ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/multipole/dielectric_new.tex
Revision: 4433
Committed: Mon Jul 11 13:23:51 2016 UTC (9 years, 2 months ago) by gezelter
Content type: application/x-tex
File size: 84691 byte(s)
Log Message:
Almost done

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     dielectric constant of simulated materials.
105 gezelter 4396
106 gezelter 4399 If we consider a system of dipolar or quadrupolar molecules under the
107     influence of an external field or field gradient, the net polarization
108     of the system will largely be proportional to the applied
109 gezelter 4417 perturbation.\cite{Chitanvis96, Stern03, SalvchovI14, SalvchovII14} In
110     simulations, the net polarization of the system is also determined by
111     the interactions \textit{between} the molecules. Therefore the
112 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     for truncation of the field and field gradient contributions from
141     other multipoles. We find that the correction formulae for dipolar
142     molecules depends not only on the methodology being used, but also on
143     whether the molecular dipoles are treated using point charges or point
144     dipoles. We derive correction factors for both cases.
145 gezelter 4399
146 gezelter 4392 In quadrupolar fluids, the relationship between quadrupolar
147     susceptibility and the dielectric constant is not as straightforward
148 gezelter 4399 as in the dipolar case. The effective dielectric constant depends on
149     the geometry of the external (or internal) field
150     perturbation.\cite{Ernst92} Significant efforts have been made to
151     increase our understanding the dielectric properties of these
152 gezelter 4400 fluids,\cite{JeonI03,JeonII03,Chitanvis96} although a general
153     correction formula has not yet been developed.
154 gezelter 4390
155 gezelter 4392 In this paper we derive general formulae for calculating the
156 gezelter 4400 quadrupolar susceptibility of quadrupolar fluids. We also evaluate the
157     correction factor for SP, GSF, and TSF methods for quadrupolar fluids
158     interacting via point charges, point dipoles or directly through
159     quadrupole-quadrupole interactions.
160 gezelter 4390
161 gezelter 4400 We also calculate the screening behavior for two ions immersed in
162     multipolar fluids to estimate the distance dependence of charge
163     screening in both dipolar and quadrupolar fluids. We use three
164     distinct methods to compare our analytical results with computer
165 gezelter 4419 simulations (see Fig. \ref{fig:schematic}):
166 gezelter 4390 \begin{enumerate}
167 gezelter 4396 \item responses of the fluid to external perturbations,
168 gezelter 4399 \item fluctuations of system multipole moments, and
169 gezelter 4396 \item potentials of mean force between solvated ions,
170 gezelter 4390 \end{enumerate}
171    
172 gezelter 4408 \begin{figure}
173 gezelter 4428 \includegraphics[width=\linewidth]{Schematic.eps}
174 gezelter 4408 \caption{Dielectric properties of a fluid measure the response to
175 gezelter 4411 external electric fields and gradients (left), or internal fields
176     and gradients generated by the molecules themselves (center), or
177     fields produced by embedded ions (right). The dielectric constant
178     ($\epsilon$) measures all three responses in dipolar fluids (top).
179     In quadrupolar liquids (bottom), the relevant bulk property is the
180     quadrupolar susceptibility ($\chi_Q$), and the geometry of the field
181 gezelter 4408 determines the effective dielectric screening.}
182     \label{fig:schematic}
183     \end{figure}
184    
185 gezelter 4400 Under the influence of weak external fields, the bulk polarization of
186     the system is primarily a linear response to the perturbation, where
187     proportionality constant depends on the electrostatic interactions
188     between the multipoles. The fluctuation formulae connect bulk
189     properties of the fluid to equilibrium fluctuations in the system
190     multipolar moments during a simulation. These fluctuations also depend
191     on the form of the electrostatic interactions between molecules.
192     Therefore, the connections between the actual bulk properties and both
193     the computed fluctuation and external field responses must be modified
194     accordingly.
195 gezelter 4390
196 gezelter 4400 The potential of mean force (PMF) allows calculation of an effective
197 gezelter 4399 dielectric constant or screening factor from the potential energy
198     between ions before and after dielectric material is introduced.
199 gezelter 4400 Computing the PMF between embedded point charges is an additional
200     check on the bulk properties computed via the other two methods.
201 mlamichh 4363
202 gezelter 4400 \section{The Real-Space Methods}
203    
204 gezelter 4404 In the first paper in this series, we derived interaction energies, as
205     well as expressions for the forces and torques for point multipoles
206     interacting via three new real-space methods.\cite{PaperI} The Taylor
207     shifted-force (TSF) method modifies the electrostatic kernel,
208     $f(r) = 1/r$, so that all forces and torques go smoothly to zero at
209     the cutoff radius,
210     \begin{equation}
211     U^{\text{TSF}}_{ab} = M_{a} M_{b} f_n(r).
212     \label{TSF}
213     \end{equation}
214     Here the multipole operator for site $a$, $M_{a}$, is expressed in
215     terms of the point charge, $C_{a}$, dipole, ${\bf D}_{a}$, and
216     quadrupole, $\mathsf{Q}_{a}$, for object $a$, etc. Because each of
217 gezelter 4432 the multipole operators includes gradient operators (one for a dipole,
218     two for a quadrupole, \textit{etc.}), an approximate electrostatic
219     kernel, $f_n(r)$ is Taylor-expanded around the cutoff radius, so that
220     $n + 1$ derivatives vanish as $r \rightarrow r_c$. This ensures
221     smooth convergence of the energy, forces, and torques as molecules
222     leave and reenter each others cutoff spheres. The order of the Taylor
223     expansion is determined by the multipolar order of the interaction.
224     That is, smooth quadrupole-quadrupole forces require the fifth
225     derivative to vanish at the cutoff radius, so the appropriate function
226     Taylor expansion will be of fifth order.
227 gezelter 4404
228     Following this procedure results in separate radial functions for each
229     of the distinct orientational contributions to the potential. For
230     example, in dipole-dipole interactions, the direct dipole dot product
231     ($\mathbf{D}_{a} \cdot \mathbf{D}_{b}$) is treated differently than
232     the dipole-distance dot products:
233     \begin{equation}
234     U_{\mathbf{D}_{a}\mathbf{D}_{b}}(r)= -\frac{1}{4\pi \epsilon_0} \left[ \left(
235     \mathbf{D}_{a} \cdot
236     \mathbf{D}_{b} \right) v_{21}(r) +
237     \left( \mathbf{D}_{a} \cdot \hat{\mathbf{r}} \right)
238     \left( \mathbf{D}_{b} \cdot \hat{\mathbf{r}} \right) v_{22}(r) \right]
239     \end{equation}
240 gezelter 4419 In standard electrostatics, the two radial functions, $v_{21}(r)$ and
241 gezelter 4432 $v_{22}(r)$, are proportional to $1/r^3$, but they have distinct radial
242 gezelter 4419 dependence in the TSF method. Careful choice of these functions makes
243     the forces and torques vanish smoothly as the molecules drift beyond
244     the cutoff radius (even when those molecules are in different
245     orientations).
