ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/multipole/dielectric_new.tex
Revision: 4439
Committed: Fri Aug 12 16:57:59 2016 UTC (9 years, 1 month ago) by gezelter
Content type: application/x-tex
File size: 85859 byte(s)
Log Message:
Latest

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 gezelter 4431 electrostatic method being utilized. One of the most important
55 gezelter 4399 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 gezelter 4439 dielectric constants 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 gezelter 4431 macroscopic polarizability obtained from a simulation depends on the
113 gezelter 4417 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 gezelter 4433 correction factors for the three new real space methods. The
139     corrections are modifications to fluctuation expressions to account
140 gezelter 4434 for truncation, shifting, and damping of the field and field gradient
141     contributions from other multipoles. We find that the correction
142     formulae for dipolar molecules depends not only on the methodology
143     being used, but also on whether the molecular dipoles are treated
144     using point charges or point dipoles. We derive correction factors for
145     both cases.
146 gezelter 4399
147 gezelter 4392 In quadrupolar fluids, the relationship between quadrupolar
148 gezelter 4434 susceptibility and dielectric screening is not as straightforward as
149     in the dipolar case. The effective dielectric constant depends on the
150     geometry of the external (or internal) field
151 gezelter 4399 perturbation.\cite{Ernst92} Significant efforts have been made to
152     increase our understanding the dielectric properties of these
153 gezelter 4400 fluids,\cite{JeonI03,JeonII03,Chitanvis96} although a general
154     correction formula has not yet been developed.
155 gezelter 4390
156 gezelter 4392 In this paper we derive general formulae for calculating the
157 gezelter 4400 quadrupolar susceptibility of quadrupolar fluids. We also evaluate the
158     correction factor for SP, GSF, and TSF methods for quadrupolar fluids
159     interacting via point charges, point dipoles or directly through
160     quadrupole-quadrupole interactions.
161 gezelter 4390
162 gezelter 4400 We also calculate the screening behavior for two ions immersed in
163     multipolar fluids to estimate the distance dependence of charge
164     screening in both dipolar and quadrupolar fluids. We use three
165     distinct methods to compare our analytical results with computer
166 gezelter 4419 simulations (see Fig. \ref{fig:schematic}):
167 gezelter 4390 \begin{enumerate}
168 gezelter 4396 \item responses of the fluid to external perturbations,
169 gezelter 4399 \item fluctuations of system multipole moments, and
170 gezelter 4396 \item potentials of mean force between solvated ions,
171 gezelter 4390 \end{enumerate}
172    
173 gezelter 4408 \begin{figure}
174 gezelter 4428 \includegraphics[width=\linewidth]{Schematic.eps}
175 gezelter 4408 \caption{Dielectric properties of a fluid measure the response to
176 gezelter 4411 external electric fields and gradients (left), or internal fields
177     and gradients generated by the molecules themselves (center), or
178     fields produced by embedded ions (right). The dielectric constant
179     ($\epsilon$) measures all three responses in dipolar fluids (top).
180     In quadrupolar liquids (bottom), the relevant bulk property is the
181     quadrupolar susceptibility ($\chi_Q$), and the geometry of the field
182 gezelter 4408 determines the effective dielectric screening.}
183     \label{fig:schematic}
184     \end{figure}
185    
186 gezelter 4400 Under the influence of weak external fields, the bulk polarization of
187     the system is primarily a linear response to the perturbation, where
188 gezelter 4439 the proportionality constant depends on the electrostatic interactions
189 gezelter 4400 between the multipoles. The fluctuation formulae connect bulk
190     properties of the fluid to equilibrium fluctuations in the system
191     multipolar moments during a simulation. These fluctuations also depend
192     on the form of the electrostatic interactions between molecules.
193     Therefore, the connections between the actual bulk properties and both
194     the computed fluctuation and external field responses must be modified
195     accordingly.
196 gezelter 4390
197 gezelter 4400 The potential of mean force (PMF) allows calculation of an effective
198 gezelter 4399 dielectric constant or screening factor from the potential energy
199     between ions before and after dielectric material is introduced.
200 gezelter 4400 Computing the PMF between embedded point charges is an additional
201     check on the bulk properties computed via the other two methods.
202 mlamichh 4363
203 gezelter 4400 \section{The Real-Space Methods}
204    
205 gezelter 4404 In the first paper in this series, we derived interaction energies, as
206     well as expressions for the forces and torques for point multipoles
207     interacting via three new real-space methods.\cite{PaperI} The Taylor
208     shifted-force (TSF) method modifies the electrostatic kernel,
209     $f(r) = 1/r$, so that all forces and torques go smoothly to zero at
210     the cutoff radius,
211     \begin{equation}
212     U^{\text{TSF}}_{ab} = M_{a} M_{b} f_n(r).
213     \label{TSF}
214     \end{equation}
215     Here the multipole operator for site $a$, $M_{a}$, is expressed in
216     terms of the point charge, $C_{a}$, dipole, ${\bf D}_{a}$, and
217     quadrupole, $\mathsf{Q}_{a}$, for object $a$, etc. Because each of
218 gezelter 4432 the multipole operators includes gradient operators (one for a dipole,
219     two for a quadrupole, \textit{etc.}), an approximate electrostatic
220     kernel, $f_n(r)$ is Taylor-expanded around the cutoff radius, so that
221     $n + 1$ derivatives vanish as $r \rightarrow r_c$. This ensures
222     smooth convergence of the energy, forces, and torques as molecules
223     leave and reenter each others cutoff spheres. The order of the Taylor
224     expansion is determined by the multipolar order of the interaction.
225     That is, smooth quadrupole-quadrupole forces require the fifth
226     derivative to vanish at the cutoff radius, so the appropriate function
227     Taylor expansion will be of fifth order.
228 gezelter 4404
229     Following this procedure results in separate radial functions for each
230     of the distinct orientational contributions to the potential. For
231     example, in dipole-dipole interactions, the direct dipole dot product
232     ($\mathbf{D}_{a} \cdot \mathbf{D}_{b}$) is treated differently than
233     the dipole-distance dot products:
234     \begin{equation}
235     U_{\mathbf{D}_{a}\mathbf{D}_{b}}(r)= -\frac{1}{4\pi \epsilon_0} \left[ \left(
236     \mathbf{D}_{a} \cdot
237     \mathbf{D}_{b} \right) v_{21}(r) +
238     \left( \mathbf{D}_{a} \cdot \hat{\mathbf{r}} \right)
239     \left( \mathbf{D}_{b} \cdot \hat{\mathbf{r}} \right) v_{22}(r) \right]
240     \end{equation}
241 gezelter 4419 In standard electrostatics, the two radial functions, $v_{21}(r)$ and
242 gezelter 4432 $v_{22}(r)$, are proportional to $1/r^3$, but they have distinct radial
243 gezelter 4419 dependence in the TSF method. Careful choice of these functions makes
244     the forces and torques vanish smoothly as the molecules drift beyond
245     the cutoff radius (even when those molecules are in different
246     orientations).
247 gezelter 4404
248     A second and somewhat simpler approach involves shifting the gradient
249     of the Coulomb potential for each particular multipole order,
250     \begin{equation}
251     U^{\text{GSF}}_{ab} = \sum \left[ U(\mathbf{r}, \mathsf{A}, \mathsf{B}) -
252     U(r_c \hat{\mathbf{r}},\mathsf{A}, \mathsf{B}) - (r-r_c)
253     \hat{\mathbf{r}} \cdot \nabla U(r_c \hat{\mathbf{r}},\mathsf{A},
254     \mathsf{B}) \right]
255     \label{generic2}
256     \end{equation}
257     where the sum describes a separate force-shifting that is applied to
258     each orientational contribution to the energy, i.e. $v_{21}$ and
259     $v_{22}$ are shifted separately. In this expression,
260     $\hat{\mathbf{r}}$ is the unit vector connecting the two multipoles
261     ($a$ and $b$) in space, and $\mathsf{A}$ and $\mathsf{B}$ represent
262 gezelter 4439 the orientations of the multipoles. Because this procedure is equivalent
263 gezelter 4404 to using the gradient of an image multipole placed at the cutoff
264     sphere for shifting the force, this method is called the gradient
265     shifted-force (GSF) approach.
266    
267     Both the TSF and GSF approaches can be thought of as multipolar
268     extensions of the original damped shifted-force (DSF) approach that
269     was developed for point charges. There is also a multipolar extension
270     of the Wolf sum that is obtained by projecting an image multipole onto
271     the surface of the cutoff sphere, and including the interactions with
272     the central multipole and the image. This effectively shifts only the
273     total potential to zero at the cutoff radius,
274     \begin{equation}
275     U^{\text{SP}}_{ab} = \sum \left[ U(\mathbf{r}, \mathsf{A}, \mathsf{B}) -
276     U(r_c \hat{\mathbf{r}},\mathsf{A}, \mathsf{B}) \right]
277     \label{eq:SP}
278     \end{equation}
279     where the sum again describes separate potential shifting that is done
280     for each orientational contribution to the energy. The potential
281     energy between a central multipole and other multipolar sites goes
282     smoothly to zero as $r \rightarrow r_c$, but the forces and torques
283     obtained from this shifted potential (SP) approach are discontinuous
284     at $r_c$.
285    
286     All three of the new real space methods share a common structure: the
287     various orientational contributions to multipolar interaction energies
288     require separate treatment of their radial functions, and these are
289     tabulated for both the raw Coulombic kernel ($1/r$) as well as the
290     damped kernel ($\mathrm{erfc}(\alpha r)/r$), in the first paper of this
291     series.\cite{PaperI} The second paper in this series evaluated the
292     fidelity with which the three new methods reproduced Ewald-based
293     results for a number of model systems.\cite{PaperII} One of the major
294     findings was that moderately-damped GSF simulations produced nearly
295     identical behavior with Ewald-based simulations, but the real-space
296     methods scale linearly with system size.
297    
298 gezelter 4399 \section{Dipolar Fluids and the Dielectric Constant}
299 mlamichh 4353
300 gezelter 4399 Dielectric properties of a fluid arise mainly from responses of the
301     fluid to either applied fields or transient fields internal to the
302     fluid. In response to an applied field, the molecules have electronic
303 gezelter 4432 polarizabilities, changes to internal bond lengths and angles, and
304     reorientations towards the direction of the applied field. There is an
305     added complication that in the presence of external field, the
306     perturbation experienced by any single molecule is not only due to
307 gezelter 4439 the external field but also to the fields produced by the all other
308 gezelter 4432 molecules in the system.
309 mlamichh 4353
310 gezelter 4399 \subsection{Response to External Perturbations}
311    
312     In the presence of uniform electric field $\mathbf{E}$, an individual
313     molecule with a permanent dipole moment $p_o$ will realign along the
314     direction of the field with an average polarization given by
315 mlamichh 4353 \begin{equation}
316 gezelter 4400 \braket{\mathbf{p}} = \epsilon_0 \alpha_p \mathbf{E},
317 mlamichh 4353 \end{equation}
318 gezelter 4400 where $\alpha_p = {p_o}^2 / 3 \epsilon_0 k_B T$ is the contribution to
319     molecular polarizability due solely to reorientation dynamics.
320     Because the applied field must overcome thermal motion, the
321     orientational polarization depends inversely on the temperature.
322 mlamichh 4353
323 gezelter 4432 A condensed phase system of permanent dipoles will also polarize along
324     the direction of an applied field. The polarization density of the
325     system is
326 mlamichh 4353 \begin{equation}
327 gezelter 4432 \textbf{P} = \epsilon_o \chi_{D} \mathbf{E},
328 mlamichh 4353 \label{pertDipole}
329     \end{equation}
330 gezelter 4432 where the constant $\chi_D$ is the dipole susceptibility, which is an
331     emergent property of the dipolar fluid, and is the quantity most
332     directly related to the static dielectric constant,
333     $\epsilon = 1 + \chi_D$.
334 mlamichh 4363
335 gezelter 4399 \subsection{Fluctuation Formula}
336 gezelter 4400 % Using the Ewald sum under tin-foil boundary conditions, $\epsilon$ can
337     % be calculated for systems of orientationally-polarized molecules,
338     % \begin{equation}
339     % \epsilon = 1 + \frac{\langle M^2\rangle}{3\epsilon_0k_\textrm{B}TV},
340     % \label{eq:staticDielectric}
341     % \end{equation}
342     % where $\epsilon_0$ is the permittivity of free space and
343     % $\langle M^2\rangle$ is the fluctuation of the system dipole
344     % moment.\cite{Allen89} The numerator in the fractional term in
345     % equation (\ref{eq:staticDielectric}) is identical to the quantity
346     % calculated in the finite-system Kirkwood $g$ factor ($G_k$):
347     % \begin{equation}
348     % G_k = \frac{\langle M^2\rangle}{N\mu^2},
349     % \label{eq:KirkwoodFactor}
350     % \end{equation}
351     % where $\mu$ is the dipole moment of a single molecule of the
352     % homogeneous system.\cite{NeumannI83,NeumannII83,Neumann84,Neumann85} The
353     % fluctuation term in both equation (\ref{eq:staticDielectric}) and
354     % (\ref{eq:KirkwoodFactor}) is calculated as follows,
355     % \begin{equation}
356     % \begin{split}
357     % \langle M^2\rangle &= \langle\bm{M}\cdot\bm{M}\rangle
358     % - \langle\bm{M}\rangle\cdot\langle\bm{M}\rangle \\
359     % &= \langle M_x^2+M_y^2+M_z^2\rangle
360     % - (\langle M_x\rangle^2 + \langle M_x\rangle^2
361     % + \langle M_x\rangle^2).
