \left(\delta_{\alpha\beta}\delta_{\gamma\delta} +
687     \delta_{\alpha\gamma}\delta_{\beta\delta}+
688     \delta_{\alpha\delta}\delta_{\beta\gamma}\right)\left(-\frac{v^\prime(r)}{r^3}
689     + \frac{v^{\prime \prime}(r)}{r^2}\right) \nonumber \\
690     & &+ \left(\delta_{\alpha\beta} r_\gamma r_\delta + 5 \mathrm{~permutations}
691     \right) \left(\frac{3v^\prime(r)}{r^5}-\frac{3v^{\prime \prime}(r)}{r^4} +
692     \frac{v^{\prime \prime \prime}(r)}{r^3}\right) \nonumber \\
693     & &+ r_\alpha r_\beta r_\gamma r_\delta
694     \left(-\frac{15v^\prime(r)}{r^7}+\frac{15v^{\prime \prime}(r)}{r^6}-\frac{6v^{\prime
695     \prime \prime}(r)}{r^5} + \frac{v^{\prime \prime \prime \prime}(r)}{r^4}\right),
696 mlamichh 4353 \label{quadCharge}
697 gezelter 4400 \end{eqnarray}
698     where $v(r)$ can either be the electrostatic kernel ($1/r$) or one of
699     the modified (Wolf or DSF) kernels.
700 mlamichh 4353
701 gezelter 4400 Similarly, when representing quadrupolar molecules with multiple point
702 gezelter 4413 \textit{dipoles}, the molecular quadrupole interaction tensor can be
703     obtained using two successive applications of the gradient operator to
704     the dipole interaction tensor,
705 gezelter 4400 \begin{eqnarray}
706     T_{\alpha\beta\gamma\delta}(\mathbf{r}) &=& \nabla_\alpha \nabla_\beta
707 gezelter 4413 T_{\gamma\delta}(\mathbf{r}) \\
708 gezelter 4400 & = & \delta_{\alpha\beta}\delta_{\gamma\delta} \frac{v^\prime_{21}(r)}{r} +
709     \left(\delta_{\alpha\gamma}\delta_{\beta\delta}+
710     \delta_{\alpha\delta}\delta_{\beta\gamma}\right)\frac{v_{22}(r)}{r^2}
711     \nonumber\\
712     & &+ \delta_{\gamma\delta} r_\alpha r_\beta
713     \left(\frac{v^{\prime \prime}_{21}(r)}{r^2}-\frac{v^\prime_{21}(r)}{r^3} \right)\nonumber \\
714     & &+\left(\delta_{\alpha\beta} r_\gamma r_\delta +
715     \delta_{\alpha\gamma} r_\beta r_\delta +\delta_{\alpha\delta}
716     r_\gamma r_\beta + \delta_{\beta\gamma} r_\alpha r_\delta
717     +\delta_{\beta\delta} r_\alpha r_\gamma \right)
718     \left(\frac{v^\prime_{22}(r)}{r^3}-\frac{2v_{22}(r)}{r^4}\right)
719     \nonumber \\
720     & &+ r_\alpha r_\beta r_\gamma r_\delta
721     \left(\frac{v^{\prime
722     \prime}_{22}(r)}{r^4}-\frac{5v^\prime_{22}(r)}{r^5}+\frac{8v_{22}(r)}{r^6}\right),
723 mlamichh 4353 \label{quadDip}
724 gezelter 4400 \end{eqnarray}
725 gezelter 4413 where $T_{\gamma\delta}(\mathbf{r})$ is a dipole-dipole interaction
726 gezelter 4400 tensor that depends on the level of the
727     approximation.\cite{PaperI,PaperII} Similarly $v_{21}(r)$ and
728     $v_{22}(r)$ are the radial function for different real space cutoff
729     methods defined in the first paper in this series.\cite{PaperI}
730    
731     For quadrupolar liquids modeled using point quadrupoles, the
732     interaction tensor can be constructed as,
733     \begin{eqnarray}
734 mlamichh 4412 T_{\alpha\beta\gamma\delta}(\mathbf{r}) &=&
735 gezelter 4400 \left(\delta_{\alpha\beta}\delta_{\gamma\delta}
736     +
737     \delta_{\alpha\gamma}\delta_{\beta\delta}+
738     \delta_{\alpha\delta}\delta_{\beta\gamma}\right)v_{41}(r)
739 mlamichh 4412 + (\delta_{\gamma\delta} r_\alpha r_\beta + \mathrm{ 5\; permutations}) \frac{v_{42}(r)}{r^2} \nonumber \\
740 gezelter 4400 & & + r_\alpha r_\beta r_\gamma r_\delta \left(\frac{v_{43}(r)}{r^4}\right),
741 mlamichh 4353 \label{quadRadial}
742 gezelter 4400 \end{eqnarray}
743     where again $v_{41}(r)$, $v_{42}(r)$, and $v_{43}(r)$ are radial
744     functions defined in Paper I of the series. \cite{PaperI} Note that
745     these radial functions have different functional forms depending on
746     the level of approximation being employed.
747    
748     The integral in equation (\ref{gradMaxwell}) can be divided into two
749     parts, $|\mathbf{r}-\mathbf{r}^\prime|\rightarrow 0 $ and
750     $|\mathbf{r}-\mathbf{r}^\prime|> 0$. Since the self-contribution to
751 gezelter 4408 the field gradient vanishes at the singularity (see the supporting
752     information), equation (\ref{gradMaxwell}) can be written as,
753 mlamichh 4353 \begin{equation}
754 gezelter 4400 \partial_\alpha E_\beta(\mathbf{r}) = \partial_\alpha {E}^\circ_\beta(\mathbf{r}) +
755 mlamichh 4412 \frac{1}{8\pi \epsilon_o}\int\limits_{|\mathbf{r}-\mathbf{r}^\prime|> 0 }
756 gezelter 4400 T_{\alpha\beta\gamma\delta}(\mathbf{r}-\mathbf{r}^\prime)
757     {Q}_{\gamma\delta}(\mathbf{r}^\prime) d\mathbf{r}^\prime.
758 mlamichh 4353 \end{equation}
759 gezelter 4400 If $\mathbf{r} = \mathbf{r}^\prime$ is excluded from the integration,
760     the total gradient can be most easily expressed in terms of
761     traceless quadrupole density as below,\cite{LoganI81}
762 mlamichh 4353 \begin{equation}
763 gezelter 4400 \partial_\alpha E_\beta(\mathbf{r}) = \partial_\alpha
764 mlamichh 4412 {E}^\circ_\beta(\mathbf{r}) + \frac{1}{24\pi
765 gezelter 4400 \epsilon_o}\int\limits_{|\mathbf{r}-\mathbf{r}^\prime|> 0 }
766     T_{\alpha\beta\gamma\delta}(\mathbf{r}-\mathbf{r}^\prime) \Theta_{\gamma\delta}(\mathbf{r}') d\mathbf{r}^\prime,
767 mlamichh 4353 \end{equation}
768 gezelter 4400 where
769     $\Theta_{\alpha\beta} = 3Q_{\alpha\beta} - \delta_{\alpha\beta}Tr(Q)$
770     is the traceless quadrupole density. In analogy to
771 gezelter 4413 Eq. (\ref{pertQuad}) above, the quadrupole polarization density may
772     now be related to the quadrupolar susceptibility, $\chi_Q$,
773 mlamichh 4353 \begin{equation}
774 gezelter 4400 \frac{1}{3}{\Theta}_{\alpha\beta}(\mathbf{r}) = \epsilon_o {\chi}_Q
775 mlamichh 4412 \left[\partial_\alpha {E}^\circ_\beta(\mathbf{r}) + \frac{1}{24\pi
776 gezelter 4400 \epsilon_o}\int\limits_{|\mathbf{r}-\mathbf{r}^\prime|> 0 }
777     T_{\alpha\beta\gamma\delta}(\mathbf{r}-\mathbf{r}^\prime)
778 mlamichh 4412 \Theta_{\gamma\delta}(\mathbf{r}^\prime) d\mathbf{r}^\prime \right].