246 gezelter 4404
247     A second and somewhat simpler approach involves shifting the gradient
248     of the Coulomb potential for each particular multipole order,
249     \begin{equation}
250     U^{\text{GSF}}_{ab} = \sum \left[ U(\mathbf{r}, \mathsf{A}, \mathsf{B}) -
251     U(r_c \hat{\mathbf{r}},\mathsf{A}, \mathsf{B}) - (r-r_c)
252     \hat{\mathbf{r}} \cdot \nabla U(r_c \hat{\mathbf{r}},\mathsf{A},
253     \mathsf{B}) \right]
254     \label{generic2}
255     \end{equation}
256     where the sum describes a separate force-shifting that is applied to
257     each orientational contribution to the energy, i.e. $v_{21}$ and
258     $v_{22}$ are shifted separately. In this expression,
259     $\hat{\mathbf{r}}$ is the unit vector connecting the two multipoles
260     ($a$ and $b$) in space, and $\mathsf{A}$ and $\mathsf{B}$ represent
261     the orientations the multipoles. Because this procedure is equivalent
262     to using the gradient of an image multipole placed at the cutoff
263     sphere for shifting the force, this method is called the gradient
264     shifted-force (GSF) approach.
265    
266     Both the TSF and GSF approaches can be thought of as multipolar
267     extensions of the original damped shifted-force (DSF) approach that
268     was developed for point charges. There is also a multipolar extension
269     of the Wolf sum that is obtained by projecting an image multipole onto
270     the surface of the cutoff sphere, and including the interactions with
271     the central multipole and the image. This effectively shifts only the
272     total potential to zero at the cutoff radius,
273     \begin{equation}
274     U^{\text{SP}}_{ab} = \sum \left[ U(\mathbf{r}, \mathsf{A}, \mathsf{B}) -
275     U(r_c \hat{\mathbf{r}},\mathsf{A}, \mathsf{B}) \right]
276     \label{eq:SP}
277     \end{equation}
278     where the sum again describes separate potential shifting that is done
279     for each orientational contribution to the energy. The potential
280     energy between a central multipole and other multipolar sites goes
281     smoothly to zero as $r \rightarrow r_c$, but the forces and torques
282     obtained from this shifted potential (SP) approach are discontinuous
283     at $r_c$.
284    
285     All three of the new real space methods share a common structure: the
286     various orientational contributions to multipolar interaction energies
287     require separate treatment of their radial functions, and these are
288     tabulated for both the raw Coulombic kernel ($1/r$) as well as the
289     damped kernel ($\mathrm{erfc}(\alpha r)/r$), in the first paper of this
290     series.\cite{PaperI} The second paper in this series evaluated the
291     fidelity with which the three new methods reproduced Ewald-based
292     results for a number of model systems.\cite{PaperII} One of the major
293     findings was that moderately-damped GSF simulations produced nearly
294     identical behavior with Ewald-based simulations, but the real-space
295     methods scale linearly with system size.
296    
297 gezelter 4399 \section{Dipolar Fluids and the Dielectric Constant}
298 mlamichh 4353
299 gezelter 4399 Dielectric properties of a fluid arise mainly from responses of the
300     fluid to either applied fields or transient fields internal to the
301     fluid. In response to an applied field, the molecules have electronic
302 gezelter 4432 polarizabilities, changes to internal bond lengths and angles, and
303     reorientations towards the direction of the applied field. There is an
304     added complication that in the presence of external field, the
305     perturbation experienced by any single molecule is not only due to
306     external field but also to the fields produced by the all other
307     molecules in the system.
308 mlamichh 4353
309 gezelter 4399 \subsection{Response to External Perturbations}
310    
311     In the presence of uniform electric field $\mathbf{E}$, an individual
312     molecule with a permanent dipole moment $p_o$ will realign along the
313     direction of the field with an average polarization given by
314 mlamichh 4353 \begin{equation}
315 gezelter 4400 \braket{\mathbf{p}} = \epsilon_0 \alpha_p \mathbf{E},
316 mlamichh 4353 \end{equation}
317 gezelter 4400 where $\alpha_p = {p_o}^2 / 3 \epsilon_0 k_B T$ is the contribution to
318     molecular polarizability due solely to reorientation dynamics.
319     Because the applied field must overcome thermal motion, the
320     orientational polarization depends inversely on the temperature.
321 mlamichh 4353
322 gezelter 4432 A condensed phase system of permanent dipoles will also polarize along
323     the direction of an applied field. The polarization density of the
324     system is
325 mlamichh 4353 \begin{equation}
326 gezelter 4432 \textbf{P} = \epsilon_o \chi_{D} \mathbf{E},
327 mlamichh 4353 \label{pertDipole}
328     \end{equation}
329 gezelter 4432 where the constant $\chi_D$ is the dipole susceptibility, which is an
330     emergent property of the dipolar fluid, and is the quantity most
331     directly related to the static dielectric constant,
332     $\epsilon = 1 + \chi_D$.
333 mlamichh 4363
334 gezelter 4399 \subsection{Fluctuation Formula}
335 gezelter 4400 % Using the Ewald sum under tin-foil boundary conditions, $\epsilon$ can
336     % be calculated for systems of orientationally-polarized molecules,
337     % \begin{equation}
338     % \epsilon = 1 + \frac{\langle M^2\rangle}{3\epsilon_0k_\textrm{B}TV},
339     % \label{eq:staticDielectric}
340     % \end{equation}
341     % where $\epsilon_0$ is the permittivity of free space and
342     % $\langle M^2\rangle$ is the fluctuation of the system dipole
343     % moment.\cite{Allen89} The numerator in the fractional term in
344     % equation (\ref{eq:staticDielectric}) is identical to the quantity
345     % calculated in the finite-system Kirkwood $g$ factor ($G_k$):
346     % \begin{equation}
347     % G_k = \frac{\langle M^2\rangle}{N\mu^2},
348     % \label{eq:KirkwoodFactor}
349     % \end{equation}
350     % where $\mu$ is the dipole moment of a single molecule of the
351     % homogeneous system.\cite{NeumannI83,NeumannII83,Neumann84,Neumann85} The
352     % fluctuation term in both equation (\ref{eq:staticDielectric}) and
353     % (\ref{eq:KirkwoodFactor}) is calculated as follows,
354     % \begin{equation}
355     % \begin{split}
356     % \langle M^2\rangle &= \langle\bm{M}\cdot\bm{M}\rangle
357     % - \langle\bm{M}\rangle\cdot\langle\bm{M}\rangle \\
358     % &= \langle M_x^2+M_y^2+M_z^2\rangle
359     % - (\langle M_x\rangle^2 + \langle M_x\rangle^2
360     % + \langle M_x\rangle^2).