362     % \end{split}
363     % \label{eq:fluctBoxDipole}
364     % \end{equation}
365     % This fluctuation term can be accumulated during a simulation; however,
366     % it converges rather slowly, thus requiring multi-nanosecond simulation
367     % times.\cite{Horn04} In the case of tin-foil boundary conditions, the
368     % dielectric/surface term of the Ewald summation is equal to zero.
369 mlamichh 4353
370 gezelter 4400 For a system of dipolar molecules at thermal equilibrium, we can
371     define both a system dipole moment, $\mathbf{M} = \sum_i \mathbf{p}_i$
372     as well as a dipole polarization density,
373 gezelter 4432 $\mathbf{P} = \braket{\mathbf{M}}/V$. The polarization density can be
374     expressed approximately in terms of fluctuations in the net dipole
375 gezelter 4400 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 gezelter 4432 When no applied electric field is present, the ensemble average of
392     both the net dipole moment $\braket{\mathbf{M}}$ and dipolar
393     polarization $\bf{P}$ tends to vanish but $\braket{\mathbf{M}^2}$ does
394     not. The bulk dipole polarizability can therefore be
395     written
396 gezelter 4401 \begin{equation}
397 gezelter 4432 \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 4432 The susceptibility ($\chi_D$) and bulk polarizability ($\alpha_D$)
400 gezelter 4434 both measure responses of a dipolar system. However, $\chi_D$ is the
401     bulk property assuming an infinite system and exact treatment of
402     electrostatic interactions, while $\alpha_D$ is relatively simple to
403     compute from numerical simulations. One of the primary aims of this
404     paper is to provide the connection between the bulk properties
405 gezelter 4432 ($\epsilon, \chi_D$) and the computed quantities ($\alpha_D$) that
406     have been adapted for the new real-space methods.
407 gezelter 4400
408 gezelter 4399 \subsection{Correction Factors}
409 gezelter 4408 \label{sec:corrFactor}
410 gezelter 4400 In the presence of a uniform external field $ \mathbf{E}^\circ$, the
411     total electric field at $\mathbf{r}$ depends on the polarization
412     density at all other points in the system,\cite{NeumannI83}
413 mlamichh 4353 \begin{equation}
414 gezelter 4400 \mathbf{E}(\mathbf{r}) = \mathbf{E}^\circ(\mathbf{r}) +
415     \frac{1}{4\pi\epsilon_o} \int d\mathbf{r}^\prime
416     \mathbf{T}(\mathbf{r}-\mathbf{r}^\prime)\cdot
417     {\mathbf{P}(\mathbf{r}^\prime)}.
418     \label{eq:localField}
419 mlamichh 4353 \end{equation}
420 gezelter 4400 $\mathbf{T}$ is the dipole interaction tensor connecting dipoles at
421 gezelter 4432 $\mathbf{r}^\prime$ with the point of interest ($\mathbf{r}$), where
422     the integral is done over all space. Because simulations utilize
423 gezelter 4434 periodic boundary conditions or spherical cutoffs, the integral is
424     normally carried out either over the domain ($0 < r < r_c$) or in
425     reciprocal space.
426 mlamichh 4353
427 gezelter 4400 In simulations of dipolar fluids, the molecular dipoles may be
428     represented either by closely-spaced point charges or by
429 gezelter 4408 point dipoles (see Fig. \ref{fig:tensor}).
430 mlamichh 4353 \begin{figure}
431 gezelter 4428 \includegraphics[width=\linewidth]{Tensors.eps}
432 gezelter 4408 \caption{In the real-space electrostatic methods, the molecular dipole
433     tensor, $\mathbf{T}_{\alpha\beta}(r)$, is not the same for
434     charge-charge interactions as for point dipoles (left panel). The
435     same holds true for the molecular quadrupole tensor (right panel),
436     $\mathbf{T}_{\alpha\beta\gamma\delta}(r)$, which can have distinct
437     forms if the molecule is represented by charges, dipoles, or point
438     quadrupoles.}
439     \label{fig:tensor}
440 mlamichh 4353 \end{figure}
441     In the case where point charges are interacting via an electrostatic
442     kernel, $v(r)$, the effective {\it molecular} dipole tensor,
443     $\mathbf{T}$ is obtained from two successive applications of the
444     gradient operator to the electrostatic kernel,
445 mlamichh 4412 \begin{eqnarray}
446 gezelter 4432 T_{\alpha \beta}(\mathbf{r}) &=& \nabla_\alpha \nabla_\beta
447 mlamichh 4412 \left(v(r)\right) \\
448 gezelter 4416 &=& \delta_{\alpha \beta}
449 mlamichh 4353 \left(\frac{1}{r} v^\prime(r) \right) + \frac{r_{\alpha}
450     r_{\beta}}{r^2} \left( v^{\prime \prime}(r) - \frac{1}{r}
451     v^{\prime}(r) \right)
452     \label{dipole-chargeTensor}
453 mlamichh 4412 \end{eqnarray}
454 mlamichh 4353 where $v(r)$ may be either the bare kernel ($1/r$) or one of the
455     modified (Wolf or DSF) kernels. This tensor describes the effective
456 gezelter 4400 interaction between molecular dipoles ($\mathbf{D}$) in Gaussian units
457     as $-\mathbf{D} \cdot \mathbf{T} \cdot \mathbf{D}$.
458 mlamichh 4353
459 gezelter 4400 When utilizing any of the three new real-space methods for point
460     \textit{dipoles}, the tensor is explicitly constructed,
461 mlamichh 4353 \begin{equation}
462 gezelter 4432 T_{\alpha \beta}(\mathbf{r}) = \delta_{\alpha \beta} v_{21}(r) +
463 mlamichh 4353 \frac{r_{\alpha} r_{\beta}}{r^2} v_{22}(r)
464     \label{dipole-diopleTensor}
465     \end{equation}
466     where the functions $v_{21}(r)$ and $v_{22}(r)$ depend on the level of
467 gezelter 4400 the approximation.\cite{PaperI,PaperII} Although the Taylor-shifted
468     (TSF) and gradient-shifted (GSF) models produce to the same $v(r)$
469     function for point charges, they have distinct forms for the
470     dipole-dipole interaction.
471 mlamichh 4353
472 gezelter 4432 Using the constitutive relation in Eq. (\ref{pertDipole}), the
473     polarization density $\mathbf{P}(\mathbf{r})$ is given by,
474 mlamichh 4353 \begin{equation}
475 gezelter 4400 \mathbf{P}(\mathbf{r}) = \epsilon_o \chi_D
476     \left(\mathbf{E}^\circ(\mathbf{r}) + \frac{1}{4\pi\epsilon_o} \int
477     d\mathbf{r}^\prime \mathbf{T}(\mathbf{r}-\mathbf{r}^\prime ) \cdot \mathbf{P}(\mathbf{r}^\prime)\right).
478 mlamichh 4353 \label{constDipole}
479     \end{equation}
480 gezelter 4434 Note that $\chi_D$ depends explicitly on the details of the dipole
481 gezelter 4400 interaction tensor. Neumann \textit{et al.}
482     \cite{NeumannI83,NeumannII83,Neumann84,Neumann85} derived an elegant
483     way to modify the fluctuation formula to correct for approximate
484     interaction tensors. This correction was derived using a Fourier
485     representation of the interaction tensor,
486     $\tilde{\mathbf{T}}(\mathbf{k})$, and involves the quantity,
487 mlamichh 4353 \begin{equation}
488 gezelter 4432 \mathsf{A} = \frac{3}{4\pi}\tilde{\mathbf{T}}(0) = \frac{3}{4\pi} \int_V
489     d\mathbf{r} \mathbf{T}(\mathbf{r})
490 mlamichh 4353 \end{equation}
491 gezelter 4432 which is the $k \rightarrow 0$ limit of
492 gezelter 4439 $\tilde{\mathbf{T}}(\mathbf{k})$. Note that the integration of the
493 gezelter 4432 dipole tensors, Eqs. (\ref{dipole-chargeTensor}) and
494     (\ref{dipole-diopleTensor}), over spherical volumes yields values only
495     along the diagonal. Additionally, the spherical symmetry of
496     $\mathbf{T}(\mathbf{r})$ insures that all diagonal elements are
497     identical. For this reason, $\mathsf{A}$ can be written as a scalar
498     constant ($A$) multiplying the unit tensor.
499    
500     Using the quantity $A$ (originally called $Q$ in
501 gezelter 4400 refs. \onlinecite{NeumannI83,NeumannII83,Neumann84,Neumann85}), the
502     dielectric constant can be computed
503 mlamichh 4353 \begin{equation}
504 gezelter 4400 \epsilon = \frac{3+(A + 2)(\epsilon_{CB}-1)}{3 + (A -1)(\epsilon_{CB}-1)}
505     \label{correctionFormula}
506 mlamichh 4353 \end{equation}
507 gezelter 4400 where $\epsilon_{CB}$ is the widely-used conducting boundary
508     expression for the dielectric constant,
509 mlamichh 4353 \begin{equation}
510 gezelter 4400 \epsilon_{CB} = 1 + \frac{\braket{\bf{M}^2}-{\braket{\bf{M}}}^2}{3
511 gezelter 4432 \epsilon_o V k_B T} = 1 + \alpha_{D}.
512 gezelter 4408 \label{conductingBoundary}
513 mlamichh 4353 \end{equation}
514 gezelter 4432 Eqs. (\ref{correctionFormula}) and (\ref{conductingBoundary}) allow
515     estimation of the static dielectric constant from fluctuations
516     computed directly from simulations, with the understanding that
517     Eq. (\ref{correctionFormula}) is extraordinarily sensitive when $A$ is
518     far from unity.
519 gezelter 4400
520     % 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}
521     % \begin{equation}
522     % \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}
523     % \label{singular}
524     % \end{equation}
525     % Substituting equation (\ref{singular}) in the equation (\ref{constDipole}) we get,
526     % \begin{equation}
527     % \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).
528     % \end{equation}
529     % For both polarization and electric field homogeneous, this can be easily solved using Fourier transformation,
530     % \begin{equation}
531     % \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}).
532     % \end{equation}
533     % For homogeneous applied field Fourier component is non-zero only if $\kappa = 0$. Therefore,
534     % \begin{equation}
535     % \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}).
536     % \label{fourierDipole}
537     % \end{equation}
538     % 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,
539     % \begin{equation}
540     % \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}
541     % \end{equation}
542     % 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,
543     % \begin{equation}
544     % \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}
545     % \label{correctionFormula}
546     % \end{equation}
547    
548 gezelter 4404 We have utilized the Neumann \textit{et al.} approach for the three
549 gezelter 4400 new real-space methods, and obtain method-dependent correction
550     factors. The expression for the correction factor also depends on
551     whether the simulation involves point charges or point dipoles to
552 gezelter 4425 represent the molecular dipoles. These corrections factors are listed
553     in Table \ref{tab:A}. We note that the GSF correction factor for
554     point dipoles has been independently derived by Stenqvist \textit{et
555     al.}\cite{Stenqvist:2015ph}
556 mlamichh 4353 \begin{table}
557 gezelter 4400 \caption{Expressions for the dipolar correction factor ($A$) for the
558     real-space electrostatic methods in terms of the damping parameter
559 mlamichh 4353 ($\alpha$) and the cutoff radius ($r_c$). The Ewald-Kornfeld result
560 gezelter 4400 derived in Refs. \onlinecite{Adams80,Adams81,NeumannI83} is shown
561     for comparison using the Ewald convergence parameter ($\kappa$)
562     and the real-space cutoff value ($r_c$). }
563 mlamichh 4353 \label{tab:A}
564 gezelter 4416 \begin{tabular}{l|c|c}
565 gezelter 4408 \toprule
566 gezelter 4416 & \multicolumn{2}{c}{Molecular Representation} \\
567 gezelter 4400 Method & point charges & point dipoles \\
568 gezelter 4408 \colrule
569 gezelter 4431 Shifted Potential (SP) & $ \mathrm{erf}(r_c \alpha) - \frac{2
570 gezelter 4411 \alpha r_c}{\sqrt{\pi}} e^{-\alpha^2
571     r_c^2}$ & $\mathrm{erf}(r_c \alpha)
572     -\frac{2 \alpha
573     r_c}{\sqrt{\pi}}\left(1+\frac{2\alpha^2
574     {r_c}^2}{3}
575     \right)e^{-\alpha^2{r_c}^2}
576     $\\ \colrule
577     Gradient-shifted (GSF) & 1 & $\mathrm{erf}(\alpha r_c)-\frac{2
578     \alpha r_c}{\sqrt{\pi}} \left(1 +
579     \frac{2 \alpha^2 r_c^2}{3} +
580     \frac{\alpha^4
581     r_c^4}{3}\right)e^{-\alpha^2 r_c^2} $ \\ \colrule
582 gezelter 4416 Taylor-shifted (TSF) &\multicolumn{2}{c}{1} \\ \colrule
583 gezelter 4432 %Spherical Cutoff (SC)& \multicolumn{2}{c}{$\mathrm{erf}(r_c \alpha) -
584     % \frac{2 \alpha r_c}{\sqrt{\pi}} e^{-\alpha^2
585     % r_c^2}$} \\
586 gezelter 4416 Ewald-Kornfeld (EK) & \multicolumn{2}{c}{$\mathrm{erf}(r_c \kappa) -
587 gezelter 4408 \frac{2 \kappa r_c}{\sqrt{\pi}} e^{-\kappa^2
588     r_c^2}$} \\
589     \botrule
590     \end{tabular}
591 mlamichh 4353 \end{table}
592 gezelter 4400 Note that for point charges, the GSF and TSF methods produce estimates
593     of the dielectric that need no correction, and the TSF method likewise
594     needs no correction for point dipoles.