779 mlamichh 4353 \end{equation}
780 mlamichh 4412 For periodic boundaries and with a uniform imposed
781 gezelter 4400 $\partial_\alpha E^\circ_\beta$, the quadrupole density
782     ${\Theta}_{\alpha\beta}$ will be uniform over the entire space. After
783     performing a Fourier transform (see the Appendix in
784     ref. \onlinecite{NeumannI83}) we obtain,
785 mlamichh 4353 \begin{equation}
786 gezelter 4400 \frac{1}{3}\tilde{\Theta}_{\alpha\beta}(\mathbf{k})=
787     \epsilon_o {\chi}_Q \left[{\partial_\alpha
788 mlamichh 4412 \tilde{E}^\circ_\beta}(\mathbf{k})+ \frac{1}{24\pi
789 gezelter 4400 \epsilon_o} \tilde{T}_{\alpha\beta\gamma\delta}(\mathbf{k})
790     \tilde{\Theta}_{\gamma\delta}(\mathbf{k})\right].
791 mlamichh 4412 \label{fourierQuad}
792 mlamichh 4353 \end{equation}
793 gezelter 4413 If the applied field gradient is homogeneous over the entire volume,
794     ${\partial_ \alpha \tilde{E}^\circ_\beta}(\mathbf{k}) = 0 $ except at
795     $ \mathbf{k} = 0$. Similarly, the quadrupolar polarization density can
796     also considered uniform over entire space. As in the dipolar case,
797     \cite{NeumannI83} the only relevant contribution from the interaction
798     tensor will also be when $\mathbf{k} = 0$. Therefore equation
799     (\ref{fourierQuad}) can be written as,
800 mlamichh 4412 \begin{equation}
801     \frac{1}{3}\tilde{\Theta}_{\alpha\beta}(\mathrm{0})=
802     \epsilon_o {\chi}_Q \left[{\partial_\alpha
803     \tilde{E}^\circ_\beta}(\mathrm{0})+ \frac{1}{24\pi
804     \epsilon_o} \tilde{T}_{\alpha\beta\gamma\delta}(\mathrm{0})
805     \tilde{\Theta}_{\gamma\delta}(\mathrm{0})\right].
806     \label{fourierQuadZeroK}
807     \end{equation}
808 gezelter 4413 The quadrupolar tensor
809 gezelter 4415 $\tilde{T}_{\alpha\beta\gamma\delta}(\mathrm{0})$ is a rank 4 tensor
810 gezelter 4413 with 81 elements. The only non-zero elements, however, are those with
811     two doubly-repeated indices, \textit{i.e.}
812     $\tilde{T}_{aabb}(\mathrm{0})$ and all permutations of these indices.
813     The special case of quadruply-repeated indices,
814     $\tilde{T}_{aaaa}(\mathrm{0})$ also survives (see appendix
815     \ref{ap:quadContraction}). Furthermore, for the both diagonal and
816     non-diagonal components of the quadrupolar polarization
817     $\tilde{\Theta}_{\alpha\beta}$, we can contract the second term in
818     equation \ref{fourierQuadZeroK} (see appendix
819     \ref{ap:quadContraction}):
820 mlamichh 4412 \begin{equation}
821 gezelter 4413 \tilde{T}_{\alpha\beta\gamma\delta}(\mathrm{0})\tilde{\Theta}_{\gamma\delta}(\mathrm{0})=
822     8 \pi \mathrm{B} \tilde{\Theta}_{\alpha\beta}(\mathrm{0}).
823 mlamichh 4412 \label{quadContraction}
824     \end{equation}
825 gezelter 4413 Here $\mathrm{B} = \tilde{T}_{abab}(\mathrm{0}) / 4 \pi$ for
826     $a \neq b$. Using this quadrupolar contraction we can solve equation
827 gezelter 4415 \ref{fourierQuadZeroK} as follows
828 gezelter 4400 \begin{eqnarray}
829 mlamichh 4412 \frac{1}{3}\tilde{\Theta}_{\alpha\beta}(\mathrm{0}) &=& \epsilon_o
830 gezelter 4400 {\chi}_Q
831     \left[{\partial_\alpha
832 mlamichh 4412 \tilde{E}^\circ_\beta}(\mathrm{0})+
833 gezelter 4413 \frac{\mathrm{B}}{3
834 mlamichh 4412 \epsilon_o}
835     {\tilde{\Theta}}_{\alpha\beta}(\mathrm{0})\right]
836     \nonumber \\
837     &=& \left[\frac{\epsilon_o {\chi}_Q} {1-{\chi}_Q \mathrm{B}}\right]
838 gezelter 4413 {\partial_\alpha \tilde{E}^\circ_\beta}(\mathrm{0}).
839 gezelter 4415 \label{fourierQuad2}
840 mlamichh 4412 \end{eqnarray}
841 gezelter 4419 In real space, the correction factor is found to be,
842 gezelter 4413 \begin{equation}
843     \mathrm{B} = \frac{1}{4 \pi} \tilde{T}_{abab}(0) = \frac{1}{4 \pi} \int_V {T}_{abab}(\mathbf{r}) d\mathbf{r},
844     \end{equation}
845 mlamichh 4412 %If the applied field gradient is homogeneous over the
846     %entire volume, ${\partial_ \alpha \tilde{E}^\circ_\beta}(\mathbf{k}) = 0 $ except at
847     %$ \mathbf{k} = 0$. As in the dipolar case, the only relevant
848     %contribution of the interaction tensor will also be when $\mathbf{k} = 0$.
849     %Therefore equation (\ref{fourierQuad}) can be written as,
850     %\begin{equation}
851     %\frac{1}{3}{\tilde{\Theta}}_{\alpha\beta}({0}) = \epsilon_o {\chi}_Q
852     %\left(1-\frac{1}{4\pi} {\chi}_Q \tilde{T}_{\alpha\beta\alpha\beta}({0})\right)^{-1} \partial_\alpha \tilde{E}^\circ_\beta({0})
853     %\label{fourierQuad2}
854     %\end{equation}
855     %where $\tilde{T}_{\alpha\beta\alpha\beta}({0})$ can be evaluated as,
856     %\begin{equation}
857     %\tilde{T}_{\alpha\beta\alpha\beta}({0}) = \int_V {T}_{\alpha\beta\alpha\beta}(\mathbf{r})d\mathbf{r}
858     %\label{realTensorQaud}
859     %\end{equation}
860     %We may now define a quantity to represent the contribution from the
861     %$\mathbf{k} = 0$ portion of the interaction tensor,
862     %\begin{equation}
863     %B = \frac{1}{4 \pi} \int_V T_{\alpha \beta \alpha \beta}(\mathbf{r})
864     %d\mathbf{r}
865     %\end{equation}
866 gezelter 4400 which has been integrated over the interaction volume $V$ and has
867     units of $\mathrm{length}^{-2}$.