361     % \end{split}
362     % \label{eq:fluctBoxDipole}
363     % \end{equation}
364     % This fluctuation term can be accumulated during a simulation; however,
365     % it converges rather slowly, thus requiring multi-nanosecond simulation
366     % times.\cite{Horn04} In the case of tin-foil boundary conditions, the
367     % dielectric/surface term of the Ewald summation is equal to zero.
368 mlamichh 4353
369 gezelter 4400 For a system of dipolar molecules at thermal equilibrium, we can
370     define both a system dipole moment, $\mathbf{M} = \sum_i \mathbf{p}_i$
371     as well as a dipole polarization density,
372 gezelter 4432 $\mathbf{P} = \braket{\mathbf{M}}/V$. The polarization density can be
373     expressed approximately in terms of fluctuations in the net dipole
374 gezelter 4400 moment,
375 mlamichh 4353 \begin{equation}
376 gezelter 4400 \mathbf{P} = \epsilon_o \frac{\braket{\mathbf{M}^2}-{\braket{\mathbf{M}}}^2}{3 \epsilon_o V k_B T}\bf{E}
377 gezelter 4399 \label{flucDipole}
378 mlamichh 4353 \end{equation}
379 gezelter 4400 This has structural similarity with the Boltzmann average for the
380     polarization of a single molecule. Here
381     $ \braket{\mathbf{M}^2}-{\braket{\mathbf{M}}}^2$ measures fluctuations
382 gezelter 4401 in the net dipole moment,
383 mlamichh 4353 \begin{equation}
384 gezelter 4401 \langle \mathbf{M}^2 \rangle - \langle \mathbf{M} \rangle^2 =
385     \langle M_x^2+M_y^2+M_z^2 \rangle - \left( \langle M_x \rangle^2 +
386     \langle M_y \rangle^2 +
387     \langle M_z \rangle^2 \right).
388     \label{eq:flucDip}
389     \end{equation}
390 gezelter 4432 When no applied electric field is present, the ensemble average of
391     both the net dipole moment $\braket{\mathbf{M}}$ and dipolar
392     polarization $\bf{P}$ tends to vanish but $\braket{\mathbf{M}^2}$ does
393     not. The bulk dipole polarizability can therefore be
394     written
395 gezelter 4401 \begin{equation}
396 gezelter 4432 \alpha_D = \frac{\braket{\mathbf{M}^2}-{\braket{\mathbf{M}}}^2}{3 \epsilon_o V k_B T}.
397 mlamichh 4353 \end{equation}
398 gezelter 4432 The susceptibility ($\chi_D$) and bulk polarizability ($\alpha_D$)
399     both measure responses of a dipolar system to an applied field.
400     However, $\chi_D$ includes contributions from local fields due to the
401     dipoles themselves, while $\alpha_D$ is relatively simple to compute
402     from numerical simulations. One of the primary aims of this paper is
403     to provide the connection between the bulk properties
404     ($\epsilon, \chi_D$) and the computed quantities ($\alpha_D$) that
405     have been adapted for the new real-space methods.
406 gezelter 4400
407 gezelter 4399 \subsection{Correction Factors}
408 gezelter 4408 \label{sec:corrFactor}
409 gezelter 4400 In the presence of a uniform external field $ \mathbf{E}^\circ$, the
410     total electric field at $\mathbf{r}$ depends on the polarization
411     density at all other points in the system,\cite{NeumannI83}
412 mlamichh 4353 \begin{equation}
413 gezelter 4400 \mathbf{E}(\mathbf{r}) = \mathbf{E}^\circ(\mathbf{r}) +
414     \frac{1}{4\pi\epsilon_o} \int d\mathbf{r}^\prime
415     \mathbf{T}(\mathbf{r}-\mathbf{r}^\prime)\cdot
416     {\mathbf{P}(\mathbf{r}^\prime)}.
417     \label{eq:localField}
418 mlamichh 4353 \end{equation}
419 gezelter 4400 $\mathbf{T}$ is the dipole interaction tensor connecting dipoles at
420 gezelter 4432 $\mathbf{r}^\prime$ with the point of interest ($\mathbf{r}$), where
421     the integral is done over all space. Because simulations utilize
422     periodic boundary conditions and spherical cutoffs, the integral is
423     carried out over the domain ($0 < r < r_c$).
424 mlamichh 4353
425 gezelter 4400 In simulations of dipolar fluids, the molecular dipoles may be
426     represented either by closely-spaced point charges or by
427 gezelter 4408 point dipoles (see Fig. \ref{fig:tensor}).
428 mlamichh 4353 \begin{figure}
429 gezelter 4428 \includegraphics[width=\linewidth]{Tensors.eps}
430 gezelter 4408 \caption{In the real-space electrostatic methods, the molecular dipole
431     tensor, $\mathbf{T}_{\alpha\beta}(r)$, is not the same for
432     charge-charge interactions as for point dipoles (left panel). The
433     same holds true for the molecular quadrupole tensor (right panel),
434     $\mathbf{T}_{\alpha\beta\gamma\delta}(r)$, which can have distinct
435     forms if the molecule is represented by charges, dipoles, or point
436     quadrupoles.}
437     \label{fig:tensor}
438 mlamichh 4353 \end{figure}
439     In the case where point charges are interacting via an electrostatic
440     kernel, $v(r)$, the effective {\it molecular} dipole tensor,
441     $\mathbf{T}$ is obtained from two successive applications of the
442     gradient operator to the electrostatic kernel,
443 mlamichh 4412 \begin{eqnarray}
444 gezelter 4432 T_{\alpha \beta}(\mathbf{r}) &=& \nabla_\alpha \nabla_\beta
445 mlamichh 4412 \left(v(r)\right) \\
446 gezelter 4416 &=& \delta_{\alpha \beta}
447 mlamichh 4353 \left(\frac{1}{r} v^\prime(r) \right) + \frac{r_{\alpha}
448     r_{\beta}}{r^2} \left( v^{\prime \prime}(r) - \frac{1}{r}
449     v^{\prime}(r) \right)
450     \label{dipole-chargeTensor}
451 mlamichh 4412 \end{eqnarray}
452 mlamichh 4353 where $v(r)$ may be either the bare kernel ($1/r$) or one of the
453     modified (Wolf or DSF) kernels. This tensor describes the effective
454 gezelter 4400 interaction between molecular dipoles ($\mathbf{D}$) in Gaussian units
455     as $-\mathbf{D} \cdot \mathbf{T} \cdot \mathbf{D}$.