595 gezelter 4399
596     \section{Quadrupolar Fluids and the Quadrupolar Susceptibility}
597     \subsection{Response to External Perturbations}
598    
599 gezelter 4400 A molecule with a permanent quadrupole, $\mathsf{q}$, will align in
600     the presence of an electric field gradient $\nabla\mathbf{E}$. The
601     anisotropic polarization of the quadrupole is given
602     by,\cite{AduGyamfi78,AduGyamfi81}
603 gezelter 4399 \begin{equation}
604 gezelter 4400 \braket{\mathsf{q}} - \frac{\mathbf{I}}{3}
605     \mathrm{Tr}(\mathsf{q}) = \epsilon_o \alpha_q \nabla\mathbf{E},
606 gezelter 4399 \end{equation}
607 gezelter 4400 where $\alpha_q = q_o^2 / 15 \epsilon_o k_B T $ is a molecular quadrupole
608 gezelter 4431 polarizability and $q_o$ is an effective quadrupole moment for the molecule,
609 gezelter 4400 \begin{equation}
610     q_o^2 = 3 \mathsf{q}:\mathsf{q} - \mathrm{Tr}(\mathsf{q})^2.
611     \end{equation}
612 gezelter 4432 Note that quadrupole calculations involve tensor contractions (double
613     dot products) between rank two tensors, which are defined as
614     \begin{equation}
615     \mathsf{A} \colon \mathsf{B} = \sum_\alpha \sum_\beta
616     A_{\alpha \beta} B_{\beta \alpha}.
617     \end{equation}
618 gezelter 4399
619 gezelter 4400 In the presence of an external field gradient, a system of quadrupolar
620     molecules also organizes with an anisotropic polarization,
621 gezelter 4399 \begin{equation}
622 gezelter 4400 \mathsf{Q} - \frac{\mathbf{I}}{3} \mathrm{Tr}(\mathsf{Q}) = \epsilon_o
623 gezelter 4432 \chi_Q \nabla\mathbf{E}
624 gezelter 4400 \end{equation}
625     where $\mathsf{Q}$ is the traced quadrupole density of the system and
626 gezelter 4432 $\chi_Q$ is a macroscopic quadrupole susceptibility which has
627 gezelter 4400 dimensions of $\mathrm{length}^{-2}$. Equivalently, the traceless form
628     may be used,
629     \begin{equation}
630 gezelter 4432 \mathsf{\Theta} = 3 \epsilon_o \chi_Q \nabla\mathbf{E},
631 gezelter 4399 \label{pertQuad}
632     \end{equation}
633 gezelter 4400 where
634     $\mathsf{\Theta} = 3\mathsf{Q} - \mathbf{I} \mathrm{Tr}(\mathsf{Q})$
635     is the traceless tensor that also describes the system quadrupole
636     density. It is this tensor that will be utilized to derive correction
637     factors below.
638    
639 gezelter 4399 \subsection{Fluctuation Formula}
640 gezelter 4400 As in the dipolar case, we may define a system quadrupole moment,
641     $\mathsf{M}_Q = \sum_i \mathsf{q}_i$ and the traced quadrupolar
642     density, $\mathsf{Q} = \mathsf{M}_Q / V$. A fluctuation formula can
643     be written for a system comprising quadrupolar
644     molecules,\cite{LoganI81,LoganII82,LoganIII82}
645 gezelter 4399 \begin{equation}
646 gezelter 4400 \mathsf{Q} - \frac{\mathbf{I}}{3} \mathrm{Tr}(\mathsf{Q}) = \epsilon_o
647     \frac{\braket{\mathsf{M}_Q^2}-{\braket{\mathsf{M}_Q}}^2}{15 \epsilon_o
648 gezelter 4419 V k_B T} \nabla\mathbf{E}.
649 gezelter 4399 \label{flucQuad}
650     \end{equation}
651 gezelter 4400 Some care is needed in the definitions of the averaged quantities. These
652     refer to the effective quadrupole moment of the system, and they are
653     computed as follows,
654     \begin{align}
655     \braket{\mathsf{M}_Q^2} &= \braket{3 \mathsf{M}_Q:\mathsf{M}_Q -
656     \mathrm{Tr}(\mathsf{M}_Q)^2 }\\
657     \braket{\mathsf{M}_Q}^2 &= 3 \braket{\mathsf{M}_Q}:\braket{\mathsf{M}_Q} -
658     \mathrm{Tr}(\braket{\mathsf{M}_Q})^2
659 gezelter 4401 \label{eq:flucQuad}
660 gezelter 4400 \end{align}
661 gezelter 4432 The bulk quadrupolarizability is given by,
662 gezelter 4399 \begin{equation}
663 gezelter 4432 \alpha_Q = \frac{\braket{\mathsf{M}_Q^2}-{\braket{\mathsf{M}_Q}}^2}{15 \epsilon_o V k_B T}.
664 gezelter 4399 \label{propConstQuad}
665     \end{equation}
666 gezelter 4432 Note that as in the dipolar case, $\alpha_Q$ and $\chi_Q$ are distinct
667 gezelter 4439 quantities. $\chi_Q$ measures the bulk response assuming an infinite
668 gezelter 4434 system and exact electrostatics, while $\alpha_Q$ is relatively simple
669     to compute from numerical simulations. As in the dipolar case,
670     estimation of the true bulk property requires correction for
671     truncation, shifting, and damping of the electrostatic interactions.
672 gezelter 4399
673     \subsection{Correction Factors}
674 gezelter 4400 In this section we generalize the treatment of Neumann \textit{et al.}
675     for quadrupolar fluids. Interactions involving multiple quadrupoles
676     are rank 4 tensors, and we therefore describe quantities in this
677     section using Einstein notation.
678    
679     In the presence of a uniform external field gradient,
680     $\partial_\alpha {E}^\circ_\beta $, the total field gradient at
681     $\mathbf{r}$ depends on the quadrupole polarization density at all
682     other points in the system,
683 mlamichh 4353 \begin{equation}
684 gezelter 4400 \partial_\alpha E_\beta(\mathbf{r}) = \partial_\alpha
685 mlamichh 4412 {E}^\circ_\beta(\mathbf{r}) + \frac{1}{8\pi \epsilon_o}\int
686 gezelter 4400 T_{\alpha\beta\gamma\delta}({\mathbf{r}-\mathbf{r}^\prime})
687     Q_{\gamma\delta}(\mathbf{r}^\prime) d\mathbf{r}^\prime
688 mlamichh 4353 \label{gradMaxwell}
689     \end{equation}
690 gezelter 4400 where and $T_{\alpha\beta\gamma\delta}$ is the quadrupole interaction
691     tensor connecting quadrupoles at $\mathbf{r}^\prime$ with the point of
692     interest ($\mathbf{r}$).
693 mlamichh 4353
694 gezelter 4408 % \begin{figure}
695     % \includegraphics[width=3in]{QuadrupoleFigure}
696     % \caption{With the real-space electrostatic methods, the molecular
697     % quadrupole tensor, $\mathbf{T}_{\alpha\beta\gamma\delta}(r)$,
698     % governing interactions between molecules is not the same for
699     % quadrupoles represented via sets of charges, point dipoles, or by
700     % single point quadrupoles.}
701     % \label{fig:quadrupolarFluid}
702     % \end{figure}
703 gezelter 4400
704     In simulations of quadrupolar fluids, the molecular quadrupoles may be
705     represented by closely-spaced point charges, by multiple point
706     dipoles, or by a single point quadrupole (see
707 gezelter 4408 Fig. \ref{fig:tensor}). In the case where point charges are
708 gezelter 4400 interacting via an electrostatic kernel, $v(r)$, the effective
709     molecular quadrupole tensor can obtained from four successive
710     applications of the gradient operator to the electrostatic kernel,
711     \begin{eqnarray}
712     T_{\alpha\beta\gamma\delta}(\mathbf{r}) &=& \nabla_\alpha \nabla_\beta
713     \nabla_\gamma \nabla_\delta v(r) \\
714     &=& \left(\delta_{\alpha\beta}\delta_{\gamma\delta} +
715     \delta_{\alpha\gamma}\delta_{\beta\delta}+
716     \delta_{\alpha\delta}\delta_{\beta\gamma}\right)\left(-\frac{v^\prime(r)}{r^3}
717     + \frac{v^{\prime \prime}(r)}{r^2}\right) \nonumber \\
718     & &+ \left(\delta_{\alpha\beta} r_\gamma r_\delta + 5 \mathrm{~permutations}
719     \right) \left(\frac{3v^\prime(r)}{r^5}-\frac{3v^{\prime \prime}(r)}{r^4} +
720     \frac{v^{\prime \prime \prime}(r)}{r^3}\right) \nonumber \\
721     & &+ r_\alpha r_\beta r_\gamma r_\delta
722     \left(-\frac{15v^\prime(r)}{r^7}+\frac{15v^{\prime \prime}(r)}{r^6}-\frac{6v^{\prime
723     \prime \prime}(r)}{r^5} + \frac{v^{\prime \prime \prime \prime}(r)}{r^4}\right),
724 mlamichh 4353 \label{quadCharge}
725 gezelter 4400 \end{eqnarray}
726     where $v(r)$ can either be the electrostatic kernel ($1/r$) or one of
727     the modified (Wolf or DSF) kernels.