868 mlamichh 4353
869 gezelter 4415 In terms of the traced quadrupole moment, equation (\ref{fourierQuad2})
870 gezelter 4413 can be written,
871 mlamichh 4353 \begin{equation}
872 gezelter 4400 \mathsf{Q} - \frac{\mathbf{I}}{3} \mathrm{Tr}(\mathsf{Q})
873 mlamichh 4412 = \frac{\epsilon_o {\chi}_Q}{1- {\chi}_Q \mathrm{B}} \nabla \mathbf{E}^\circ
874 mlamichh 4353 \label{tracedConstQuad}
875     \end{equation}
876 gezelter 4400 Comparing (\ref{tracedConstQuad}) and (\ref{flucQuad}) we obtain,
877 mlamichh 4353 \begin{equation}
878 gezelter 4400 \frac{\braket{\mathsf{M}_Q^2}-{\braket{\mathsf{M}_Q}}^2}{15 \epsilon_o
879 mlamichh 4412 V k_B T} = \frac{\chi_Q} {1 - \chi_Q \mathrm{B}},
880 mlamichh 4353 \end{equation}
881 gezelter 4400 or equivalently,
882 mlamichh 4353 \begin{equation}
883 gezelter 4400 \chi_Q = \frac{\braket{\mathsf{M}_Q^2}-{\braket{\mathsf{M}_Q}}^2}{15 \epsilon_o
884 mlamichh 4412 V k_B T} \left(1 + \mathrm{B} \frac{\braket{\mathsf{M}_Q^2}-{\braket{\mathsf{M}_Q}}^2}{15 \epsilon_o
885 gezelter 4400 V k_B T} \right)^{-1}
886     \label{eq:finalForm}
887 mlamichh 4353 \end{equation}
888 gezelter 4400 Eq. (\ref{eq:finalForm}) now expresses a bulk property (the
889 gezelter 4413 quadrupolar susceptibility, $\chi_Q$) in terms of a fluctuation in the
890     system quadrupole moment and a quadrupolar correction factor
891     ($\mathrm{B}$). The correction factors depend on the cutoff method
892     being employed in the simulation, and these are listed in Table
893     \ref{tab:B}.
894 mlamichh 4353
895 gezelter 4400 In terms of the macroscopic quadrupole polarizability, $\alpha_Q$,
896     which may be thought of as the ``conducting boundary'' version of the
897     susceptibility,
898 mlamichh 4353 \begin{equation}
899 mlamichh 4412 \chi_Q = \frac{\alpha_Q}{1 + \mathrm{B} \alpha_Q}.
900 gezelter 4408 \label{eq:quadrupolarSusceptiblity}
901 mlamichh 4353 \end{equation}
902 mlamichh 4412 If an electrostatic method produces $\mathrm{B} \rightarrow 0$, the computed
903 gezelter 4400 quadrupole polarizability and quadrupole susceptibility converge to
904     the same value.
905    
906     \begin{sidewaystable}
907 gezelter 4415 \caption{Expressions for the quadrupolar correction factor
908     ($\mathrm{B}$) for the real-space electrostatic methods in terms
909     of the damping parameter ($\alpha$) and the cutoff radius
910     ($r_c$). The units of the correction factor are
911     $ \mathrm{length}^{-2}$ for quadrupolar fluids.}
912 mlamichh 4353 \label{tab:B}
913 gezelter 4416 \begin{tabular}{l|c|c|c} \toprule
914     \multirow{2}{*}{Method} & \multicolumn{3}{c}{Molecular Representation} \\
915     & charges & dipoles & quadrupoles \\\colrule
916 mlamichh 4412 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}}$ \\
917     Gradient-shifted (GSF) & $- \frac{8 \alpha^5 {r_c}^3e^{-\alpha^2 r_c^2}}{15\sqrt{\pi}} $ & 0 & $-\frac{4{\alpha}^7{r_c}^5 e^{-\alpha^2 r_c^2}(-1+2\alpha ^2 r_c^2)}{45\sqrt{\pi}}$\\
918     Taylor-shifted (TSF) & $ -\frac{8 \alpha^5 {r_c}^3e^{-\alpha^2 r_c^2}}{15\sqrt{\pi}} $ & $\frac{4\;\mathrm{erfc(\alpha r_c)}}{{5r_c}^2} + \frac{8\alpha e^{-\alpha^2{r_c}^2}\left(3+ 2\alpha^2 {r_c}^2 +\alpha^4{r_c}^4 \right)}{15\sqrt{\pi}r_c}$ & $\frac{2\;\mathrm{erfc}(\alpha r_c )}{{r_c}^2} + \frac{4{\alpha}e^{-\alpha^2 r_c^2}\left(45 + 30\alpha ^2 {r_c}^2 + 12\alpha^4 {r_c}^4 + 3\alpha^6 {r_c}^6 + 2 \alpha^8 {r_c}^8\right)}{45\sqrt{\pi}{r_c}}$ \\
919 gezelter 4408 \colrule
920 gezelter 4416 Spherical Cutoff (SC) & \multicolumn{3}{c}{$ -\frac{8 \alpha^5
921     {r_c}^3e^{-\alpha^2 r_c^2}}{15\sqrt{\pi}} $}\\
922     Ewald-Kornfeld (EK) & \multicolumn{3}{c}{$ -\frac{8 \kappa^5 {r_c}^3e^{-\kappa^2 r_c^2}}{15\sqrt{\pi}}$} \\
923 gezelter 4408 \botrule
924 gezelter 4400 \end{tabular}
925     \end{sidewaystable}
926 gezelter 4399
927     \section{Screening of Charges by Multipolar Fluids}
928 gezelter 4401 \label{sec:PMF}
929     In a dipolar fluid, the static dielectric constant is also a measure
930     of the ability of the fluid to screen charges from one another. A set
931     of point charges creates an inhomogeneous field in the fluid, and the
932     fluid responds to this field as if it was created externally or via
933     local polarization fluctuations. For this reason, the dielectric
934     constant can be used to estimate an effective potential between two
935     point charges ($C_i$ and $C_j$) embedded in the fluid,
936 gezelter 4399 \begin{equation}
937 gezelter 4401 U_\mathrm{effective} = \frac{C_i C_j}{4 \pi \epsilon_0 \epsilon
938     r_{ij}}.
939     \label{eq:effectivePot}
940 gezelter 4399 \end{equation}
941 gezelter 4401
942     The same set of point charges can also create an inhomogeneous field
943     \textit{gradient}, and this will cause a response in a quadrupolar
944     fluid that will also cause an effective screening. As discussed above,
945     however, the relevant phyiscal property in quadrupolar fluids is the
946     susceptibility, $\chi_Q$. The screening dielectric associated with
947     the quadrupolar susceptibility is defined as,\cite{Ernst92}
948 gezelter 4399 \begin{equation}
949 gezelter 4401 \epsilon = 1 + \chi_Q G = 1 + G \frac{\alpha_Q}{1 + \alpha_Q B}
950     \label{eq:dielectricFromQuadrupoles}
951 gezelter 4399 \end{equation}
952 gezelter 4401 where $G$ is a geometrical factor that depends on the geometry of the
953     field perturbation,
954     \begin{equation}
955     G = \frac{\int_V \left| \nabla \mathbf{E}^\circ \right|^2 d\mathbf{r}}
956     {\int_V \left|\mathbf{E}^\circ\right|^2 d\mathbf{r}}
957     \end{equation}
958     integrated over the interaction volume. Note that this geometrical
959     factor is also required to compute effective dielectric constants even
960     when the field gradient is homogeneous over the entire sample.