456 mlamichh 4353
457 gezelter 4400 When utilizing any of the three new real-space methods for point
458     \textit{dipoles}, the tensor is explicitly constructed,
459 mlamichh 4353 \begin{equation}
460 gezelter 4432 T_{\alpha \beta}(\mathbf{r}) = \delta_{\alpha \beta} v_{21}(r) +
461 mlamichh 4353 \frac{r_{\alpha} r_{\beta}}{r^2} v_{22}(r)
462     \label{dipole-diopleTensor}
463     \end{equation}
464     where the functions $v_{21}(r)$ and $v_{22}(r)$ depend on the level of
465 gezelter 4400 the approximation.\cite{PaperI,PaperII} Although the Taylor-shifted
466     (TSF) and gradient-shifted (GSF) models produce to the same $v(r)$
467     function for point charges, they have distinct forms for the
468     dipole-dipole interaction.
469 mlamichh 4353
470 gezelter 4432 Using the constitutive relation in Eq. (\ref{pertDipole}), the
471     polarization density $\mathbf{P}(\mathbf{r})$ is given by,
472 mlamichh 4353 \begin{equation}
473 gezelter 4400 \mathbf{P}(\mathbf{r}) = \epsilon_o \chi_D
474     \left(\mathbf{E}^\circ(\mathbf{r}) + \frac{1}{4\pi\epsilon_o} \int
475     d\mathbf{r}^\prime \mathbf{T}(\mathbf{r}-\mathbf{r}^\prime ) \cdot \mathbf{P}(\mathbf{r}^\prime)\right).
476 mlamichh 4353 \label{constDipole}
477     \end{equation}
478 gezelter 4400 Note that $\chi_D$ explicitly depends on the details of the dipole
479     interaction tensor. Neumann \textit{et al.}
480     \cite{NeumannI83,NeumannII83,Neumann84,Neumann85} derived an elegant
481     way to modify the fluctuation formula to correct for approximate
482     interaction tensors. This correction was derived using a Fourier
483     representation of the interaction tensor,
484     $\tilde{\mathbf{T}}(\mathbf{k})$, and involves the quantity,
485 mlamichh 4353 \begin{equation}
486 gezelter 4432 \mathsf{A} = \frac{3}{4\pi}\tilde{\mathbf{T}}(0) = \frac{3}{4\pi} \int_V
487     d\mathbf{r} \mathbf{T}(\mathbf{r})
488 mlamichh 4353 \end{equation}
489 gezelter 4432 which is the $k \rightarrow 0$ limit of
490     $\tilde{\mathbf{T}}(\mathbf{r})$. Note that the integration of the
491     dipole tensors, Eqs. (\ref{dipole-chargeTensor}) and
492     (\ref{dipole-diopleTensor}), over spherical volumes yields values only
493     along the diagonal. Additionally, the spherical symmetry of
494     $\mathbf{T}(\mathbf{r})$ insures that all diagonal elements are
495     identical. For this reason, $\mathsf{A}$ can be written as a scalar
496     constant ($A$) multiplying the unit tensor.
497    
498     Using the quantity $A$ (originally called $Q$ in
499 gezelter 4400 refs. \onlinecite{NeumannI83,NeumannII83,Neumann84,Neumann85}), the
500     dielectric constant can be computed
501 mlamichh 4353 \begin{equation}
502 gezelter 4400 \epsilon = \frac{3+(A + 2)(\epsilon_{CB}-1)}{3 + (A -1)(\epsilon_{CB}-1)}
503     \label{correctionFormula}
504 mlamichh 4353 \end{equation}
505 gezelter 4400 where $\epsilon_{CB}$ is the widely-used conducting boundary
506     expression for the dielectric constant,
507 mlamichh 4353 \begin{equation}
508 gezelter 4400 \epsilon_{CB} = 1 + \frac{\braket{\bf{M}^2}-{\braket{\bf{M}}}^2}{3
509 gezelter 4432 \epsilon_o V k_B T} = 1 + \alpha_{D}.
510 gezelter 4408 \label{conductingBoundary}
511 mlamichh 4353 \end{equation}
512 gezelter 4432 Eqs. (\ref{correctionFormula}) and (\ref{conductingBoundary}) allow
513     estimation of the static dielectric constant from fluctuations
514     computed directly from simulations, with the understanding that
515     Eq. (\ref{correctionFormula}) is extraordinarily sensitive when $A$ is
516     far from unity.
517 gezelter 4400
518     % 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}
519     % \begin{equation}
520     % \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}
521     % \label{singular}
522     % \end{equation}
523     % Substituting equation (\ref{singular}) in the equation (\ref{constDipole}) we get,
524     % \begin{equation}
525     % \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).
526     % \end{equation}
527     % For both polarization and electric field homogeneous, this can be easily solved using Fourier transformation,
528     % \begin{equation}
529     % \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}).
530     % \end{equation}
531     % For homogeneous applied field Fourier component is non-zero only if $\kappa = 0$. Therefore,
532     % \begin{equation}
533     % \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}).
534     % \label{fourierDipole}
535     % \end{equation}
536     % 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,
537     % \begin{equation}
538     % \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}
539     % \end{equation}
540     % 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,
541     % \begin{equation}
542     % \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}
543     % \label{correctionFormula}
544     % \end{equation}
545    
546 gezelter 4404 We have utilized the Neumann \textit{et al.} approach for the three
547 gezelter 4400 new real-space methods, and obtain method-dependent correction
548     factors. The expression for the correction factor also depends on
549     whether the simulation involves point charges or point dipoles to
550 gezelter 4425 represent the molecular dipoles. These corrections factors are listed
551     in Table \ref{tab:A}. We note that the GSF correction factor for
552     point dipoles has been independently derived by Stenqvist \textit{et
553     al.}\cite{Stenqvist:2015ph}
554 mlamichh 4353 \begin{table}
555 gezelter 4400 \caption{Expressions for the dipolar correction factor ($A$) for the
556     real-space electrostatic methods in terms of the damping parameter
557 mlamichh 4353 ($\alpha$) and the cutoff radius ($r_c$). The Ewald-Kornfeld result
558 gezelter 4400 derived in Refs. \onlinecite{Adams80,Adams81,NeumannI83} is shown
559     for comparison using the Ewald convergence parameter ($\kappa$)
560     and the real-space cutoff value ($r_c$). }
561 mlamichh 4353 \label{tab:A}
562 gezelter 4416 \begin{tabular}{l|c|c}
563 gezelter 4408 \toprule
564 gezelter 4416 & \multicolumn{2}{c}{Molecular Representation} \\
565 gezelter 4400 Method & point charges & point dipoles \\
566 gezelter 4408 \colrule
567 gezelter 4431 Shifted Potential (SP) & $ \mathrm{erf}(r_c \alpha) - \frac{2
568 gezelter 4411 \alpha r_c}{\sqrt{\pi}} e^{-\alpha^2
569     r_c^2}$ & $\mathrm{erf}(r_c \alpha)
570     -\frac{2 \alpha
571     r_c}{\sqrt{\pi}}\left(1+\frac{2\alpha^2
572     {r_c}^2}{3}
573     \right)e^{-\alpha^2{r_c}^2}
574     $\\ \colrule
575     Gradient-shifted (GSF) & 1 & $\mathrm{erf}(\alpha r_c)-\frac{2
576     \alpha r_c}{\sqrt{\pi}} \left(1 +
577     \frac{2 \alpha^2 r_c^2}{3} +
578     \frac{\alpha^4
579     r_c^4}{3}\right)e^{-\alpha^2 r_c^2} $ \\ \colrule
580 gezelter 4416 Taylor-shifted (TSF) &\multicolumn{2}{c}{1} \\ \colrule
581 gezelter 4432 %Spherical Cutoff (SC)& \multicolumn{2}{c}{$\mathrm{erf}(r_c \alpha) -
582     % \frac{2 \alpha r_c}{\sqrt{\pi}} e^{-\alpha^2
583     % r_c^2}$} \\
584 gezelter 4416 Ewald-Kornfeld (EK) & \multicolumn{2}{c}{$\mathrm{erf}(r_c \kappa) -
585 gezelter 4408 \frac{2 \kappa r_c}{\sqrt{\pi}} e^{-\kappa^2
586     r_c^2}$} \\
587     \botrule
588     \end{tabular}
589 mlamichh 4353 \end{table}
590 gezelter 4400 Note that for point charges, the GSF and TSF methods produce estimates
591     of the dielectric that need no correction, and the TSF method likewise
592     needs no correction for point dipoles.