728 mlamichh 4353
729 gezelter 4400 Similarly, when representing quadrupolar molecules with multiple point
730 gezelter 4413 \textit{dipoles}, the molecular quadrupole interaction tensor can be
731     obtained using two successive applications of the gradient operator to
732     the dipole interaction tensor,
733 gezelter 4400 \begin{eqnarray}
734     T_{\alpha\beta\gamma\delta}(\mathbf{r}) &=& \nabla_\alpha \nabla_\beta
735 gezelter 4413 T_{\gamma\delta}(\mathbf{r}) \\
736 gezelter 4400 & = & \delta_{\alpha\beta}\delta_{\gamma\delta} \frac{v^\prime_{21}(r)}{r} +
737     \left(\delta_{\alpha\gamma}\delta_{\beta\delta}+
738     \delta_{\alpha\delta}\delta_{\beta\gamma}\right)\frac{v_{22}(r)}{r^2}
739     \nonumber\\
740     & &+ \delta_{\gamma\delta} r_\alpha r_\beta
741     \left(\frac{v^{\prime \prime}_{21}(r)}{r^2}-\frac{v^\prime_{21}(r)}{r^3} \right)\nonumber \\
742     & &+\left(\delta_{\alpha\beta} r_\gamma r_\delta +
743     \delta_{\alpha\gamma} r_\beta r_\delta +\delta_{\alpha\delta}
744     r_\gamma r_\beta + \delta_{\beta\gamma} r_\alpha r_\delta
745     +\delta_{\beta\delta} r_\alpha r_\gamma \right)
746     \left(\frac{v^\prime_{22}(r)}{r^3}-\frac{2v_{22}(r)}{r^4}\right)
747     \nonumber \\
748     & &+ r_\alpha r_\beta r_\gamma r_\delta
749     \left(\frac{v^{\prime
750     \prime}_{22}(r)}{r^4}-\frac{5v^\prime_{22}(r)}{r^5}+\frac{8v_{22}(r)}{r^6}\right),
751 mlamichh 4353 \label{quadDip}
752 gezelter 4400 \end{eqnarray}
753 gezelter 4413 where $T_{\gamma\delta}(\mathbf{r})$ is a dipole-dipole interaction
754 gezelter 4432 tensor that depends on the level of the approximation (see
755     Eq. (\ref{dipole-diopleTensor})).\cite{PaperI,PaperII} Similarly
756 gezelter 4439 $v_{21}(r)$ and $v_{22}(r)$ are the radial functions for different real
757 gezelter 4432 space cutoff methods defined in the first paper in this
758     series.\cite{PaperI}
759 gezelter 4400
760     For quadrupolar liquids modeled using point quadrupoles, the
761     interaction tensor can be constructed as,
762     \begin{eqnarray}
763 mlamichh 4412 T_{\alpha\beta\gamma\delta}(\mathbf{r}) &=&
764 gezelter 4400 \left(\delta_{\alpha\beta}\delta_{\gamma\delta}
765     +
766     \delta_{\alpha\gamma}\delta_{\beta\delta}+
767     \delta_{\alpha\delta}\delta_{\beta\gamma}\right)v_{41}(r)
768 mlamichh 4412 + (\delta_{\gamma\delta} r_\alpha r_\beta + \mathrm{ 5\; permutations}) \frac{v_{42}(r)}{r^2} \nonumber \\
769 gezelter 4400 & & + r_\alpha r_\beta r_\gamma r_\delta \left(\frac{v_{43}(r)}{r^4}\right),
770 mlamichh 4353 \label{quadRadial}
771 gezelter 4400 \end{eqnarray}
772     where again $v_{41}(r)$, $v_{42}(r)$, and $v_{43}(r)$ are radial
773     functions defined in Paper I of the series. \cite{PaperI} Note that
774     these radial functions have different functional forms depending on
775     the level of approximation being employed.
776    
777 gezelter 4432 The integral in Eq. (\ref{gradMaxwell}) can be divided into two
778 gezelter 4400 parts, $|\mathbf{r}-\mathbf{r}^\prime|\rightarrow 0 $ and
779     $|\mathbf{r}-\mathbf{r}^\prime|> 0$. Since the self-contribution to
780 gezelter 4432 the field gradient vanishes at the singularity (see the supplemental
781 gezelter 4439 material), Eq. (\ref{gradMaxwell}) can be
782 gezelter 4432 written as,
783 mlamichh 4353 \begin{equation}
784 gezelter 4400 \partial_\alpha E_\beta(\mathbf{r}) = \partial_\alpha {E}^\circ_\beta(\mathbf{r}) +
785 mlamichh 4412 \frac{1}{8\pi \epsilon_o}\int\limits_{|\mathbf{r}-\mathbf{r}^\prime|> 0 }
786 gezelter 4400 T_{\alpha\beta\gamma\delta}(\mathbf{r}-\mathbf{r}^\prime)
787     {Q}_{\gamma\delta}(\mathbf{r}^\prime) d\mathbf{r}^\prime.
788 mlamichh 4353 \end{equation}
789 gezelter 4400 If $\mathbf{r} = \mathbf{r}^\prime$ is excluded from the integration,
790     the total gradient can be most easily expressed in terms of
791     traceless quadrupole density as below,\cite{LoganI81}
792 mlamichh 4353 \begin{equation}
793 gezelter 4400 \partial_\alpha E_\beta(\mathbf{r}) = \partial_\alpha
794 mlamichh 4412 {E}^\circ_\beta(\mathbf{r}) + \frac{1}{24\pi
795 gezelter 4400 \epsilon_o}\int\limits_{|\mathbf{r}-\mathbf{r}^\prime|> 0 }
796     T_{\alpha\beta\gamma\delta}(\mathbf{r}-\mathbf{r}^\prime) \Theta_{\gamma\delta}(\mathbf{r}') d\mathbf{r}^\prime,
797 mlamichh 4353 \end{equation}
798 gezelter 4400 where
799     $\Theta_{\alpha\beta} = 3Q_{\alpha\beta} - \delta_{\alpha\beta}Tr(Q)$
800     is the traceless quadrupole density. In analogy to
801 gezelter 4413 Eq. (\ref{pertQuad}) above, the quadrupole polarization density may
802     now be related to the quadrupolar susceptibility, $\chi_Q$,
803 mlamichh 4353 \begin{equation}
804 gezelter 4400 \frac{1}{3}{\Theta}_{\alpha\beta}(\mathbf{r}) = \epsilon_o {\chi}_Q
805 mlamichh 4412 \left[\partial_\alpha {E}^\circ_\beta(\mathbf{r}) + \frac{1}{24\pi
806 gezelter 4400 \epsilon_o}\int\limits_{|\mathbf{r}-\mathbf{r}^\prime|> 0 }
807     T_{\alpha\beta\gamma\delta}(\mathbf{r}-\mathbf{r}^\prime)
808 mlamichh 4412 \Theta_{\gamma\delta}(\mathbf{r}^\prime) d\mathbf{r}^\prime \right].
809 mlamichh 4353 \end{equation}
810 mlamichh 4412 For periodic boundaries and with a uniform imposed
811 gezelter 4400 $\partial_\alpha E^\circ_\beta$, the quadrupole density
812     ${\Theta}_{\alpha\beta}$ will be uniform over the entire space. After
813     performing a Fourier transform (see the Appendix in
814     ref. \onlinecite{NeumannI83}) we obtain,
815 mlamichh 4353 \begin{equation}
816 gezelter 4400 \frac{1}{3}\tilde{\Theta}_{\alpha\beta}(\mathbf{k})=
817     \epsilon_o {\chi}_Q \left[{\partial_\alpha
818 mlamichh 4412 \tilde{E}^\circ_\beta}(\mathbf{k})+ \frac{1}{24\pi
819 gezelter 4400 \epsilon_o} \tilde{T}_{\alpha\beta\gamma\delta}(\mathbf{k})
820     \tilde{\Theta}_{\gamma\delta}(\mathbf{k})\right].
821 mlamichh 4412 \label{fourierQuad}
822 mlamichh 4353 \end{equation}
823 gezelter 4413 If the applied field gradient is homogeneous over the entire volume,
824     ${\partial_ \alpha \tilde{E}^\circ_\beta}(\mathbf{k}) = 0 $ except at
825     $ \mathbf{k} = 0$. Similarly, the quadrupolar polarization density can
826     also considered uniform over entire space. As in the dipolar case,
827     \cite{NeumannI83} the only relevant contribution from the interaction
828 gezelter 4432 tensor will also be when $\mathbf{k} = 0$. Therefore Eq.
829 gezelter 4413 (\ref{fourierQuad}) can be written as,
830 mlamichh 4412 \begin{equation}
831     \frac{1}{3}\tilde{\Theta}_{\alpha\beta}(\mathrm{0})=
832     \epsilon_o {\chi}_Q \left[{\partial_\alpha
833     \tilde{E}^\circ_\beta}(\mathrm{0})+ \frac{1}{24\pi
834     \epsilon_o} \tilde{T}_{\alpha\beta\gamma\delta}(\mathrm{0})
835     \tilde{\Theta}_{\gamma\delta}(\mathrm{0})\right].
836     \label{fourierQuadZeroK}
837     \end{equation}
838 gezelter 4413 The quadrupolar tensor
839 gezelter 4415 $\tilde{T}_{\alpha\beta\gamma\delta}(\mathrm{0})$ is a rank 4 tensor
840 gezelter 4413 with 81 elements. The only non-zero elements, however, are those with
841     two doubly-repeated indices, \textit{i.e.}
842     $\tilde{T}_{aabb}(\mathrm{0})$ and all permutations of these indices.
843     The special case of quadruply-repeated indices,
844     $\tilde{T}_{aaaa}(\mathrm{0})$ also survives (see appendix
845     \ref{ap:quadContraction}). Furthermore, for the both diagonal and
846     non-diagonal components of the quadrupolar polarization
847     $\tilde{\Theta}_{\alpha\beta}$, we can contract the second term in
848 gezelter 4432 Eq. \ref{fourierQuadZeroK} (see appendix
849 gezelter 4413 \ref{ap:quadContraction}):
850 mlamichh 4412 \begin{equation}
851 gezelter 4413 \tilde{T}_{\alpha\beta\gamma\delta}(\mathrm{0})\tilde{\Theta}_{\gamma\delta}(\mathrm{0})=
852     8 \pi \mathrm{B} \tilde{\Theta}_{\alpha\beta}(\mathrm{0}).
853 mlamichh 4412 \label{quadContraction}
854     \end{equation}
855 gezelter 4413 Here $\mathrm{B} = \tilde{T}_{abab}(\mathrm{0}) / 4 \pi$ for
856 gezelter 4432 $a \neq b$. Using this quadrupolar contraction we can solve Eq.
857 gezelter 4415 \ref{fourierQuadZeroK} as follows
858 gezelter 4400 \begin{eqnarray}
859 mlamichh 4412 \frac{1}{3}\tilde{\Theta}_{\alpha\beta}(\mathrm{0}) &=& \epsilon_o
860 gezelter 4400 {\chi}_Q
861     \left[{\partial_\alpha
862 mlamichh 4412 \tilde{E}^\circ_\beta}(\mathrm{0})+
863 gezelter 4413 \frac{\mathrm{B}}{3
864 mlamichh 4412 \epsilon_o}
865     {\tilde{\Theta}}_{\alpha\beta}(\mathrm{0})\right]
866     \nonumber \\
867     &=& \left[\frac{\epsilon_o {\chi}_Q} {1-{\chi}_Q \mathrm{B}}\right]
868 gezelter 4413 {\partial_\alpha \tilde{E}^\circ_\beta}(\mathrm{0}).
869 gezelter 4415 \label{fourierQuad2}
870 mlamichh 4412 \end{eqnarray}
871 gezelter 4419 In real space, the correction factor is found to be,
872 gezelter 4413 \begin{equation}
873     \mathrm{B} = \frac{1}{4 \pi} \tilde{T}_{abab}(0) = \frac{1}{4 \pi} \int_V {T}_{abab}(\mathbf{r}) d\mathbf{r},
874     \end{equation}
875 mlamichh 4412 %If the applied field gradient is homogeneous over the
876     %entire volume, ${\partial_ \alpha \tilde{E}^\circ_\beta}(\mathbf{k}) = 0 $ except at
877     %$ \mathbf{k} = 0$. As in the dipolar case, the only relevant
878     %contribution of the interaction tensor will also be when $\mathbf{k} = 0$.
879     %Therefore equation (\ref{fourierQuad}) can be written as,
880     %\begin{equation}
881     %\frac{1}{3}{\tilde{\Theta}}_{\alpha\beta}({0}) = \epsilon_o {\chi}_Q
882     %\left(1-\frac{1}{4\pi} {\chi}_Q \tilde{T}_{\alpha\beta\alpha\beta}({0})\right)^{-1} \partial_\alpha \tilde{E}^\circ_\beta({0})
883     %\label{fourierQuad2}
884     %\end{equation}
885     %where $\tilde{T}_{\alpha\beta\alpha\beta}({0})$ can be evaluated as,
886     %\begin{equation}
887     %\tilde{T}_{\alpha\beta\alpha\beta}({0}) = \int_V {T}_{\alpha\beta\alpha\beta}(\mathbf{r})d\mathbf{r}
888     %\label{realTensorQaud}
889     %\end{equation}
890     %We may now define a quantity to represent the contribution from the
891     %$\mathbf{k} = 0$ portion of the interaction tensor,
892     %\begin{equation}
893     %B = \frac{1}{4 \pi} \int_V T_{\alpha \beta \alpha \beta}(\mathbf{r})
894     %d\mathbf{r}
895     %\end{equation}
896 gezelter 4400 which has been integrated over the interaction volume $V$ and has
897     units of $\mathrm{length}^{-2}$.
898 mlamichh 4353
899 gezelter 4432 In terms of the traced quadrupole moment, Eq. (\ref{fourierQuad2})
900 gezelter 4413 can be written,
901 mlamichh 4353 \begin{equation}
902 gezelter 4400 \mathsf{Q} - \frac{\mathbf{I}}{3} \mathrm{Tr}(\mathsf{Q})
903 mlamichh 4412 = \frac{\epsilon_o {\chi}_Q}{1- {\chi}_Q \mathrm{B}} \nabla \mathbf{E}^\circ
904 mlamichh 4353 \label{tracedConstQuad}
905     \end{equation}
906 gezelter 4400 Comparing (\ref{tracedConstQuad}) and (\ref{flucQuad}) we obtain,
907 mlamichh 4353 \begin{equation}
908 gezelter 4400 \frac{\braket{\mathsf{M}_Q^2}-{\braket{\mathsf{M}_Q}}^2}{15 \epsilon_o
909 mlamichh 4412 V k_B T} = \frac{\chi_Q} {1 - \chi_Q \mathrm{B}},
910 mlamichh 4353 \end{equation}
911 gezelter 4400 or equivalently,
912 mlamichh 4353 \begin{equation}
913 gezelter 4400 \chi_Q = \frac{\braket{\mathsf{M}_Q^2}-{\braket{\mathsf{M}_Q}}^2}{15 \epsilon_o
914 mlamichh 4412 V k_B T} \left(1 + \mathrm{B} \frac{\braket{\mathsf{M}_Q^2}-{\braket{\mathsf{M}_Q}}^2}{15 \epsilon_o
915 gezelter 4432 V k_B T} \right)^{-1}.