961 gezelter 4399
962 gezelter 4401 To measure effective screening in a multipolar fluid, we compute an
963     effective interaction potential, the potential of mean force (PMF),
964     between positively and negatively charged ions when they screened by
965     the intervening fluid. The PMF is obtained from a sequence of
966     simulations in which two ions are constrained to a fixed distance, and
967     the average constraint force to hold them at a fixed distance $r$ is
968     collected during a long simulation,\cite{Wilfred07}
969 gezelter 4400 \begin{equation}
970 gezelter 4401 w(r) = \int_{r_o}^{r}\left\langle \frac{\partial f}{\partial r^\prime}
971     \right\rangle_{r^\prime} dr^\prime + 2k_BT \ln\left(\frac{r}{r_o}\right) + w(r_o),
972     \label{eq:pmf}
973 gezelter 4400 \end{equation}
974 gezelter 4401 where $\braket{\partial f/\partial r^\prime}_{r^\prime}$ is the mean
975     constraint force required to hold the ions at distance $r^\prime$,
976     $2k_BT \log(r/r_o)$ is the Fixman factor,\cite{Fixman:1974fk} and
977     $r_o$ is a reference position (usually taken as a large separation
978     betwen the ions). If the dielectric constant is a good measure of the
979     screening at all inter-ion separations, we would expect $w(r)$ to have
980     the form in Eq. (\ref{eq:effectivePot}). Because real fluids are not
981     continuum dielectrics, the effective dielectric constant is a function
982     of the interionic separation,
983 gezelter 4400 \begin{equation}
984 gezelter 4425 \epsilon(r) = \frac{u_\mathrm{raw}(r) - u_\mathrm{raw}(r_o) }{w(r) - w(r_o)}
985 gezelter 4400 \end{equation}
986 gezelter 4401 where $u_\mathrm{raw}(r)$ is the direct charge-charge interaction
987     potential that is in use during the simulation. $\epsilon(r)$ may
988     vary considerably from the bulk estimates at short distances, although
989     it should converge to the bulk value as the separation between the
990     ions increases.
991 gezelter 4400
992 gezelter 4399 \section{Simulation Methodology}
993 mlamichh 4422
994 gezelter 4401 To test the formalism developed in the preceding sections, we have
995     carried out computer simulations using three different techniques: i)
996     simulations in the presence of external fields, ii) equilibrium
997     calculations of box moment fluctuations, and iii) potentials of mean
998     force (PMF) between embedded ions. In all cases, the fluids were
999     composed of point multipoles protected by a Lennard-Jones potential.
1000     The parameters used in the test systems are given in table
1001     \ref{Tab:C}.
1002 mlamichh 4422
1003     \begin{sidewaystable}
1004 gezelter 4401 \caption{\label{Tab:C}The parameters used in simulations to evaluate
1005     the dielectric response of the new real-space methods.}
1006 mlamichh 4389 \begin{tabularx}{\textwidth}{r|cc|YYccc|Yccc} \hline
1007     & \multicolumn{2}{c|}{LJ parameters} &
1008     \multicolumn{5}{c|}{Electrostatic moments} & & & & \\
1009     Test system & $\sigma$& $\epsilon$ & $C$ & $D$ &
1010     $Q_{xx}$ & $Q_{yy}$ & $Q_{zz}$ & mass & $I_{xx}$ & $I_{yy}$ &
1011     $I_{zz}$ \\ \cline{6-8}\cline{10-12}
1012     & (\AA) & (kcal/mol) & (e) & (debye) & \multicolumn{3}{c|}{(debye \AA)} & (amu) & \multicolumn{3}{c}{(amu
1013     \AA\textsuperscript{2})} \\ \hline
1014 gezelter 4401 Dipolar fluid & 3.41 & 0.2381 & - & 1.4026 &-&-&-& 39.948 & 11.613 & 11.613 & 0.0 \\
1015     Quadrupolar fluid & 2.985 & 0.265 & - & - & 0.0 & 0.0 &-2.139 & 18.0153 & 43.0565 & 43.0565 & 0.0 \\
1016 mlamichh 4389 \ce{q+} & 1.0 & 0.1 & +1 & - & - & - & - & 22.98 & - & - & - \\
1017     \ce{q-} & 1.0 & 0.1 & -1 & - & - & - & - & 22.98 & - & - & - \\ \hline
1018     \end{tabularx}
1019 mlamichh 4422 \end{sidewaystable}
1020 mlamichh 4363
1021 gezelter 4401 The first of the test systems consists entirely of fluids of point
1022     dipolar or quadrupolar molecules in the presence of constant field or
1023     field gradients. Since there are no isolated charges within the
1024     system, the divergence of the field will be zero, \textit{i.e.}
1025     $\vec{\nabla} \cdot \mathbf{E} = 0$. This condition can be satisfied
1026     by using the relatively simple applied potential as described in the
1027     supporting information.
1028 mlamichh 4353
1029 gezelter 4401 When a constant electric field or field gradient is applied to the
1030     system, the molecules align along the direction of the applied field,
1031     and polarize to a degree determined both by the strength of the field
1032     and the fluid's polarizibility. We have calculated ensemble averages
1033     of the box dipole and quadrupole moments as a function of the strength
1034     of the applied fields. If the fields are sufficiently weak, the
1035     response is linear in the field strength, and one can easily compute
1036     the polarizability directly from the simulations.
1037 mlamichh 4389
1038 gezelter 4401 The second set of test systems consists of equilibrium simulations of
1039     fluids of point dipolar or quadrupolar molecules simulated in the
1040     absence of any external perturbation. The fluctuation of the ensemble
1041     averages of the box multipolar moment was calculated for each of the
1042     multipolar fluids. The box multipolar moments were computed as simple
1043     sums over the instantaneous molecular moments, and fluctuations in
1044     these quantities were obtained from Eqs. (\ref{eq:flucDip}) and
1045     (\ref{eq:flucQuad}). The macroscopic polarizabilities of the system at
1046     a were derived using Eqs.(\ref{flucDipole}) and (\ref{flucQuad}).
1047 mlamichh 4389
1048 gezelter 4401 The final system consists of dipolar or quadrupolar fluids with two
1049     oppositely charged ions embedded within the fluid. These ions are
1050     constrained to be at fixed distance throughout a simulation, although
1051     they are allowed to move freely throughout the fluid while satisfying
1052     that constraint. Separate simulations were run at a range of
1053     constraint distances. A dielectric screening factor was computed using
1054     the ratio between the potential between the two ions in the absence of
1055     the fluid medium and the PMF obtained from the simulations.