593 gezelter 4399
594     \section{Quadrupolar Fluids and the Quadrupolar Susceptibility}
595     \subsection{Response to External Perturbations}
596    
597 gezelter 4400 A molecule with a permanent quadrupole, $\mathsf{q}$, will align in
598     the presence of an electric field gradient $\nabla\mathbf{E}$. The
599     anisotropic polarization of the quadrupole is given
600     by,\cite{AduGyamfi78,AduGyamfi81}
601 gezelter 4399 \begin{equation}
602 gezelter 4400 \braket{\mathsf{q}} - \frac{\mathbf{I}}{3}
603     \mathrm{Tr}(\mathsf{q}) = \epsilon_o \alpha_q \nabla\mathbf{E},
604 gezelter 4399 \end{equation}
605 gezelter 4400 where $\alpha_q = q_o^2 / 15 \epsilon_o k_B T $ is a molecular quadrupole
606 gezelter 4431 polarizability and $q_o$ is an effective quadrupole moment for the molecule,
607 gezelter 4400 \begin{equation}
608     q_o^2 = 3 \mathsf{q}:\mathsf{q} - \mathrm{Tr}(\mathsf{q})^2.
609     \end{equation}
610 gezelter 4432 Note that quadrupole calculations involve tensor contractions (double
611     dot products) between rank two tensors, which are defined as
612     \begin{equation}
613     \mathsf{A} \colon \mathsf{B} = \sum_\alpha \sum_\beta
614     A_{\alpha \beta} B_{\beta \alpha}.
615     \end{equation}
616 gezelter 4399
617 gezelter 4400 In the presence of an external field gradient, a system of quadrupolar
618     molecules also organizes with an anisotropic polarization,
619 gezelter 4399 \begin{equation}
620 gezelter 4400 \mathsf{Q} - \frac{\mathbf{I}}{3} \mathrm{Tr}(\mathsf{Q}) = \epsilon_o
621 gezelter 4432 \chi_Q \nabla\mathbf{E}
622 gezelter 4400 \end{equation}
623     where $\mathsf{Q}$ is the traced quadrupole density of the system and
624 gezelter 4432 $\chi_Q$ is a macroscopic quadrupole susceptibility which has
625 gezelter 4400 dimensions of $\mathrm{length}^{-2}$. Equivalently, the traceless form
626     may be used,
627     \begin{equation}
628 gezelter 4432 \mathsf{\Theta} = 3 \epsilon_o \chi_Q \nabla\mathbf{E},
629 gezelter 4399 \label{pertQuad}
630     \end{equation}
631 gezelter 4400 where
632     $\mathsf{\Theta} = 3\mathsf{Q} - \mathbf{I} \mathrm{Tr}(\mathsf{Q})$
633     is the traceless tensor that also describes the system quadrupole
634     density. It is this tensor that will be utilized to derive correction
635     factors below.
636    
637 gezelter 4399 \subsection{Fluctuation Formula}
638 gezelter 4400 As in the dipolar case, we may define a system quadrupole moment,
639     $\mathsf{M}_Q = \sum_i \mathsf{q}_i$ and the traced quadrupolar
640     density, $\mathsf{Q} = \mathsf{M}_Q / V$. A fluctuation formula can
641     be written for a system comprising quadrupolar
642     molecules,\cite{LoganI81,LoganII82,LoganIII82}
643 gezelter 4399 \begin{equation}
644 gezelter 4400 \mathsf{Q} - \frac{\mathbf{I}}{3} \mathrm{Tr}(\mathsf{Q}) = \epsilon_o
645     \frac{\braket{\mathsf{M}_Q^2}-{\braket{\mathsf{M}_Q}}^2}{15 \epsilon_o
646 gezelter 4419 V k_B T} \nabla\mathbf{E}.
647 gezelter 4399 \label{flucQuad}
648     \end{equation}
649 gezelter 4400 Some care is needed in the definitions of the averaged quantities. These
650     refer to the effective quadrupole moment of the system, and they are
651     computed as follows,
652     \begin{align}
653     \braket{\mathsf{M}_Q^2} &= \braket{3 \mathsf{M}_Q:\mathsf{M}_Q -
654     \mathrm{Tr}(\mathsf{M}_Q)^2 }\\
655     \braket{\mathsf{M}_Q}^2 &= 3 \braket{\mathsf{M}_Q}:\braket{\mathsf{M}_Q} -
656     \mathrm{Tr}(\braket{\mathsf{M}_Q})^2
657 gezelter 4401 \label{eq:flucQuad}
658 gezelter 4400 \end{align}
659 gezelter 4432 The bulk quadrupolarizability is given by,
660 gezelter 4399 \begin{equation}
661 gezelter 4432 \alpha_Q = \frac{\braket{\mathsf{M}_Q^2}-{\braket{\mathsf{M}_Q}}^2}{15 \epsilon_o V k_B T}.
662 gezelter 4399 \label{propConstQuad}
663     \end{equation}
664 gezelter 4432 Note that as in the dipolar case, $\alpha_Q$ and $\chi_Q$ are distinct
665     quantities. $\chi_Q$ includes contributions from local field
666     gradients due to the quadrupoles themselves, while $\alpha_Q$ is
667     relatively simple to compute from numerical simulations.
668 gezelter 4399
669     \subsection{Correction Factors}
670 gezelter 4400 In this section we generalize the treatment of Neumann \textit{et al.}
671     for quadrupolar fluids. Interactions involving multiple quadrupoles
672     are rank 4 tensors, and we therefore describe quantities in this
673     section using Einstein notation.