916 gezelter 4400 \label{eq:finalForm}
917 mlamichh 4353 \end{equation}
918 gezelter 4400 Eq. (\ref{eq:finalForm}) now expresses a bulk property (the
919 gezelter 4413 quadrupolar susceptibility, $\chi_Q$) in terms of a fluctuation in the
920     system quadrupole moment and a quadrupolar correction factor
921     ($\mathrm{B}$). The correction factors depend on the cutoff method
922     being employed in the simulation, and these are listed in Table
923     \ref{tab:B}.
924 mlamichh 4353
925 gezelter 4400 In terms of the macroscopic quadrupole polarizability, $\alpha_Q$,
926     which may be thought of as the ``conducting boundary'' version of the
927     susceptibility,
928 mlamichh 4353 \begin{equation}
929 mlamichh 4412 \chi_Q = \frac{\alpha_Q}{1 + \mathrm{B} \alpha_Q}.
930 gezelter 4408 \label{eq:quadrupolarSusceptiblity}
931 mlamichh 4353 \end{equation}
932 mlamichh 4412 If an electrostatic method produces $\mathrm{B} \rightarrow 0$, the computed
933 gezelter 4400 quadrupole polarizability and quadrupole susceptibility converge to
934     the same value.
935    
936     \begin{sidewaystable}
937 gezelter 4415 \caption{Expressions for the quadrupolar correction factor
938     ($\mathrm{B}$) for the real-space electrostatic methods in terms
939     of the damping parameter ($\alpha$) and the cutoff radius
940     ($r_c$). The units of the correction factor are
941     $ \mathrm{length}^{-2}$ for quadrupolar fluids.}
942 mlamichh 4353 \label{tab:B}
943 gezelter 4416 \begin{tabular}{l|c|c|c} \toprule
944     \multirow{2}{*}{Method} & \multicolumn{3}{c}{Molecular Representation} \\
945     & charges & dipoles & quadrupoles \\\colrule
946 gezelter 4431 Shifted Potential (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}}$ \\
947 mlamichh 4412 Gradient-shifted (GSF) & $- \frac{8 \alpha^5 {r_c}^3e^{-\alpha^2 r_c^2}}{15\sqrt{\pi}} $ & 0 & $-\frac{4{\alpha}^7{r_c}^5 e^{-\alpha^2 r_c^2}(-1+2\alpha ^2 r_c^2)}{45\sqrt{\pi}}$\\
948     Taylor-shifted (TSF) & $ -\frac{8 \alpha^5 {r_c}^3e^{-\alpha^2 r_c^2}}{15\sqrt{\pi}} $ & $\frac{4\;\mathrm{erfc(\alpha r_c)}}{{5r_c}^2} + \frac{8\alpha e^{-\alpha^2{r_c}^2}\left(3+ 2\alpha^2 {r_c}^2 +\alpha^4{r_c}^4 \right)}{15\sqrt{\pi}r_c}$ & $\frac{2\;\mathrm{erfc}(\alpha r_c )}{{r_c}^2} + \frac{4{\alpha}e^{-\alpha^2 r_c^2}\left(45 + 30\alpha ^2 {r_c}^2 + 12\alpha^4 {r_c}^4 + 3\alpha^6 {r_c}^6 + 2 \alpha^8 {r_c}^8\right)}{45\sqrt{\pi}{r_c}}$ \\
949 gezelter 4408 \colrule
950 gezelter 4432 %Spherical Cutoff (SC) & \multicolumn{3}{c}{$ -\frac{8 \alpha^5
951     % {r_c}^3e^{-\alpha^2 r_c^2}}{15\sqrt{\pi}} $}\\
952 gezelter 4416 Ewald-Kornfeld (EK) & \multicolumn{3}{c}{$ -\frac{8 \kappa^5 {r_c}^3e^{-\kappa^2 r_c^2}}{15\sqrt{\pi}}$} \\
953 gezelter 4408 \botrule
954 gezelter 4400 \end{tabular}
955     \end{sidewaystable}
956 gezelter 4399
957     \section{Screening of Charges by Multipolar Fluids}
958 gezelter 4401 \label{sec:PMF}
959     In a dipolar fluid, the static dielectric constant is also a measure
960     of the ability of the fluid to screen charges from one another. A set
961     of point charges creates an inhomogeneous field in the fluid, and the
962     fluid responds to this field as if it was created externally or via
963     local polarization fluctuations. For this reason, the dielectric
964     constant can be used to estimate an effective potential between two
965     point charges ($C_i$ and $C_j$) embedded in the fluid,
966 gezelter 4399 \begin{equation}
967 gezelter 4401 U_\mathrm{effective} = \frac{C_i C_j}{4 \pi \epsilon_0 \epsilon
968     r_{ij}}.
969     \label{eq:effectivePot}
970 gezelter 4399 \end{equation}
971 gezelter 4401
972     The same set of point charges can also create an inhomogeneous field
973     \textit{gradient}, and this will cause a response in a quadrupolar
974     fluid that will also cause an effective screening. As discussed above,
975 gezelter 4431 however, the relevant physical property in quadrupolar fluids is the
976 gezelter 4401 susceptibility, $\chi_Q$. The screening dielectric associated with
977     the quadrupolar susceptibility is defined as,\cite{Ernst92}
978 gezelter 4399 \begin{equation}
979 gezelter 4401 \epsilon = 1 + \chi_Q G = 1 + G \frac{\alpha_Q}{1 + \alpha_Q B}
980     \label{eq:dielectricFromQuadrupoles}
981 gezelter 4399 \end{equation}
982 gezelter 4401 where $G$ is a geometrical factor that depends on the geometry of the
983     field perturbation,
984     \begin{equation}
985     G = \frac{\int_V \left| \nabla \mathbf{E}^\circ \right|^2 d\mathbf{r}}
986     {\int_V \left|\mathbf{E}^\circ\right|^2 d\mathbf{r}}
987     \end{equation}
988     integrated over the interaction volume. Note that this geometrical
989     factor is also required to compute effective dielectric constants even
990     when the field gradient is homogeneous over the entire sample.
991 gezelter 4399
992 gezelter 4401 To measure effective screening in a multipolar fluid, we compute an
993     effective interaction potential, the potential of mean force (PMF),
994 gezelter 4432 between positively and negatively charged ions when they are screened
995     by the intervening fluid. The PMF is obtained from a sequence of
996 gezelter 4401 simulations in which two ions are constrained to a fixed distance, and
997     the average constraint force to hold them at a fixed distance $r$ is
998     collected during a long simulation,\cite{Wilfred07}
999 gezelter 4400 \begin{equation}
1000 gezelter 4401 w(r) = \int_{r_o}^{r}\left\langle \frac{\partial f}{\partial r^\prime}
1001     \right\rangle_{r^\prime} dr^\prime + 2k_BT \ln\left(\frac{r}{r_o}\right) + w(r_o),
1002     \label{eq:pmf}
1003 gezelter 4400 \end{equation}
1004 gezelter 4401 where $\braket{\partial f/\partial r^\prime}_{r^\prime}$ is the mean
1005     constraint force required to hold the ions at distance $r^\prime$,
1006     $2k_BT \log(r/r_o)$ is the Fixman factor,\cite{Fixman:1974fk} and
1007     $r_o$ is a reference position (usually taken as a large separation
1008 gezelter 4431 between the ions). If the dielectric constant is a good measure of the
1009 gezelter 4401 screening at all inter-ion separations, we would expect $w(r)$ to have
1010     the form in Eq. (\ref{eq:effectivePot}). Because real fluids are not
1011     continuum dielectrics, the effective dielectric constant is a function
1012     of the interionic separation,
1013 gezelter 4400 \begin{equation}
1014 gezelter 4425 \epsilon(r) = \frac{u_\mathrm{raw}(r) - u_\mathrm{raw}(r_o) }{w(r) - w(r_o)}
1015 gezelter 4400 \end{equation}
1016 gezelter 4401 where $u_\mathrm{raw}(r)$ is the direct charge-charge interaction
1017     potential that is in use during the simulation. $\epsilon(r)$ may
1018     vary considerably from the bulk estimates at short distances, although
1019     it should converge to the bulk value as the separation between the
1020     ions increases.
1021 gezelter 4400
1022 gezelter 4399 \section{Simulation Methodology}
1023 mlamichh 4422
1024 gezelter 4401 To test the formalism developed in the preceding sections, we have
1025     carried out computer simulations using three different techniques: i)
1026     simulations in the presence of external fields, ii) equilibrium
1027     calculations of box moment fluctuations, and iii) potentials of mean
1028     force (PMF) between embedded ions. In all cases, the fluids were
1029     composed of point multipoles protected by a Lennard-Jones potential.
1030     The parameters used in the test systems are given in table
1031     \ref{Tab:C}.
1032 mlamichh 4422
1033     \begin{sidewaystable}
1034 gezelter 4401 \caption{\label{Tab:C}The parameters used in simulations to evaluate
1035     the dielectric response of the new real-space methods.}
1036 mlamichh 4389 \begin{tabularx}{\textwidth}{r|cc|YYccc|Yccc} \hline
1037     & \multicolumn{2}{c|}{LJ parameters} &
1038     \multicolumn{5}{c|}{Electrostatic moments} & & & & \\
1039     Test system & $\sigma$& $\epsilon$ & $C$ & $D$ &
1040     $Q_{xx}$ & $Q_{yy}$ & $Q_{zz}$ & mass & $I_{xx}$ & $I_{yy}$ &
1041     $I_{zz}$ \\ \cline{6-8}\cline{10-12}
1042     & (\AA) & (kcal/mol) & (e) & (debye) & \multicolumn{3}{c|}{(debye \AA)} & (amu) & \multicolumn{3}{c}{(amu
1043     \AA\textsuperscript{2})} \\ \hline
1044 gezelter 4401 Dipolar fluid & 3.41 & 0.2381 & - & 1.4026 &-&-&-& 39.948 & 11.613 & 11.613 & 0.0 \\
1045     Quadrupolar fluid & 2.985 & 0.265 & - & - & 0.0 & 0.0 &-2.139 & 18.0153 & 43.0565 & 43.0565 & 0.0 \\
1046 mlamichh 4389 \ce{q+} & 1.0 & 0.1 & +1 & - & - & - & - & 22.98 & - & - & - \\
1047     \ce{q-} & 1.0 & 0.1 & -1 & - & - & - & - & 22.98 & - & - & - \\ \hline
1048     \end{tabularx}
1049 mlamichh 4422 \end{sidewaystable}
1050 mlamichh 4363
1051 gezelter 4401 The first of the test systems consists entirely of fluids of point
1052     dipolar or quadrupolar molecules in the presence of constant field or
1053     field gradients. Since there are no isolated charges within the
1054     system, the divergence of the field will be zero, \textit{i.e.}
1055 gezelter 4432 $\nabla \cdot \mathbf{E} = 0$. This condition can be satisfied
1056 gezelter 4401 by using the relatively simple applied potential as described in the
1057 gezelter 4439 supplemental material.
1058 mlamichh 4353
1059 gezelter 4401 When a constant electric field or field gradient is applied to the
1060     system, the molecules align along the direction of the applied field,
1061     and polarize to a degree determined both by the strength of the field
1062 gezelter 4431 and the fluid's polarizability. We have calculated ensemble averages
1063 gezelter 4401 of the box dipole and quadrupole moments as a function of the strength
1064     of the applied fields. If the fields are sufficiently weak, the
1065     response is linear in the field strength, and one can easily compute
1066     the polarizability directly from the simulations.
1067 mlamichh 4389
1068 gezelter 4401 The second set of test systems consists of equilibrium simulations of
1069     fluids of point dipolar or quadrupolar molecules simulated in the
1070     absence of any external perturbation. The fluctuation of the ensemble
1071     averages of the box multipolar moment was calculated for each of the
1072     multipolar fluids. The box multipolar moments were computed as simple
1073     sums over the instantaneous molecular moments, and fluctuations in
1074     these quantities were obtained from Eqs. (\ref{eq:flucDip}) and
1075 gezelter 4439 (\ref{eq:flucQuad}). The macroscopic polarizabilities of the system
1076     were derived using Eqs.(\ref{flucDipole}) and (\ref{flucQuad}).