1056 mlamichh 4389
1057 gezelter 4401 We carried out these simulations for all three of the new real-space
1058     electrostatic methods (SP, GSF, and TSF) that were developed in the
1059     first paper (Ref. \onlinecite{PaperI}) in the series. The radius of
1060     the cutoff sphere was taken to be 12~\AA. Each of the real space
1061     methods also depends on an adjustable damping parameter $\alpha$ (in
1062     units of $\mathrm{length}^{-1}$). We have selected ten different
1063     values of damping parameter: 0.0, 0.05, 0.1, 0.15, 0.175, 0.2, 0.225,
1064     0.25, 0.3, and 0.35~\AA$^{-1}$ in our simulations of the dipolar
1065     liquids, while four values were chosen for the quadrupolar fluids:
1066     0.0, 0.1, 0.2, and 0.3~\AA$^{-1}$.
1067 mlamichh 4389
1068 gezelter 4401 For each of the methods and systems listed above, a reference
1069     simulation was carried out using a multipolar implementation of the
1070     Ewald sum.\cite{Smith82,Smith98} A default tolerance of
1071     $1 \times 10^{-8}$~kcal/mol was used in all Ewald calculations,
1072     resulting in Ewald coefficient 0.3119~\AA$^{-1}$ for a cutoff radius
1073     of 12~\AA. All of the electrostatics and constraint methods were
1074     implemented in our group's open source molecular simulation program,
1075     OpenMD,\cite{Meineke05,openmd} which was used for all calculations in
1076     this work.
1077    
1078     Dipolar systems contained 2048 Lennard-Jones-protected point dipolar
1079     (Stockmayer) molecules with reduced density $\rho^* = 0.822$,
1080     temperature $T^* = 1.15$, moment of inertia $I^* = 0.025$, and dipole
1081     moment $\mu^* = \sqrt{3.0}$. These systems were equilibrated for
1082     0.5~ns in the canonical (NVT) ensemble. Data collection was carried
1083     out over a 1~ns simulation in the microcanonical (NVE) ensemble. Box
1084     dipole moments were sampled every fs. For simulations with external
1085 gezelter 4408 perturbations, field strengths ranging from $0 - 10 \times
1086 gezelter 4415 10^{-4}$~V/\AA\ with increments of $ 10^{-4}$~V/\AA\ were carried out
1087 gezelter 4408 for each system.
1088 gezelter 4401
1089     Quadrupolar systems contained 4000 linear point quadrupoles with a
1090 gezelter 4408 density $2.338 \mathrm{~g/cm}^3$ at a temperature of 500~K. These
1091 gezelter 4401 systems were equilibrated for 200~ps in a canonical (NVT) ensemble.
1092     Data collection was carried out over a 500~ps simulation in the
1093 gezelter 4408 microcanonical (NVE) ensemble. Components of box quadrupole moments
1094     were sampled every 100 fs. For quadrupolar simulations with external
1095 gezelter 4401 field gradients, field strengths ranging from
1096 gezelter 4408 $0 - 9 \times 10^{-2}$~V/\AA$^2$ with increments of
1097     $10^{-2}$~V/\AA$^2$ were carried out for each system.
1098 gezelter 4401
1099     To carry out the PMF simulations, two of the multipolar molecules in
1100     the test system were converted into \ce{q+} and \ce{q-} ions and
1101     constrained to remain at a fixed distance for the duration of the
1102 gezelter 4408 simulation. The constrained distance was then varied from 5--12~\AA.
1103     In the PMF calculations, all simulations were equilibrated for 500 ps
1104     in the NVT ensemble and run for 5 ns in the microcanonical (NVE)
1105     ensemble. Constraint forces were sampled every 20~fs.
1106 mlamichh 4389
1107     \section{Results}
1108 mlamichh 4397 \subsection{Dipolar fluid}
1109     \begin{figure}
1110 gezelter 4428 \includegraphics[width=5in]{dielectricFinal_Dipole.eps}
1111 gezelter 4401 \caption{The polarizability ($\alpha_D$), correction factor ($A$), and
1112     dielectric constant ($\epsilon$) for the test dipolar fluid. The
1113     left panels were computed using external fields, and those on the
1114     right are the result of equilibrium fluctuations. In the GSF and SP
1115     methods, the corrections are large in with small values of $\alpha$,
1116     and a optimal damping coefficient is evident around 0.25 \AA$^{-1}$.
1117     The dashed lines in the upper panel indicate back-calculation of the
1118 gezelter 4404 polarizability using the Ewald estimate (Refs. \onlinecite{Adams81}
1119     and \onlinecite{NeumannI83}) for the dielectric constant.}
1120 mlamichh 4397 \label{fig:dielectricDipole}
1121     \end{figure}
1122 gezelter 4401 The macroscopic polarizability ($\alpha_D$) for the dipolar fluid is
1123 gezelter 4408 shown in the upper panels in Fig. \ref{fig:dielectricDipole}. The
1124 gezelter 4401 polarizability obtained from the both perturbation and fluctuation
1125     approaches are in excellent agreement with each other. The data also
1126     show a stong dependence on the damping parameter for both the Shifted
1127     Potential (SP) and Gradient Shifted force (GSF) methods, while Taylor
1128     shifted force (TSF) is largely independent of the damping
1129     parameter.
1130 mlamichh 4397
1131 gezelter 4401 The calculated correction factors discussed in section
1132 gezelter 4408 \ref{sec:corrFactor} are shown in the middle panels. Because the TSF
1133 gezelter 4401 method has $A = 1$ for all values of the damping parameter, the
1134     computed polarizabilities need no correction for the dielectric
1135     calculation. The value of $A$ varies with the damping parameter in
1136     both the SP and GSF methods, and inclusion of the correction yields
1137     dielectric estimates (shown in the lower panel) that are generally too
1138 gezelter 4408 large until the damping reaches $\sim$~0.25~\AA$^{-1}$. Above this
1139     value, the dielectric constants are generally in good agreement with
1140     previous simulation results.\cite{NeumannI83}
1141 mlamichh 4397
1142 gezelter 4401 Figure \ref{fig:dielectricDipole} also contains back-calculations of
1143     the polarizability using the reference (Ewald) simulation
1144     results.\cite{NeumannI83} These are indicated with dashed lines in the
1145     upper panels. It is clear that the expected polarizability for the SP
1146     and GSF methods are quite close to results obtained from the
1147     simulations. This indicates that the correction formula for the
1148     dipolar fluid (Eq. \ref{correctionFormula}) is quite sensitive when
1149     the value of $\mathrm{A}$ deviates significantly from unity.
1150    
1151     These results also suggest an optimal value for the damping parameter
1152 gezelter 4419 of ($\alpha \sim 0.2-0.3$~\AA$^{-1}$) when evaluating dielectric
1153 gezelter 4408 constants of point dipolar fluids using either the perturbation and
1154     fluctuation approaches for the new real-space methods.
1155 gezelter 4401
1156 mlamichh 4397 \begin{figure}
1157 gezelter 4428 \includegraphics[width=4in]{screen_pmf.eps}
1158 gezelter 4401 \caption{The distance-dependent screening factor, $\epsilon(r)$,
1159     between two ions immersed in the dipolar fluid. The new methods are
1160     shown in separate panels, and different values of the damping
1161     parameter ($\alpha$) are indicated with different symbols. All of
1162     the methods appear to be converging to the bulk dielectric constant
1163 gezelter 4415 ($\sim 65$) at large ion separations.}
1164 mlamichh 4397 \label{fig:ScreeningFactor_Dipole}
1165     \end{figure}
1166 gezelter 4401 We have also evaluated the distance-dependent screening factor,
1167     $\epsilon(r)$, between two oppositely charged ions when they are
1168     placed in the dipolar fluid. These results were computed using
1169     Eq. \ref{eq:pmf} and are shown in Fig.
1170     \ref{fig:ScreeningFactor_Dipole}.