674    
675     In the presence of a uniform external field gradient,
676     $\partial_\alpha {E}^\circ_\beta $, the total field gradient at
677     $\mathbf{r}$ depends on the quadrupole polarization density at all
678     other points in the system,
679 mlamichh 4353 \begin{equation}
680 gezelter 4400 \partial_\alpha E_\beta(\mathbf{r}) = \partial_\alpha
681 mlamichh 4412 {E}^\circ_\beta(\mathbf{r}) + \frac{1}{8\pi \epsilon_o}\int
682 gezelter 4400 T_{\alpha\beta\gamma\delta}({\mathbf{r}-\mathbf{r}^\prime})
683     Q_{\gamma\delta}(\mathbf{r}^\prime) d\mathbf{r}^\prime
684 mlamichh 4353 \label{gradMaxwell}
685     \end{equation}
686 gezelter 4400 where and $T_{\alpha\beta\gamma\delta}$ is the quadrupole interaction
687     tensor connecting quadrupoles at $\mathbf{r}^\prime$ with the point of
688     interest ($\mathbf{r}$).
689 mlamichh 4353
690 gezelter 4408 % \begin{figure}
691     % \includegraphics[width=3in]{QuadrupoleFigure}
692     % \caption{With the real-space electrostatic methods, the molecular
693     % quadrupole tensor, $\mathbf{T}_{\alpha\beta\gamma\delta}(r)$,
694     % governing interactions between molecules is not the same for
695     % quadrupoles represented via sets of charges, point dipoles, or by
696     % single point quadrupoles.}
697     % \label{fig:quadrupolarFluid}
698     % \end{figure}
699 gezelter 4400
700     In simulations of quadrupolar fluids, the molecular quadrupoles may be
701     represented by closely-spaced point charges, by multiple point
702     dipoles, or by a single point quadrupole (see
703 gezelter 4408 Fig. \ref{fig:tensor}). In the case where point charges are
704 gezelter 4400 interacting via an electrostatic kernel, $v(r)$, the effective
705     molecular quadrupole tensor can obtained from four successive
706     applications of the gradient operator to the electrostatic kernel,
707     \begin{eqnarray}
708     T_{\alpha\beta\gamma\delta}(\mathbf{r}) &=& \nabla_\alpha \nabla_\beta
709     \nabla_\gamma \nabla_\delta v(r) \\
710     &=& \left(\delta_{\alpha\beta}\delta_{\gamma\delta} +
711     \delta_{\alpha\gamma}\delta_{\beta\delta}+
712     \delta_{\alpha\delta}\delta_{\beta\gamma}\right)\left(-\frac{v^\prime(r)}{r^3}
713     + \frac{v^{\prime \prime}(r)}{r^2}\right) \nonumber \\
714     & &+ \left(\delta_{\alpha\beta} r_\gamma r_\delta + 5 \mathrm{~permutations}
715     \right) \left(\frac{3v^\prime(r)}{r^5}-\frac{3v^{\prime \prime}(r)}{r^4} +
716     \frac{v^{\prime \prime \prime}(r)}{r^3}\right) \nonumber \\
717     & &+ r_\alpha r_\beta r_\gamma r_\delta
718     \left(-\frac{15v^\prime(r)}{r^7}+\frac{15v^{\prime \prime}(r)}{r^6}-\frac{6v^{\prime
719     \prime \prime}(r)}{r^5} + \frac{v^{\prime \prime \prime \prime}(r)}{r^4}\right),
720 mlamichh 4353 \label{quadCharge}
721 gezelter 4400 \end{eqnarray}
722     where $v(r)$ can either be the electrostatic kernel ($1/r$) or one of
723     the modified (Wolf or DSF) kernels.
724 mlamichh 4353
725 gezelter 4400 Similarly, when representing quadrupolar molecules with multiple point
726 gezelter 4413 \textit{dipoles}, the molecular quadrupole interaction tensor can be
727     obtained using two successive applications of the gradient operator to
728     the dipole interaction tensor,
729 gezelter 4400 \begin{eqnarray}
730     T_{\alpha\beta\gamma\delta}(\mathbf{r}) &=& \nabla_\alpha \nabla_\beta
731 gezelter 4413 T_{\gamma\delta}(\mathbf{r}) \\
732 gezelter 4400 & = & \delta_{\alpha\beta}\delta_{\gamma\delta} \frac{v^\prime_{21}(r)}{r} +
733     \left(\delta_{\alpha\gamma}\delta_{\beta\delta}+
734     \delta_{\alpha\delta}\delta_{\beta\gamma}\right)\frac{v_{22}(r)}{r^2}
735     \nonumber\\
736     & &+ \delta_{\gamma\delta} r_\alpha r_\beta
737     \left(\frac{v^{\prime \prime}_{21}(r)}{r^2}-\frac{v^\prime_{21}(r)}{r^3} \right)\nonumber \\
738     & &+\left(\delta_{\alpha\beta} r_\gamma r_\delta +
739     \delta_{\alpha\gamma} r_\beta r_\delta +\delta_{\alpha\delta}
740     r_\gamma r_\beta + \delta_{\beta\gamma} r_\alpha r_\delta
741     +\delta_{\beta\delta} r_\alpha r_\gamma \right)
742     \left(\frac{v^\prime_{22}(r)}{r^3}-\frac{2v_{22}(r)}{r^4}\right)
743     \nonumber \\
744     & &+ r_\alpha r_\beta r_\gamma r_\delta
745     \left(\frac{v^{\prime
746     \prime}_{22}(r)}{r^4}-\frac{5v^\prime_{22}(r)}{r^5}+\frac{8v_{22}(r)}{r^6}\right),
747 mlamichh 4353 \label{quadDip}
748 gezelter 4400 \end{eqnarray}
749 gezelter 4413 where $T_{\gamma\delta}(\mathbf{r})$ is a dipole-dipole interaction
750 gezelter 4432 tensor that depends on the level of the approximation (see
751     Eq. (\ref{dipole-diopleTensor})).\cite{PaperI,PaperII} Similarly
752     $v_{21}(r)$ and $v_{22}(r)$ are the radial function for different real
753     space cutoff methods defined in the first paper in this
754     series.\cite{PaperI}
755 gezelter 4400
756     For quadrupolar liquids modeled using point quadrupoles, the
757     interaction tensor can be constructed as,
758     \begin{eqnarray}
759 mlamichh 4412 T_{\alpha\beta\gamma\delta}(\mathbf{r}) &=&
760 gezelter 4400 \left(\delta_{\alpha\beta}\delta_{\gamma\delta}
761     +
762     \delta_{\alpha\gamma}\delta_{\beta\delta}+
763     \delta_{\alpha\delta}\delta_{\beta\gamma}\right)v_{41}(r)
764 mlamichh 4412 + (\delta_{\gamma\delta} r_\alpha r_\beta + \mathrm{ 5\; permutations}) \frac{v_{42}(r)}{r^2} \nonumber \\
765 gezelter 4400 & & + r_\alpha r_\beta r_\gamma r_\delta \left(\frac{v_{43}(r)}{r^4}\right),
766 mlamichh 4353 \label{quadRadial}
767 gezelter 4400 \end{eqnarray}
768     where again $v_{41}(r)$, $v_{42}(r)$, and $v_{43}(r)$ are radial
769     functions defined in Paper I of the series. \cite{PaperI} Note that
770     these radial functions have different functional forms depending on
771     the level of approximation being employed.