1077 mlamichh 4389
1078 gezelter 4401 The final system consists of dipolar or quadrupolar fluids with two
1079     oppositely charged ions embedded within the fluid. These ions are
1080     constrained to be at fixed distance throughout a simulation, although
1081     they are allowed to move freely throughout the fluid while satisfying
1082     that constraint. Separate simulations were run at a range of
1083     constraint distances. A dielectric screening factor was computed using
1084     the ratio between the potential between the two ions in the absence of
1085     the fluid medium and the PMF obtained from the simulations.
1086 mlamichh 4389
1087 gezelter 4401 We carried out these simulations for all three of the new real-space
1088     electrostatic methods (SP, GSF, and TSF) that were developed in the
1089     first paper (Ref. \onlinecite{PaperI}) in the series. The radius of
1090     the cutoff sphere was taken to be 12~\AA. Each of the real space
1091     methods also depends on an adjustable damping parameter $\alpha$ (in
1092     units of $\mathrm{length}^{-1}$). We have selected ten different
1093     values of damping parameter: 0.0, 0.05, 0.1, 0.15, 0.175, 0.2, 0.225,
1094     0.25, 0.3, and 0.35~\AA$^{-1}$ in our simulations of the dipolar
1095     liquids, while four values were chosen for the quadrupolar fluids:
1096     0.0, 0.1, 0.2, and 0.3~\AA$^{-1}$.
1097 mlamichh 4389
1098 gezelter 4401 For each of the methods and systems listed above, a reference
1099     simulation was carried out using a multipolar implementation of the
1100     Ewald sum.\cite{Smith82,Smith98} A default tolerance of
1101     $1 \times 10^{-8}$~kcal/mol was used in all Ewald calculations,
1102     resulting in Ewald coefficient 0.3119~\AA$^{-1}$ for a cutoff radius
1103     of 12~\AA. All of the electrostatics and constraint methods were
1104     implemented in our group's open source molecular simulation program,
1105     OpenMD,\cite{Meineke05,openmd} which was used for all calculations in
1106     this work.
1107    
1108     Dipolar systems contained 2048 Lennard-Jones-protected point dipolar
1109     (Stockmayer) molecules with reduced density $\rho^* = 0.822$,
1110     temperature $T^* = 1.15$, moment of inertia $I^* = 0.025$, and dipole
1111     moment $\mu^* = \sqrt{3.0}$. These systems were equilibrated for
1112     0.5~ns in the canonical (NVT) ensemble. Data collection was carried
1113     out over a 1~ns simulation in the microcanonical (NVE) ensemble. Box
1114     dipole moments were sampled every fs. For simulations with external
1115 gezelter 4432 perturbations, field strengths ranging from
1116 gezelter 4439 0 to 10 $10^{-3}$~V/\AA\ with increments of $ 10^{-4}$~V/\AA\
1117 gezelter 4432 were carried out for each system. For dipolar systems the interaction
1118     potential between molecules $i$ and $j$,
1119     \begin{equation}
1120     u_{ij}(\mathbf{r}_{ij}, \mathbf{D}_i, \mathbf{D}_j) = 4 \epsilon \left(
1121     \left(\frac{\sigma}{r_{ij}} \right)^{12} - \left(\frac{\sigma}{r_{ij}}
1122     \right)^{6} \right) - \mathbf{D}_i \cdot
1123     \mathbf{T}(\mathbf{r}_{ij}) \cdot \mathbf{D}_j
1124     \end{equation}
1125     where the dipole interaction tensor, $\mathbf{T}(\mathbf{r})$, is
1126     given in Eq. (\ref{dipole-diopleTensor}).
1127 gezelter 4401
1128     Quadrupolar systems contained 4000 linear point quadrupoles with a
1129 gezelter 4408 density $2.338 \mathrm{~g/cm}^3$ at a temperature of 500~K. These
1130 gezelter 4401 systems were equilibrated for 200~ps in a canonical (NVT) ensemble.
1131     Data collection was carried out over a 500~ps simulation in the
1132 gezelter 4408 microcanonical (NVE) ensemble. Components of box quadrupole moments
1133     were sampled every 100 fs. For quadrupolar simulations with external
1134 gezelter 4401 field gradients, field strengths ranging from
1135 gezelter 4408 $0 - 9 \times 10^{-2}$~V/\AA$^2$ with increments of
1136 gezelter 4432 $10^{-2}$~V/\AA$^2$ were carried out for each system. For quadrupolar
1137     systems the interaction potential between molecules $i$ and $j$,
1138     \begin{equation}
1139     u_{ij}(\mathbf{r}_{ij}, \mathsf{Q}_i, \mathsf{Q}_j) = 4 \epsilon \left(
1140     \left(\frac{\sigma}{r_{ij}} \right)^{12} - \left(\frac{\sigma}{r_{ij}}
1141     \right)^{6} \right) + \mathsf{Q}_i \colon
1142     \mathbf{T}(\mathbf{r}_{ij}) \colon \mathsf{Q}_j
1143     \end{equation}
1144     where the quadrupole interaction tensor is given in
1145     Eq. (\ref{quadRadial}).
1146 gezelter 4401
1147     To carry out the PMF simulations, two of the multipolar molecules in
1148     the test system were converted into \ce{q+} and \ce{q-} ions and
1149     constrained to remain at a fixed distance for the duration of the
1150 gezelter 4408 simulation. The constrained distance was then varied from 5--12~\AA.
1151     In the PMF calculations, all simulations were equilibrated for 500 ps
1152     in the NVT ensemble and run for 5 ns in the microcanonical (NVE)
1153     ensemble. Constraint forces were sampled every 20~fs.
1154 mlamichh 4389
1155     \section{Results}
1156 mlamichh 4397 \subsection{Dipolar fluid}
1157     \begin{figure}
1158 gezelter 4428 \includegraphics[width=5in]{dielectricFinal_Dipole.eps}
1159 gezelter 4401 \caption{The polarizability ($\alpha_D$), correction factor ($A$), and
1160     dielectric constant ($\epsilon$) for the test dipolar fluid. The
1161     left panels were computed using external fields, and those on the
1162     right are the result of equilibrium fluctuations. In the GSF and SP
1163 gezelter 4432 methods, the corrections are large for small values of $\alpha$, and
1164     an optimal damping coefficient is evident around 0.25 \AA$^{-1}$.
1165 gezelter 4401 The dashed lines in the upper panel indicate back-calculation of the
1166 gezelter 4404 polarizability using the Ewald estimate (Refs. \onlinecite{Adams81}
1167     and \onlinecite{NeumannI83}) for the dielectric constant.}
1168 mlamichh 4397 \label{fig:dielectricDipole}
1169     \end{figure}
1170 gezelter 4432 The bulk polarizability ($\alpha_D$) for the dipolar fluid is
1171 gezelter 4408 shown in the upper panels in Fig. \ref{fig:dielectricDipole}. The
1172 gezelter 4401 polarizability obtained from the both perturbation and fluctuation
1173     approaches are in excellent agreement with each other. The data also
1174 gezelter 4431 show a strong dependence on the damping parameter for both the Shifted
1175 gezelter 4401 Potential (SP) and Gradient Shifted force (GSF) methods, while Taylor
1176     shifted force (TSF) is largely independent of the damping
1177     parameter.
1178 mlamichh 4397
1179 gezelter 4401 The calculated correction factors discussed in section
1180 gezelter 4408 \ref{sec:corrFactor} are shown in the middle panels. Because the TSF
1181 gezelter 4401 method has $A = 1$ for all values of the damping parameter, the
1182     computed polarizabilities need no correction for the dielectric
1183     calculation. The value of $A$ varies with the damping parameter in
1184     both the SP and GSF methods, and inclusion of the correction yields
1185     dielectric estimates (shown in the lower panel) that are generally too
1186 gezelter 4408 large until the damping reaches $\sim$~0.25~\AA$^{-1}$. Above this
1187 gezelter 4433 value, the dielectric constants are in reasonable agreement with
1188 gezelter 4408 previous simulation results.\cite{NeumannI83}
1189 mlamichh 4397
1190 gezelter 4401 Figure \ref{fig:dielectricDipole} also contains back-calculations of
1191     the polarizability using the reference (Ewald) simulation
1192     results.\cite{NeumannI83} These are indicated with dashed lines in the
1193     upper panels. It is clear that the expected polarizability for the SP
1194     and GSF methods are quite close to results obtained from the
1195     simulations. This indicates that the correction formula for the
1196 gezelter 4433 dipolar fluid (Eq. \ref{correctionFormula}) is extraordinarily
1197 gezelter 4434 sensitive when the value of $A$ deviates significantly from unity. It
1198     is also apparent that Gaussian damping is essential for capturing the
1199     field effects from other dipoles. Eq. (\ref{correctionFormula}) works
1200     well when real-space methods employ moderate damping, but is not
1201     capable of providing adequate correction for undamped or weakly-damped
1202     multipoles.
1203 gezelter 4401
1204 gezelter 4432 Because the dielectric correction in Eq. (\ref{correctionFormula}) is
1205     so sensitive to $A$ values away from unity, the entries in table
1206     \ref{tab:A} can provide an effective minimum on the values of $\alpha$
1207     that should be used. With a minimum $A=0.995$ and a cutoff radius of
1208     12~\AA, the minimum $\alpha$ values are 0.241~\AA$^{-1}$ (SP) or
1209 gezelter 4433 0.268~\AA$^{-1}$ (GSF). The TSF method is not sensitive to the choice
1210 gezelter 4432 of damping parameter.
1211 gezelter 4401
1212 mlamichh 4397 \begin{figure}
1213 gezelter 4428 \includegraphics[width=4in]{screen_pmf.eps}
1214 gezelter 4401 \caption{The distance-dependent screening factor, $\epsilon(r)$,
1215     between two ions immersed in the dipolar fluid. The new methods are
1216     shown in separate panels, and different values of the damping
1217     parameter ($\alpha$) are indicated with different symbols. All of
1218     the methods appear to be converging to the bulk dielectric constant
1219 gezelter 4432 ($\sim 65$) for higher values of $\alpha$ and at large ion
1220     separations.}
1221 mlamichh 4397 \label{fig:ScreeningFactor_Dipole}
1222     \end{figure}
1223 gezelter 4401 We have also evaluated the distance-dependent screening factor,
1224     $\epsilon(r)$, between two oppositely charged ions when they are
1225     placed in the dipolar fluid. These results were computed using
1226     Eq. \ref{eq:pmf} and are shown in Fig.
1227     \ref{fig:ScreeningFactor_Dipole}.
1228 mlamichh 4397
1229 gezelter 4401 The screening factor is similar to the dielectric constant, but
1230     measures a local property of the ions in the fluid and depends on both
1231     ion-dipole and dipole-dipole interactions. These interactions depend
1232     on the distance between ions as well as the electrostatic interaction
1233     methods utilized in the simulations. The screening should converge to
1234     the dielectric constant when the field due to ions is small. This
1235     occurs when the ions are separated (or when the damping parameter is
1236     large). In Fig. \ref{fig:ScreeningFactor_Dipole} we observe that for
1237     the higher value of damping alpha \textit{i.e.}
1238 gezelter 4434 $\alpha > 0.2$~\AA$^{-1}$ and large separation between ions, the
1239     screening factor does indeed approach the correct dielectric constant.
1240 mlamichh 4397
1241 gezelter 4401 It is also notable that the TSF method again displays smaller
1242     perturbations away from the correct dielectric screening behavior. We
1243 gezelter 4439 also observe that for TSF, the method yields high dielectric screening
1244     even for lower values of $\alpha$.
1245 gezelter 4401
1246     At short distances, the presence of the ions creates a strong local
1247     field that acts to align nearby dipoles nearly perfectly in opposition
1248     to the field from the ions. This has the effect of increasing the
1249     effective screening when the ions are brought close to one another.
1250 gezelter 4428 This effect is present even in the full Ewald treatment, and indicates
1251     that the local ordering behavior is being captured by all of the
1252     moderately-damped real-space methods.