1171 mlamichh 4397
1172 gezelter 4401 The screening factor is similar to the dielectric constant, but
1173     measures a local property of the ions in the fluid and depends on both
1174     ion-dipole and dipole-dipole interactions. These interactions depend
1175     on the distance between ions as well as the electrostatic interaction
1176     methods utilized in the simulations. The screening should converge to
1177     the dielectric constant when the field due to ions is small. This
1178     occurs when the ions are separated (or when the damping parameter is
1179     large). In Fig. \ref{fig:ScreeningFactor_Dipole} we observe that for
1180     the higher value of damping alpha \textit{i.e.}
1181 gezelter 4408 $\alpha = 0.2$~\AA$^{-1}$ and $0.3$~\AA$^{-1}$ and large separation
1182 gezelter 4401 between ions, the screening factor does indeed approach the correct
1183     dielectric constant.
1184 mlamichh 4397
1185 gezelter 4401 It is also notable that the TSF method again displays smaller
1186     perturbations away from the correct dielectric screening behavior. We
1187     also observe that for TSF method yields high dielectric screening even
1188 gezelter 4428 for lower values of $\alpha$.
1189 gezelter 4401
1190     At short distances, the presence of the ions creates a strong local
1191     field that acts to align nearby dipoles nearly perfectly in opposition
1192     to the field from the ions. This has the effect of increasing the
1193     effective screening when the ions are brought close to one another.
1194 gezelter 4428 This effect is present even in the full Ewald treatment, and indicates
1195     that the local ordering behavior is being captured by all of the
1196     moderately-damped real-space methods.
1197 gezelter 4401
1198 mlamichh 4397 \subsection{Quadrupolar fluid}
1199     \begin{figure}
1200 gezelter 4428 \includegraphics[width=4in]{polarizabilityFinal_Quad.eps}
1201 gezelter 4401 \caption{The quadrupole polarizability ($\alpha_Q$), correction factor
1202     ($B$), and susceptibility ($\chi_Q$) for the test quadrupolar
1203     fluid. The left panels were computed using external field gradients,
1204     and those on the right are the result of equilibrium fluctuations.
1205     The GSF and SP methods allow nearly unmodified use of the
1206     ``conducting boundary'' or polarizability results in place of the
1207 gezelter 4404 bulk susceptibility.}
1208 mlamichh 4397 \label{fig:dielectricQuad}
1209     \end{figure}
1210 gezelter 4401 The polarizability ($\alpha_Q$), correction factor ($B$), and
1211     susceptibility ($\chi_Q$) for the quadrupolar fluid is plotted against
1212     damping parameter Fig. \ref{fig:dielectricQuad}. In quadrupolar
1213     fluids, both the polarizability and susceptibility have units of
1214 gezelter 4419 $\mathrm{length}^2$. Although the susceptibility has dimensionality, it
1215 gezelter 4401 is the relevant measure of macroscopic quadrupolar
1216 gezelter 4419 properties.\cite{JeonI03, JeonII03} The left panel in
1217 gezelter 4401 Fig. \ref{fig:dielectricQuad} shows results obtained from the applied
1218     field gradient simulations whereas the results from the equilibrium
1219     fluctuation formula are plotted in the right panels.
1220 mlamichh 4397
1221 gezelter 4401 The susceptibility for the quadrupolar fluid is obtained from
1222     quadrupolarizability and a correction factor using
1223     Eq. (\ref{eq:quadrupolarSusceptiblity}). The susceptibilities are
1224     shown in the bottom panels of Fig. \ref{fig:dielectricQuad}. All three
1225     methods: (SP, GSF, and TSF) produce similar susceptibilities over the
1226     range of damping parameters. This shows that susceptibility derived
1227     using the quadrupolarizability and the correction factors are
1228     essentially independent of the electrostatic method utilized in the
1229     simulation.
1230 mlamichh 4397
1231     \begin{figure}
1232 gezelter 4428 \includegraphics[width=\linewidth]{screen_all.eps}
1233 gezelter 4401 \caption{\label{fig:screenQuad} The distance-dependent screening
1234     factor, $\epsilon(r)$, between two ions immersed in the quadrupolar
1235     fluid. Results from the perturbation and fluctuation methods are
1236     shown in right and central panels. Here the susceptibility is
1237     calculated from the simulation and the geometrical factor is
1238     evaluated analytically, using the field and field-gradient produced
1239     by ions. The right hand panel shows the screening factor obtained
1240     from the PMF calculations.}
1241 mlamichh 4397 \end{figure}
1242 gezelter 4401 A more difficult test of the quadrupolar susceptibility is made by
1243     comparing with direct calculation of the electrostatic screening using
1244     the potential of mean force (PMF). Since the effective dielectric
1245     constant for a quadrupolar fluid depends on the geometry of the field
1246     and field gradient, this is not a physical property of the quadrupolar
1247     fluid.
1248 mlamichh 4397
1249 gezelter 4401 The geometrical factor for embedded ions changes with the ion
1250     separation distance. It is therefore reasonable to treat the
1251     dielectric constant as a distance-dependent screening factor. Since
1252     the quadrupolar molecules couple with the gradient of the field, the
1253     distribution of the quadrupoles will be inhomogeneously distributed
1254     around the point charges. Hence the distribution of quadrupolar
1255     molecules should be taken into account when computing the geometrical
1256     factors in the presence of this perturbation,
1257     \begin{eqnarray}
1258     G &=& \frac{\int_V g(\mathbf{r}) \left|\nabla \mathbf{E}^\circ
1259     \right|^2 d\mathbf{r}}{\int_V\left|\mathbf{E}^\circ\right|^2
1260     d\mathbf{r}} \nonumber \\ \nonumber \\
1261     &=& \frac{ 2 \pi \int_{-1}^{1} \int_{0}^{R} r^2 g(r,
1262 gezelter 4419 \cos\theta) \left|\nabla \mathbf{E}^\circ \right|^2 dr d(\cos\theta) }{\int_V\left|\mathbf{E}^\circ\right|^2 d\mathbf{r}}
1263 gezelter 4401 \label{eq:geometricalFactor}
1264     \end{eqnarray}
1265     where $g(r,\cos\theta)$ is a distribution function for the quadrupoles
1266     with respect to an origin at midpoint of a line joining the two probe
1267     charges.
1268    
1269     The effective screening factor is plotted against ion separation
1270     distance in Fig. \ref{fig:screenQuad}. The screening evaluated from
1271     the perturbation and fluctuation methods are shown in right and
1272     central panels. Here the susceptibility is calculated from the
1273     simulation and the geometrical factor is evaluated analytically, using
1274     the field and field-gradient produced by ions. The right hand panel
1275     shows the screening factor obtained from the PMF calculations.
1276    
1277     We note that the screening factor obtained from both the perturbation
1278     and fluctuation formula show good agreement with the screening factor
1279     calculated using PMF method. As there are no large differences in
1280     quadrupole-quadruople interactions for various real-space
1281     methods,\cite{PaperI,PaperII} we generally good agreement for the
1282     screening factors using any of the real space methods.