772    
773 gezelter 4432 The integral in Eq. (\ref{gradMaxwell}) can be divided into two
774 gezelter 4400 parts, $|\mathbf{r}-\mathbf{r}^\prime|\rightarrow 0 $ and
775     $|\mathbf{r}-\mathbf{r}^\prime|> 0$. Since the self-contribution to
776 gezelter 4432 the field gradient vanishes at the singularity (see the supplemental
777     material),\cite{supplemental} Eq. (\ref{gradMaxwell}) can be
778     written as,
779 mlamichh 4353 \begin{equation}
780 gezelter 4400 \partial_\alpha E_\beta(\mathbf{r}) = \partial_\alpha {E}^\circ_\beta(\mathbf{r}) +
781 mlamichh 4412 \frac{1}{8\pi \epsilon_o}\int\limits_{|\mathbf{r}-\mathbf{r}^\prime|> 0 }
782 gezelter 4400 T_{\alpha\beta\gamma\delta}(\mathbf{r}-\mathbf{r}^\prime)
783     {Q}_{\gamma\delta}(\mathbf{r}^\prime) d\mathbf{r}^\prime.
784 mlamichh 4353 \end{equation}
785 gezelter 4400 If $\mathbf{r} = \mathbf{r}^\prime$ is excluded from the integration,
786     the total gradient can be most easily expressed in terms of
787     traceless quadrupole density as below,\cite{LoganI81}
788 mlamichh 4353 \begin{equation}
789 gezelter 4400 \partial_\alpha E_\beta(\mathbf{r}) = \partial_\alpha
790 mlamichh 4412 {E}^\circ_\beta(\mathbf{r}) + \frac{1}{24\pi
791 gezelter 4400 \epsilon_o}\int\limits_{|\mathbf{r}-\mathbf{r}^\prime|> 0 }
792     T_{\alpha\beta\gamma\delta}(\mathbf{r}-\mathbf{r}^\prime) \Theta_{\gamma\delta}(\mathbf{r}') d\mathbf{r}^\prime,
793 mlamichh 4353 \end{equation}
794 gezelter 4400 where
795     $\Theta_{\alpha\beta} = 3Q_{\alpha\beta} - \delta_{\alpha\beta}Tr(Q)$
796     is the traceless quadrupole density. In analogy to
797 gezelter 4413 Eq. (\ref{pertQuad}) above, the quadrupole polarization density may
798     now be related to the quadrupolar susceptibility, $\chi_Q$,
799 mlamichh 4353 \begin{equation}
800 gezelter 4400 \frac{1}{3}{\Theta}_{\alpha\beta}(\mathbf{r}) = \epsilon_o {\chi}_Q
801 mlamichh 4412 \left[\partial_\alpha {E}^\circ_\beta(\mathbf{r}) + \frac{1}{24\pi
802 gezelter 4400 \epsilon_o}\int\limits_{|\mathbf{r}-\mathbf{r}^\prime|> 0 }
803     T_{\alpha\beta\gamma\delta}(\mathbf{r}-\mathbf{r}^\prime)
804 mlamichh 4412 \Theta_{\gamma\delta}(\mathbf{r}^\prime) d\mathbf{r}^\prime \right].
805 mlamichh 4353 \end{equation}
806 mlamichh 4412 For periodic boundaries and with a uniform imposed
807 gezelter 4400 $\partial_\alpha E^\circ_\beta$, the quadrupole density
808     ${\Theta}_{\alpha\beta}$ will be uniform over the entire space. After
809     performing a Fourier transform (see the Appendix in
810     ref. \onlinecite{NeumannI83}) we obtain,
811 mlamichh 4353 \begin{equation}
812 gezelter 4400 \frac{1}{3}\tilde{\Theta}_{\alpha\beta}(\mathbf{k})=
813     \epsilon_o {\chi}_Q \left[{\partial_\alpha
814 mlamichh 4412 \tilde{E}^\circ_\beta}(\mathbf{k})+ \frac{1}{24\pi
815 gezelter 4400 \epsilon_o} \tilde{T}_{\alpha\beta\gamma\delta}(\mathbf{k})
816     \tilde{\Theta}_{\gamma\delta}(\mathbf{k})\right].
817 mlamichh 4412 \label{fourierQuad}
818 mlamichh 4353 \end{equation}
819 gezelter 4413 If the applied field gradient is homogeneous over the entire volume,
820     ${\partial_ \alpha \tilde{E}^\circ_\beta}(\mathbf{k}) = 0 $ except at
821     $ \mathbf{k} = 0$. Similarly, the quadrupolar polarization density can
822     also considered uniform over entire space. As in the dipolar case,
823     \cite{NeumannI83} the only relevant contribution from the interaction
824 gezelter 4432 tensor will also be when $\mathbf{k} = 0$. Therefore Eq.
825 gezelter 4413 (\ref{fourierQuad}) can be written as,
826 mlamichh 4412 \begin{equation}
827     \frac{1}{3}\tilde{\Theta}_{\alpha\beta}(\mathrm{0})=
828     \epsilon_o {\chi}_Q \left[{\partial_\alpha
829     \tilde{E}^\circ_\beta}(\mathrm{0})+ \frac{1}{24\pi
830     \epsilon_o} \tilde{T}_{\alpha\beta\gamma\delta}(\mathrm{0})
831     \tilde{\Theta}_{\gamma\delta}(\mathrm{0})\right].
832     \label{fourierQuadZeroK}
833     \end{equation}
834 gezelter 4413 The quadrupolar tensor
835 gezelter 4415 $\tilde{T}_{\alpha\beta\gamma\delta}(\mathrm{0})$ is a rank 4 tensor
836 gezelter 4413 with 81 elements. The only non-zero elements, however, are those with
837     two doubly-repeated indices, \textit{i.e.}
838     $\tilde{T}_{aabb}(\mathrm{0})$ and all permutations of these indices.
839     The special case of quadruply-repeated indices,
840     $\tilde{T}_{aaaa}(\mathrm{0})$ also survives (see appendix
841     \ref{ap:quadContraction}). Furthermore, for the both diagonal and
842     non-diagonal components of the quadrupolar polarization
843     $\tilde{\Theta}_{\alpha\beta}$, we can contract the second term in
844 gezelter 4432 Eq. \ref{fourierQuadZeroK} (see appendix
845 gezelter 4413 \ref{ap:quadContraction}):
846 mlamichh 4412 \begin{equation}
847 gezelter 4413 \tilde{T}_{\alpha\beta\gamma\delta}(\mathrm{0})\tilde{\Theta}_{\gamma\delta}(\mathrm{0})=
848     8 \pi \mathrm{B} \tilde{\Theta}_{\alpha\beta}(\mathrm{0}).