1253 gezelter 4401
1254 gezelter 4439 \subsubsection*{Distance-dependent Kirkwood factors}
1255 gezelter 4432 One of the most sensitive measures of dipolar ordering in a liquid is
1256     the disance dependent Kirkwood factor,
1257     \begin{equation}
1258     G_K(r) = \left< \frac{1}{N} \sum_{i} \sum_{\substack{j \\ r_{ij} < r}}
1259     \frac{\mathbf{D}_i \cdot \mathbf{D}_j}{\left| D_i \right| \left| D_j
1260     \right|} \right>
1261     \label{eq:kirkwood}
1262     \end{equation}
1263     which measures the net orientational (cosine) ordering of dipoles
1264     inside a sphere of radius $r$. The outer brackets denote a
1265     configurational average. Figure \ref{fig:kirkwood} shows $G_K(r)$ for
1266 gezelter 4434 the three real space methods with $r_c = 3.52 \sigma = 12$~\AA~ and
1267     for the Ewald sum. These results were obtained from unperturbed 5~ns
1268     simulations of the dipolar fluid in the microcanonical (NVE)
1269     ensemble. For SP and GSF, the underdamped cases exhibit the ``hole''
1270     at $r_c$ that is sometimes seen in cutoff-based method simulations of
1271     liquid water,\cite{Mark:2002rt,Fukuda:2012fr} but for values of
1272     $\alpha > 0.225$ \AA$^{-1}$, agreement with the Ewald results is good.
1273 gezelter 4432 \begin{figure}
1274     \includegraphics[width=\linewidth]{kirkwood.eps}
1275     \caption{The distance-dependent Kirkwood factors of the dipolar system
1276     for the three real space methods at a range of Gaussian damping
1277     parameters ($\alpha$) with a cutoff $r_c = 3.52 \sigma$.}
1278     \label{fig:kirkwood}
1279     \end{figure}
1280     Note that like the dielectric constant, $G_K(r)$ can also be corrected
1281     using the expressions for $A$ in table \ref{tab:A}. This is discussed
1282 gezelter 4439 in more detail in the supplemental material.
1283 gezelter 4432
1284 mlamichh 4397 \subsection{Quadrupolar fluid}
1285     \begin{figure}
1286 gezelter 4428 \includegraphics[width=4in]{polarizabilityFinal_Quad.eps}
1287 gezelter 4401 \caption{The quadrupole polarizability ($\alpha_Q$), correction factor
1288     ($B$), and susceptibility ($\chi_Q$) for the test quadrupolar
1289     fluid. The left panels were computed using external field gradients,
1290     and those on the right are the result of equilibrium fluctuations.
1291     The GSF and SP methods allow nearly unmodified use of the
1292     ``conducting boundary'' or polarizability results in place of the
1293 gezelter 4404 bulk susceptibility.}
1294 mlamichh 4397 \label{fig:dielectricQuad}
1295     \end{figure}
1296 gezelter 4434 The polarizability ($\alpha_Q$), correction factor ($\mathrm{B}$), and
1297 gezelter 4401 susceptibility ($\chi_Q$) for the quadrupolar fluid is plotted against
1298     damping parameter Fig. \ref{fig:dielectricQuad}. In quadrupolar
1299     fluids, both the polarizability and susceptibility have units of
1300 gezelter 4434 $\mathrm{length}^2$. Although the susceptibility has dimensionality,
1301     it is the relevant measure of macroscopic quadrupolar
1302 gezelter 4419 properties.\cite{JeonI03, JeonII03} The left panel in
1303 gezelter 4401 Fig. \ref{fig:dielectricQuad} shows results obtained from the applied
1304     field gradient simulations whereas the results from the equilibrium
1305     fluctuation formula are plotted in the right panels.
1306 mlamichh 4397
1307 gezelter 4401 The susceptibility for the quadrupolar fluid is obtained from
1308     quadrupolarizability and a correction factor using
1309     Eq. (\ref{eq:quadrupolarSusceptiblity}). The susceptibilities are
1310     shown in the bottom panels of Fig. \ref{fig:dielectricQuad}. All three
1311 gezelter 4434 methods: (SP, GSF, and TSF) produce small correction factors,
1312     $\mathrm{B}$, so all show similar susceptibilities over the range of
1313     damping parameters. This shows that susceptibility derived using the
1314 gezelter 4432 quadrupolarizability and the correction factors are essentially
1315     independent of the electrostatic method utilized in the simulation.
1316 mlamichh 4397
1317 gezelter 4433 There is a notable difference in the dependence on $\alpha$ for the
1318     quadrupolar correction compared with the dipolar correction. This is
1319     due to the reduced range of the quadrupole-quadrupole interaction when
1320     compared with dipolar interactions. The effects of the Gaussian
1321     damping for dipoles are significant near the cutoff radius, which can
1322     be observed in Fig. \ref{fig:kirkwood}, while for quadrupoles, most of
1323 gezelter 4434 the interaction is naturally diminished by that point. Because
1324     overdamping can obscure orientational preferences, quadrupolar fluids
1325     can be safely simulated with smaller values of $\alpha$ than a similar
1326 gezelter 4433 dipolar fluid.
1327    
1328 mlamichh 4397 \begin{figure}
1329 gezelter 4428 \includegraphics[width=\linewidth]{screen_all.eps}
1330 gezelter 4401 \caption{\label{fig:screenQuad} The distance-dependent screening
1331     factor, $\epsilon(r)$, between two ions immersed in the quadrupolar
1332 gezelter 4434 fluid. Results from the perturbation and fluctuation methods are
1333 gezelter 4432 shown in left and central panels. Here the susceptibility is
1334 gezelter 4434 calculated from the bulk simulations and the geometrical factor is
1335     evaluated using Eq. (\ref{eq:geometricalFactor}) using the field and
1336     field-gradient produced by the two ions. The right hand panel shows
1337     the screening factor obtained from the PMF calculations.}
1338 mlamichh 4397 \end{figure}
1339 gezelter 4401 A more difficult test of the quadrupolar susceptibility is made by
1340     comparing with direct calculation of the electrostatic screening using
1341     the potential of mean force (PMF). Since the effective dielectric
1342     constant for a quadrupolar fluid depends on the geometry of the field
1343     and field gradient, this is not a physical property of the quadrupolar
1344     fluid.
1345 mlamichh 4397
1346 gezelter 4401 The geometrical factor for embedded ions changes with the ion
1347     separation distance. It is therefore reasonable to treat the
1348     dielectric constant as a distance-dependent screening factor. Since
1349     the quadrupolar molecules couple with the gradient of the field, the
1350     distribution of the quadrupoles will be inhomogeneously distributed
1351     around the point charges. Hence the distribution of quadrupolar
1352     molecules should be taken into account when computing the geometrical
1353     factors in the presence of this perturbation,
1354     \begin{eqnarray}
1355     G &=& \frac{\int_V g(\mathbf{r}) \left|\nabla \mathbf{E}^\circ
1356     \right|^2 d\mathbf{r}}{\int_V\left|\mathbf{E}^\circ\right|^2
1357     d\mathbf{r}} \nonumber \\ \nonumber \\
1358     &=& \frac{ 2 \pi \int_{-1}^{1} \int_{0}^{R} r^2 g(r,
1359 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}}
1360 gezelter 4401 \label{eq:geometricalFactor}
1361     \end{eqnarray}
1362     where $g(r,\cos\theta)$ is a distribution function for the quadrupoles
1363     with respect to an origin at midpoint of a line joining the two probe
1364     charges.
1365    
1366     The effective screening factor is plotted against ion separation
1367     distance in Fig. \ref{fig:screenQuad}. The screening evaluated from
1368 gezelter 4433 the perturbation and fluctuation methods are shown in the left and
1369     central panels. Here the susceptibilities are calculated from bulk
1370     fluid simulations and the geometrical factors are evaluated using the
1371 gezelter 4434 field and field gradients produced by the ions. The field gradients
1372 gezelter 4433 have been weighted by the $g(r, \cos\theta)$ from the PMF calculations
1373     (Eq. (\ref{eq:geometricalFactor})). The right hand panel shows the
1374     screening factor obtained directly from the PMF calculations.
1375 gezelter 4401
1376     We note that the screening factor obtained from both the perturbation
1377 gezelter 4433 and fluctuation methods are in good agreement with each other at
1378     similar values of $\alpha$, and agree with Ewald for
1379     $\alpha = 0.2$~\AA$^{-1}$. The magnitude of these screening factors
1380     depends strongly on the $g(r, \cos\theta)$ weighting originating in
1381     the PMF calculations.
1382 gezelter 4401
1383 gezelter 4433 In Ewald-based simulations, the PMF calculations include interactions
1384     between periodic replicas of the ions, and there is a significant
1385     reduction in the screening factor because of this effect. Because the
1386     real-space methods do not include coupling to periodic replicas, both
1387     the magnitude and distance-dependent decay of the PMF are
1388     significantly larger. For moderate damping ($\alpha \sim 0.2-0.3$
1389     \AA$^{-1}$), screening factors for GSF, TSF, and SP are converging to
1390     similar values at large ion separations, and this value is the same as
1391     the large-separation estimate from the perturbation and fluctuation
1392     simulations for $\alpha \sim 0.2$ \AA$^{-1}$. The PMF calculations
1393 gezelter 4434 also show signs of coalescence of the ion solvation shells at
1394 gezelter 4433 separations smaller than 7~\AA. At larger separations, the
1395     $\alpha = 0.2$~\AA$^{-1}$ PMF calculations appear to be reproducing
1396     the bulk screening values. These results suggest that using either
1397     TSF or GSF with moderate damping is a relatively safe way to predict
1398     screening in quadrupolar fluids.
1399    
1400 gezelter 4404 \section{Conclusions}
1401     We have used both perturbation and fluctuation approaches to evaluate
1402     dielectric properties for dipolar and quadrupolar fluids. The static
1403     dielectric constant is the relevant bulk property for dipolar fluids,
1404     while the quadrupolar susceptibility plays a similar role for
1405     quadrupoles. Corrections to both the static dielectric constant and
1406     the quadrupolar susceptibility were derived for three new real space
1407     electrostatic methods, and these corrections were tested against a
1408     third measure of dielectric screening, the potential of mean force
1409     between two ions immersed in the fluids.
1410 gezelter 4401
1411 gezelter 4404 For the dipolar fluids, we find that the polarizability evaluated
1412     using the perturbation and fluctuation methods show excellent
1413     agreement, indicating that equilibrium calculations of the dipole
1414 gezelter 4435 fluctuations are good measures of bulk polarizability.
1415 mlamichh 4353
1416 gezelter 4435 One of the findings of the second paper in this series is that the
1417     moderately damped GSF and SP methods were most suitable for molecular
1418     dynamics and Monte Carlo simulations, respectively.\cite{PaperII} Our
1419     current results show that dielectic properties like $\epsilon$ and
1420     $G_K(r)$ are sensitive probes of local treatment of electrostatic
1421     damping for the new real space methods, as well as for the Ewald
1422     sum. Choosing a Gaussian damping parameter ($\alpha$) in a reasonable
1423     range is therefore essential for obtaining agreement between the
1424     electrostatic methods. A physical explanation of this rests on the
1425     local orientational preferences of other molecules around a central
1426     dipole. The orientational contributions to dipolar interactions are
1427     weighted by two radial functions ($v_{21}(r)$ and $v_{22}(r)$). The
1428     relative magnitudes of these functions, and therefore the
1429     orientational preferences of local dipoles, are quite sensitive to the
1430     value of $\alpha$. With moderate damping, the ratio approaches the
1431     orientational preferences of Ewald-based simulations, removing the
1432     ``hole'' in $G_K(r)$ for underdamped SP and GSF simulations
1433     (Fig. \ref{fig:kirkwood}).
1434 mlamichh 4353
1435 gezelter 4435 The derived correction formulae can approximate bulk properties from
1436     non-optimal parameter choices, as long as the methods are used in a
1437     relatively ``safe'' range of damping. The newly-derived entries in
1438     table \ref{tab:A} can provide an effective minimum on the values of
1439     $\alpha$ that should be used in simulations. With a cutoff radius of
1440     12 \AA, $\alpha = 0.241$ \AA$^{-1}$ (SP) or $0.268$ \AA$^{-1}$ (GSF)
1441     would capture dielectric screening with reasonable fidelity. The
1442     sensitivity of the dielectric screening is also observed in the
1443     effective screening of ions embedded in the fluid.
1444    
1445 gezelter 4432 With good choices of $\alpha$, the dielectric constant evaluated using
1446     the computed polarizability and correction factors agrees well with
1447     the previous Ewald-based simulation results.\cite{Adams81,NeumannI83}
1448 gezelter 4404 Although the TSF method alters many dynamic and structural features in
1449     multipolar liquids,\cite{PaperII} it is surprisingly good at computing
1450     bulk dielectric properties at nearly all ranges of the damping
1451     parameter. In fact, the correction factor, $A = 1$, for the TSF
1452     method so the conducting boundary formula is essentially correct when
1453     using this method for point dipolar fluids.
1454 mlamichh 4397
1455 gezelter 4404 As in the dipolar case, the quadpole polarizability evaluated from
1456 gezelter 4433 both perturbation and fluctuation simulations show good agreement,
1457     again confirming that equilibrium fluctuation calculations are
1458     sufficient to reproduce bulk dielectric properties in these fluids.