1283    
1284 gezelter 4404 \section{Conclusions}
1285     We have used both perturbation and fluctuation approaches to evaluate
1286     dielectric properties for dipolar and quadrupolar fluids. The static
1287     dielectric constant is the relevant bulk property for dipolar fluids,
1288     while the quadrupolar susceptibility plays a similar role for
1289     quadrupoles. Corrections to both the static dielectric constant and
1290     the quadrupolar susceptibility were derived for three new real space
1291     electrostatic methods, and these corrections were tested against a
1292     third measure of dielectric screening, the potential of mean force
1293     between two ions immersed in the fluids.
1294 gezelter 4401
1295 gezelter 4404 For the dipolar fluids, we find that the polarizability evaluated
1296     using the perturbation and fluctuation methods show excellent
1297     agreement, indicating that equilibrium calculations of the dipole
1298     fluctuations are good measures of bulk polarizability. One of the
1299     findings of the second paper in this series is that the moderately
1300     damped GSF and SP methods were most suitable for molecular dynamics
1301     and Monte Carlo simulations, respectively.\cite{PaperII}
1302 mlamichh 4353
1303 gezelter 4404 The dielectric constant evaluated using the computed polarizability
1304     and correction factors agrees well with the previous Ewald-based
1305     simulation results \cite{Adams81,NeumannI83} for moderate damping
1306 gezelter 4408 parameters in the range 0.25--0.3~\AA~$^{-1}$.
1307 mlamichh 4353
1308 gezelter 4404 Although the TSF method alters many dynamic and structural features in
1309     multipolar liquids,\cite{PaperII} it is surprisingly good at computing
1310     bulk dielectric properties at nearly all ranges of the damping
1311     parameter. In fact, the correction factor, $A = 1$, for the TSF
1312     method so the conducting boundary formula is essentially correct when
1313     using this method for point dipolar fluids.
1314 mlamichh 4397
1315 gezelter 4404 The dielectric correction formula (equation ~\ref{correctionFormula})
1316     is extremely sensitive when the correction factor ($A$) deviates from
1317     1, and this behavior is also seen in the effective screening of ions
1318     embedded in the fluid.
1319 mlamichh 4353
1320 gezelter 4404 As in the dipolar case, the quadpole polarizability evaluated from
1321     both perturbation and fluctuation simulations show excellent
1322     agreement, again confirming that equilibrium fluctuation calculations
1323     are sufficient to reproduce bulk dielectric properties in these
1324     fluids. The quadrupolar susceptibility calculated via our derived
1325     correction factors tends to produce the same result for all three real
1326     space methods. Similarly, the screening factor calculated using the
1327     susceptibility and a weighted geometric factor provides good agreement
1328     with results obtained directly via potentials of mean force. For
1329     quadrupolar fluids, the distance dependence of the electrostatic
1330     interaction is significantly reduced and the correction factors are
1331     all small. These points suggest that how an electrostatic method
1332     treats the cutoff radius become less consequential for higher order
1333     multipoles.
1334    
1335     For this reason, we renew our recommendation that the
1336     moderately-damped GSF method is an excellent choice for molecular
1337     dynamics simulation where point-multipole interactions are being
1338     utilized to compute bulk properties of fluids.
1339 mlamichh 4353
1340 mlamichh 4412 \appendix
1341 gezelter 4413 \section{Contraction of the quadrupolar tensor with the traceless
1342     quadrupole moment }
1343 mlamichh 4412 \label{ap:quadContraction}
1344 gezelter 4413 For quadrupolar liquids modeled using point quadrupoles, the
1345     interaction tensor is shown in Eq. (\ref{quadRadial}). The Fourier
1346     transformation of this tensor for $ \mathbf{k} = 0$ is,
1347 mlamichh 4412 \begin{equation}
1348     \tilde{T}_{\alpha\beta\gamma\delta}(0) = \int_V T_{\alpha\beta\gamma\delta}(\mathbf{r}) d \mathbf{r}
1349     \end{equation}
1350 gezelter 4413 On the basis of symmetry, the 81 elements can be placed in four
1351     different groups: $\tilde{T}_{aaaa}$, $\tilde{T}_{aaab}$,
1352     $\tilde{T}_{aabb}$, and $\tilde{T}_{aabc}$, where $a$, $b$, and $c$,
1353 gezelter 4414 and can take on distinct values from the set $\left\{x, y, z\right\}$.
1354     The elements belonging to each of these groups can be obtained using
1355     permutations of the indices. Integration of all of the elements shows
1356     that only the groups with indices ${aaaa}$ and ${aabb}$ are non-zero.
1357 gezelter 4413
1358     We can derive values of the components of $\tilde{T}_{aaaa}$ and
1359     $\tilde{T}_{aabb}$ as follows;
1360 mlamichh 4412 \begin{eqnarray}
1361 gezelter 4413 \tilde{T}_{xxxx}(0) &=&
1362 mlamichh 4412 \int_{\textrm{V}}
1363 gezelter 4414 \left[ 3v_{41}(R)+6x^2v_{42}(r)/r^2 + x^4\,v_{43}(r)/r^4 \right] d\mathbf{r} \nonumber \\
1364 mlamichh 4412 &=&12\pi \int_0^{r_c}
1365 gezelter 4414 \left[ v_{41}(r)+\frac{2}{3} v_{42}(r) + \frac{1}{15}v_{43}(r) \right] r^2\,dr =
1366 gezelter 4413 \mathrm{12 \pi B}
1367 mlamichh 4412 \end{eqnarray}
1368     and
1369     \begin{eqnarray}
1370 gezelter 4413 \tilde{T}_{xxyy}(0)&=&
1371     \int_{\textrm{V}}
1372 gezelter 4414 \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 \\
1373 gezelter 4413 &=&4\pi \int_0^{r_c}
1374 gezelter 4414 \left[ v_{41}(r)+\frac{2}{3} v_{42}(r) + \frac{1}{15}v_{43}(r) \right] r^2\,dr =
1375 gezelter 4413 \mathrm{4 \pi B}.
1376 mlamichh 4412 \end{eqnarray}
1377 gezelter 4413 These integrals yield the same values for all permutations of the
1378     indices in both tensor element groups. In equation
1379     \ref{fourierQuadZeroK}, for a particular value of the quadrupolar
1380     polarization $\tilde{\Theta}_{aa}$ we can contract
1381     $\tilde{T}_{aa\gamma\delta}(0)$ with $\tilde{\Theta}_{\gamma\delta}$,
1382     using the traceless properties of the quadrupolar moment,
1383 mlamichh 4412 \begin{eqnarray}
1384 gezelter 4413 \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 \\
1385     &=& 12 \pi \mathrm{B}\tilde{\Theta}_{xx}(0) +
1386     4 \pi \mathrm{B}\tilde{\Theta}_{yy}(0) +
1387     4 \pi \mathrm{B}\tilde{\Theta}_{zz}(0) \nonumber \\
1388     &=& 8 \pi \mathrm{B}\tilde{\Theta}_{xx}(0) + 4 \pi
1389     \mathrm{B}\left(\tilde{\Theta}_{xx}(0)+\tilde{\Theta}_{yy}(0) +
1390     \tilde{\Theta}_{zz}(0)\right) \nonumber \\
1391     &=& 8 \pi \mathrm{B}\tilde{\Theta}_{xx}(0)
1392 mlamichh 4412 \end{eqnarray}
1393 gezelter 4413 Similarly for a quadrupolar polarization $\tilde{\Theta}_{xy}$ in
1394     equation \ref{fourierQuadZeroK}, we can contract
1395     $\tilde{T}_{xy\gamma\delta}(0)$ with $\tilde{\Theta}_{\gamma\delta}$,
1396     using the only surviving terms of the tensor,
1397 mlamichh 4412 \begin{eqnarray}
1398 gezelter 4413 \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 \\
1399     &=& 4 \pi \mathrm{B}\tilde{\Theta}_{xy}(0) +
1400     4 \pi \mathrm{B}\tilde{\Theta}_{yx}(0) \nonumber \\
1401     &=& 8 \pi \mathrm{B}\tilde{\Theta}_{xy}(0)
1402 mlamichh 4412 \end{eqnarray}
1403 gezelter 4413 Here, we have used the symmetry of the quadrupole tensor to combine
1404     the symmetric terms. Therefore we can write matrix contraction for
1405     $\tilde{T}_{\alpha\beta\gamma\delta}(\mathrm{0})$ and
1406     $ \tilde{\Theta}_{\gamma\delta}(\mathrm{0})$ in a general form,
1407 mlamichh 4412 \begin{equation}
1408 gezelter 4413 \tilde{T}_{\alpha\beta\gamma\delta}(\mathrm{0})\tilde{\Theta}_{\gamma\delta}(\mathrm{0})
1409     = 8 \pi \mathrm{B} \tilde{\Theta}_{\alpha\beta}(\mathrm{0}),
1410 mlamichh 4412 \label{contract}
1411     \end{equation}
1412 gezelter 4413 which is the same as Eq. (\ref{quadContraction}).
1413 mlamichh 4412
1414 gezelter 4413 When the molecular quadrupoles are represented by point charges, the
1415     symmetry of the quadrupolar tensor is same as for point quadrupoles
1416     (see Eqs.~\ref{quadCharge} and \ref{quadRadial}). However, for
1417     molecular quadrupoles represented by point dipoles, the symmetry of
1418     the quadrupolar tensor must be handled separately (compare
1419     equations~\ref{quadDip} and~\ref{quadRadial}). Although there is a
1420 gezelter 4414 difference in symmetry, the final result (Eq.~\ref{contract}) also holds
1421     true for dipolar representations.
1422 gezelter 4413
1423 gezelter 4414 \section{Quadrupolar correction factor for the Ewald-Kornfeld (EK)
1424     method}
1425     The interaction tensor between two point quadrupoles in the Ewald
1426     method may be expressed,\cite{Smith98,NeumannII83}
1427     \begin{align}
1428     {T}_{\alpha\beta\gamma\delta}(\mathbf{r}) = &\frac{4\pi}{V
1429     }\sum_{k\neq0}^{\infty}
1430     e^{-k^2 / 4
1431     \kappa^2} e^{-i\mathbf{k}\cdot
1432     \mathbf{r}} \left(\frac{r_\alpha r_\beta k_\delta k_\gamma}{k^2}\right) \nonumber \\
1433     &+ \left(\delta_{\alpha\beta}\delta_{\gamma\delta}+\delta_{\alpha\gamma}\delta_{\beta\delta}+\delta_{\alpha\delta}\delta_{\beta\gamma}\right)
1434     B_2(r) \nonumber \\
1435     &- \left(\delta_{\gamma\delta} r_\alpha r_\beta + \mathrm{ 5\; permutations}\right) B_3(r) \nonumber \\
1436     &+ \left(r_\alpha r_\beta r_\gamma r_\delta \right) B_4(r)
1437 mlamichh 4412 \label{ewaldTensor}
1438 gezelter 4414 \end{align}
1439     where $B_n(r)$ are radial functions defined in reference
1440     \onlinecite{Smith98},
1441     \begin{align}
1442     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) \\
1443     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) \\
1444     B_4(r) =& \frac{105}{r^9}\left(\frac{2r\kappa e^{-r^2
1445     \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}}
1446     + \frac{16r^7\kappa^7 e^{-r^2 \kappa^2}}{105\sqrt{\pi}} + \mathrm{erfc(\kappa r)} \right)
1447     \end{align}
1448    
1449     We can divide ${T}_{\alpha\beta\gamma\delta}(\mathbf{r})$ into three
1450     parts:
1451 mlamichh 4412 \begin{eqnarray}
1452 gezelter 4414 & & \mathbf{T}(\mathbf{r}) =
1453     \mathbf{T}^\mathrm{K}(\mathbf{r}) +
1454     \mathbf{T}^\mathrm{R1}(\mathbf{r}) +
1455     \mathbf{T}^\mathrm{R2}(\mathbf{r})
1456 mlamichh 4412 \end{eqnarray}
1457 gezelter 4414 where the first term is the reciprocal space portion. Since the
1458     quadrupolar correction factor $B = \tilde{T}_{abab}(0) / 4\pi$ and
1459 gezelter 4415 $\mathbf{k} = 0 $ is excluded from the reciprocal space sum,
1460 gezelter 4414 $\mathbf{T}^\mathrm{K}$ will not contribute.\cite{NeumannII83} The
1461     remaining terms,
1462     \begin{equation}
1463     \mathbf{T}^\mathrm{R1}(\mathbf{r}) = \mathbf{T}^\mathrm{bare}(\mathbf{r}) \left(\frac{2r\kappa e^{-r^2
1464     \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}}
1465     + \frac{16r^7\kappa^7 e^{-r^2 \kappa^2}}{105\sqrt{\pi}} + \mathrm{erfc(\kappa r)} \right)
1466     \end{equation}
1467 mlamichh 4412 and
1468     \begin{eqnarray}
1469 gezelter 4414 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 \\
1470     &-&\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)
1471 mlamichh 4412 \end{eqnarray}
1472 gezelter 4414 are contributions from the real space
1473     sum.\cite{Adams76,Adams80,Adams81} Here
1474     $\mathbf{T}^\mathrm{bare}(\mathbf{r})$ is the unmodified quadrupolar
1475     tensor (for undamped quadrupoles). Due to the angular symmetry of the
1476     unmodified tensor, the integral of
1477     $\mathbf{T}^\mathrm{R1}(\mathbf{r})$ will vanish when integrated over
1478     a spherical region. The only term contributing to the correction
1479     factor (B) is therefore
1480     $T^\mathrm{R2}_{\alpha\beta\gamma\delta}(\mathbf{r})$, which allows us
1481     to derive the correction factor for the Ewald-Kornfeld (EK) method,
1482 mlamichh 4412 \begin{eqnarray}
1483 gezelter 4414 \mathrm{B} &=& \frac{1}{4\pi} \int_V T^\mathrm{R2}_{abab}(\mathbf{r}) \nonumber \\
1484     &=& -\frac{8r_c^3 \kappa^5 e^{-\kappa^2 r_c^2}}{15\sqrt{\pi}}
1485 mlamichh 4412 \end{eqnarray}
1486 gezelter 4414 which is essentially identical with the correction factor from the
1487     direct spherical cutoff (SC) method.
1488 gezelter 4428
1489     \begin{acknowledgments}
1490     Support for this project was provided by the National Science Foundation
1491     under grant CHE-1362211. Computational time was provided by the
1492     Center for Research Computing (CRC) at the University of Notre Dame.
1493     \end{acknowledgments}
1494    
1495 mlamichh 4353 \newpage
1496 mlamichh 4397 \bibliography{dielectric_new}
1497 mlamichh 4353
1498     \end{document}