1459     The quadrupolar susceptibility calculated via our derived correction
1460 gezelter 4434 factors produces similar results for all three real space
1461 gezelter 4433 methods. Similarly, with good choices of the damping parameter, the
1462     screening factor calculated using the susceptibility and a weighted
1463     geometric factor provides good agreement with results obtained
1464     directly via potentials of mean force. For quadrupolar fluids, the
1465     distance dependence of the electrostatic interaction is significantly
1466     reduced and the correction factors are all small. These points
1467     suggest that how an electrostatic method treats the cutoff radius
1468     become less consequential for higher order multipoles.
1469 gezelter 4404
1470 gezelter 4434 For this reason, our recommendation is that the moderately-damped
1471 gezelter 4439 ($\alpha = 0.25-0.27$~\AA$^{-1}$) GSF method is a good choice for
1472 gezelter 4434 molecular dynamics simulations where point-multipole interactions are
1473 gezelter 4435 being utilized to compute bulk dielectric properties of fluids.
1474 mlamichh 4353
1475 gezelter 4439 \section*{Supplementary Material}
1476     See supplementary material for information on interactions with
1477     spatially varying fields, Boltzmann averages, self-contributions from
1478     quadrupoles, and corrections to distance-dependent Kirkwood factors.
1479 gezelter 4435
1480 gezelter 4439 \begin{acknowledgments}
1481     Support for this project was provided by the National Science Foundation
1482     under grant CHE-1362211. Computational time was provided by the
1483     Center for Research Computing (CRC) at the University of Notre
1484     Dame. The authors would like to thank the reviewer for helpful
1485     comments and suggestions.
1486     \end{acknowledgments}
1487    
1488    
1489 mlamichh 4412 \appendix
1490 gezelter 4413 \section{Contraction of the quadrupolar tensor with the traceless
1491     quadrupole moment }
1492 mlamichh 4412 \label{ap:quadContraction}
1493 gezelter 4413 For quadrupolar liquids modeled using point quadrupoles, the
1494     interaction tensor is shown in Eq. (\ref{quadRadial}). The Fourier
1495     transformation of this tensor for $ \mathbf{k} = 0$ is,
1496 mlamichh 4412 \begin{equation}
1497     \tilde{T}_{\alpha\beta\gamma\delta}(0) = \int_V T_{\alpha\beta\gamma\delta}(\mathbf{r}) d \mathbf{r}
1498     \end{equation}
1499 gezelter 4413 On the basis of symmetry, the 81 elements can be placed in four
1500     different groups: $\tilde{T}_{aaaa}$, $\tilde{T}_{aaab}$,
1501     $\tilde{T}_{aabb}$, and $\tilde{T}_{aabc}$, where $a$, $b$, and $c$,
1502 gezelter 4414 and can take on distinct values from the set $\left\{x, y, z\right\}$.
1503     The elements belonging to each of these groups can be obtained using
1504     permutations of the indices. Integration of all of the elements shows
1505     that only the groups with indices ${aaaa}$ and ${aabb}$ are non-zero.
1506 gezelter 4413
1507     We can derive values of the components of $\tilde{T}_{aaaa}$ and
1508     $\tilde{T}_{aabb}$ as follows;
1509 mlamichh 4412 \begin{eqnarray}
1510 gezelter 4413 \tilde{T}_{xxxx}(0) &=&
1511 mlamichh 4412 \int_{\textrm{V}}
1512 gezelter 4432 \left[ 3v_{41}(r)+6x^2v_{42}(r)/r^2 + x^4\,v_{43}(r)/r^4 \right] d\mathbf{r} \nonumber \\
1513 mlamichh 4412 &=&12\pi \int_0^{r_c}
1514 gezelter 4414 \left[ v_{41}(r)+\frac{2}{3} v_{42}(r) + \frac{1}{15}v_{43}(r) \right] r^2\,dr =
1515 gezelter 4413 \mathrm{12 \pi B}
1516 mlamichh 4412 \end{eqnarray}
1517     and
1518     \begin{eqnarray}
1519 gezelter 4413 \tilde{T}_{xxyy}(0)&=&
1520     \int_{\textrm{V}}
1521 gezelter 4432 \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 \\
1522 gezelter 4413 &=&4\pi \int_0^{r_c}
1523 gezelter 4414 \left[ v_{41}(r)+\frac{2}{3} v_{42}(r) + \frac{1}{15}v_{43}(r) \right] r^2\,dr =
1524 gezelter 4413 \mathrm{4 \pi B}.
1525 mlamichh 4412 \end{eqnarray}
1526 gezelter 4413 These integrals yield the same values for all permutations of the
1527 gezelter 4432 indices in both tensor element groups. In Eq.
1528 gezelter 4413 \ref{fourierQuadZeroK}, for a particular value of the quadrupolar
1529     polarization $\tilde{\Theta}_{aa}$ we can contract
1530     $\tilde{T}_{aa\gamma\delta}(0)$ with $\tilde{\Theta}_{\gamma\delta}$,
1531     using the traceless properties of the quadrupolar moment,
1532 mlamichh 4412 \begin{eqnarray}
1533 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 \\
1534     &=& 12 \pi \mathrm{B}\tilde{\Theta}_{xx}(0) +
1535     4 \pi \mathrm{B}\tilde{\Theta}_{yy}(0) +
1536     4 \pi \mathrm{B}\tilde{\Theta}_{zz}(0) \nonumber \\
1537     &=& 8 \pi \mathrm{B}\tilde{\Theta}_{xx}(0) + 4 \pi
1538     \mathrm{B}\left(\tilde{\Theta}_{xx}(0)+\tilde{\Theta}_{yy}(0) +
1539     \tilde{\Theta}_{zz}(0)\right) \nonumber \\
1540     &=& 8 \pi \mathrm{B}\tilde{\Theta}_{xx}(0)
1541 mlamichh 4412 \end{eqnarray}
1542 gezelter 4413 Similarly for a quadrupolar polarization $\tilde{\Theta}_{xy}$ in
1543 gezelter 4432 Eq. \ref{fourierQuadZeroK}, we can contract
1544 gezelter 4413 $\tilde{T}_{xy\gamma\delta}(0)$ with $\tilde{\Theta}_{\gamma\delta}$,
1545     using the only surviving terms of the tensor,
1546 mlamichh 4412 \begin{eqnarray}
1547 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 \\
1548     &=& 4 \pi \mathrm{B}\tilde{\Theta}_{xy}(0) +
1549     4 \pi \mathrm{B}\tilde{\Theta}_{yx}(0) \nonumber \\
1550     &=& 8 \pi \mathrm{B}\tilde{\Theta}_{xy}(0)
1551 mlamichh 4412 \end{eqnarray}
1552 gezelter 4413 Here, we have used the symmetry of the quadrupole tensor to combine
1553     the symmetric terms. Therefore we can write matrix contraction for
1554     $\tilde{T}_{\alpha\beta\gamma\delta}(\mathrm{0})$ and
1555     $ \tilde{\Theta}_{\gamma\delta}(\mathrm{0})$ in a general form,
1556 mlamichh 4412 \begin{equation}
1557 gezelter 4413 \tilde{T}_{\alpha\beta\gamma\delta}(\mathrm{0})\tilde{\Theta}_{\gamma\delta}(\mathrm{0})
1558     = 8 \pi \mathrm{B} \tilde{\Theta}_{\alpha\beta}(\mathrm{0}),
1559 mlamichh 4412 \label{contract}
1560     \end{equation}
1561 gezelter 4413 which is the same as Eq. (\ref{quadContraction}).
1562 mlamichh 4412
1563 gezelter 4413 When the molecular quadrupoles are represented by point charges, the
1564     symmetry of the quadrupolar tensor is same as for point quadrupoles
1565     (see Eqs.~\ref{quadCharge} and \ref{quadRadial}). However, for
1566     molecular quadrupoles represented by point dipoles, the symmetry of
1567     the quadrupolar tensor must be handled separately (compare
1568 gezelter 4432 Eqs.~\ref{quadDip} and~\ref{quadRadial}). Although there is a
1569 gezelter 4414 difference in symmetry, the final result (Eq.~\ref{contract}) also holds
1570     true for dipolar representations.
1571 gezelter 4413
1572 gezelter 4414 \section{Quadrupolar correction factor for the Ewald-Kornfeld (EK)
1573     method}
1574     The interaction tensor between two point quadrupoles in the Ewald
1575     method may be expressed,\cite{Smith98,NeumannII83}
1576     \begin{align}
1577     {T}_{\alpha\beta\gamma\delta}(\mathbf{r}) = &\frac{4\pi}{V
1578     }\sum_{k\neq0}^{\infty}
1579     e^{-k^2 / 4
1580     \kappa^2} e^{-i\mathbf{k}\cdot
1581     \mathbf{r}} \left(\frac{r_\alpha r_\beta k_\delta k_\gamma}{k^2}\right) \nonumber \\
1582     &+ \left(\delta_{\alpha\beta}\delta_{\gamma\delta}+\delta_{\alpha\gamma}\delta_{\beta\delta}+\delta_{\alpha\delta}\delta_{\beta\gamma}\right)
1583     B_2(r) \nonumber \\
1584     &- \left(\delta_{\gamma\delta} r_\alpha r_\beta + \mathrm{ 5\; permutations}\right) B_3(r) \nonumber \\
1585     &+ \left(r_\alpha r_\beta r_\gamma r_\delta \right) B_4(r)
1586 mlamichh 4412 \label{ewaldTensor}
1587 gezelter 4414 \end{align}
1588     where $B_n(r)$ are radial functions defined in reference
1589     \onlinecite{Smith98},
1590     \begin{align}
1591     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) \\
1592     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) \\
1593     B_4(r) =& \frac{105}{r^9}\left(\frac{2r\kappa e^{-r^2
1594     \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}}
1595     + \frac{16r^7\kappa^7 e^{-r^2 \kappa^2}}{105\sqrt{\pi}} + \mathrm{erfc(\kappa r)} \right)
1596     \end{align}
1597    
1598     We can divide ${T}_{\alpha\beta\gamma\delta}(\mathbf{r})$ into three
1599     parts:
1600 mlamichh 4412 \begin{eqnarray}
1601 gezelter 4414 & & \mathbf{T}(\mathbf{r}) =
1602     \mathbf{T}^\mathrm{K}(\mathbf{r}) +
1603     \mathbf{T}^\mathrm{R1}(\mathbf{r}) +
1604     \mathbf{T}^\mathrm{R2}(\mathbf{r})
1605 mlamichh 4412 \end{eqnarray}
1606 gezelter 4414 where the first term is the reciprocal space portion. Since the
1607     quadrupolar correction factor $B = \tilde{T}_{abab}(0) / 4\pi$ and
1608 gezelter 4415 $\mathbf{k} = 0 $ is excluded from the reciprocal space sum,
1609 gezelter 4414 $\mathbf{T}^\mathrm{K}$ will not contribute.\cite{NeumannII83} The
1610     remaining terms,
1611     \begin{equation}
1612     \mathbf{T}^\mathrm{R1}(\mathbf{r}) = \mathbf{T}^\mathrm{bare}(\mathbf{r}) \left(\frac{2r\kappa e^{-r^2
1613     \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}}
1614     + \frac{16r^7\kappa^7 e^{-r^2 \kappa^2}}{105\sqrt{\pi}} + \mathrm{erfc(\kappa r)} \right)
1615     \end{equation}
1616 mlamichh 4412 and
1617     \begin{eqnarray}
1618 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 \\
1619     &-&\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)
1620 mlamichh 4412 \end{eqnarray}
1621 gezelter 4414 are contributions from the real space
1622     sum.\cite{Adams76,Adams80,Adams81} Here
1623     $\mathbf{T}^\mathrm{bare}(\mathbf{r})$ is the unmodified quadrupolar
1624     tensor (for undamped quadrupoles). Due to the angular symmetry of the
1625     unmodified tensor, the integral of
1626     $\mathbf{T}^\mathrm{R1}(\mathbf{r})$ will vanish when integrated over
1627     a spherical region. The only term contributing to the correction
1628     factor (B) is therefore
1629     $T^\mathrm{R2}_{\alpha\beta\gamma\delta}(\mathbf{r})$, which allows us
1630     to derive the correction factor for the Ewald-Kornfeld (EK) method,
1631 mlamichh 4412 \begin{eqnarray}
1632 gezelter 4414 \mathrm{B} &=& \frac{1}{4\pi} \int_V T^\mathrm{R2}_{abab}(\mathbf{r}) \nonumber \\
1633 gezelter 4432 &=& -\frac{8r_c^3 \kappa^5 e^{-\kappa^2 r_c^2}}{15\sqrt{\pi}}.
1634 mlamichh 4412 \end{eqnarray}
1635 gezelter 4428
1636    
1637 mlamichh 4353 \newpage
1638 mlamichh 4397 \bibliography{dielectric_new}
1639 mlamichh 4353
1640     \end{document}