849 mlamichh 4412 \label{quadContraction}
850     \end{equation}
851 gezelter 4413 Here $\mathrm{B} = \tilde{T}_{abab}(\mathrm{0}) / 4 \pi$ for
852 gezelter 4432 $a \neq b$. Using this quadrupolar contraction we can solve Eq.
853 gezelter 4415 \ref{fourierQuadZeroK} as follows
854 gezelter 4400 \begin{eqnarray}
855 mlamichh 4412 \frac{1}{3}\tilde{\Theta}_{\alpha\beta}(\mathrm{0}) &=& \epsilon_o
856 gezelter 4400 {\chi}_Q
857     \left[{\partial_\alpha
858 mlamichh 4412 \tilde{E}^\circ_\beta}(\mathrm{0})+
859 gezelter 4413 \frac{\mathrm{B}}{3
860 mlamichh 4412 \epsilon_o}
861     {\tilde{\Theta}}_{\alpha\beta}(\mathrm{0})\right]
862     \nonumber \\
863     &=& \left[\frac{\epsilon_o {\chi}_Q} {1-{\chi}_Q \mathrm{B}}\right]
864 gezelter 4413 {\partial_\alpha \tilde{E}^\circ_\beta}(\mathrm{0}).
865 gezelter 4415 \label{fourierQuad2}
866 mlamichh 4412 \end{eqnarray}
867 gezelter 4419 In real space, the correction factor is found to be,
868 gezelter 4413 \begin{equation}
869     \mathrm{B} = \frac{1}{4 \pi} \tilde{T}_{abab}(0) = \frac{1}{4 \pi} \int_V {T}_{abab}(\mathbf{r}) d\mathbf{r},
870     \end{equation}
871 mlamichh 4412 %If the applied field gradient is homogeneous over the
872     %entire volume, ${\partial_ \alpha \tilde{E}^\circ_\beta}(\mathbf{k}) = 0 $ except at
873     %$ \mathbf{k} = 0$. As in the dipolar case, the only relevant
874     %contribution of the interaction tensor will also be when $\mathbf{k} = 0$.
875     %Therefore equation (\ref{fourierQuad}) can be written as,
876     %\begin{equation}
877     %\frac{1}{3}{\tilde{\Theta}}_{\alpha\beta}({0}) = \epsilon_o {\chi}_Q
878     %\left(1-\frac{1}{4\pi} {\chi}_Q \tilde{T}_{\alpha\beta\alpha\beta}({0})\right)^{-1} \partial_\alpha \tilde{E}^\circ_\beta({0})
879     %\label{fourierQuad2}
880     %\end{equation}
881     %where $\tilde{T}_{\alpha\beta\alpha\beta}({0})$ can be evaluated as,
882     %\begin{equation}
883     %\tilde{T}_{\alpha\beta\alpha\beta}({0}) = \int_V {T}_{\alpha\beta\alpha\beta}(\mathbf{r})d\mathbf{r}
884     %\label{realTensorQaud}
885     %\end{equation}
886     %We may now define a quantity to represent the contribution from the
887     %$\mathbf{k} = 0$ portion of the interaction tensor,
888     %\begin{equation}
889     %B = \frac{1}{4 \pi} \int_V T_{\alpha \beta \alpha \beta}(\mathbf{r})
890     %d\mathbf{r}
891     %\end{equation}
892 gezelter 4400 which has been integrated over the interaction volume $V$ and has
893     units of $\mathrm{length}^{-2}$.
894 mlamichh 4353
895 gezelter 4432 In terms of the traced quadrupole moment, Eq. (\ref{fourierQuad2})
896 gezelter 4413 can be written,
897 mlamichh 4353 \begin{equation}
898 gezelter 4400 \mathsf{Q} - \frac{\mathbf{I}}{3} \mathrm{Tr}(\mathsf{Q})
899 mlamichh 4412 = \frac{\epsilon_o {\chi}_Q}{1- {\chi}_Q \mathrm{B}} \nabla \mathbf{E}^\circ
900 mlamichh 4353 \label{tracedConstQuad}
901     \end{equation}
902 gezelter 4400 Comparing (\ref{tracedConstQuad}) and (\ref{flucQuad}) we obtain,
903 mlamichh 4353 \begin{equation}
904 gezelter 4400 \frac{\braket{\mathsf{M}_Q^2}-{\braket{\mathsf{M}_Q}}^2}{15 \epsilon_o
905 mlamichh 4412 V k_B T} = \frac{\chi_Q} {1 - \chi_Q \mathrm{B}},
906 mlamichh 4353 \end{equation}
907 gezelter 4400 or equivalently,
908 mlamichh 4353 \begin{equation}
909 gezelter 4400 \chi_Q = \frac{\braket{\mathsf{M}_Q^2}-{\braket{\mathsf{M}_Q}}^2}{15 \epsilon_o
910 mlamichh 4412 V k_B T} \left(1 + \mathrm{B} \frac{\braket{\mathsf{M}_Q^2}-{\braket{\mathsf{M}_Q}}^2}{15 \epsilon_o
911 gezelter 4432 V k_B T} \right)^{-1}.
912 gezelter 4400 \label{eq:finalForm}
913 mlamichh 4353 \end{equation}
914 gezelter 4400 Eq. (\ref{eq:finalForm}) now expresses a bulk property (the
915 gezelter 4413 quadrupolar susceptibility, $\chi_Q$) in terms of a fluctuation in the
916     system quadrupole moment and a quadrupolar correction factor
917     ($\mathrm{B}$). The correction factors depend on the cutoff method
918     being employed in the simulation, and these are listed in Table
919     \ref{tab:B}.
920 mlamichh 4353
921 gezelter 4400 In terms of the macroscopic quadrupole polarizability, $\alpha_Q$,
922     which may be thought of as the ``conducting boundary'' version of the
923     susceptibility,
924 mlamichh 4353 \begin{equation}
925 mlamichh 4412 \chi_Q = \frac{\alpha_Q}{1 + \mathrm{B} \alpha_Q}.
926 gezelter 4408 \label{eq:quadrupolarSusceptiblity}
927 mlamichh 4353 \end{equation}
928 mlamichh 4412 If an electrostatic method produces $\mathrm{B} \rightarrow 0$, the computed
929 gezelter 4400 quadrupole polarizability and quadrupole susceptibility converge to
930     the same value.
931    
932     \begin{sidewaystable}
933 gezelter 4415 \caption{Expressions for the quadrupolar correction factor
934     ($\mathrm{B}$) for the real-space electrostatic methods in terms
935     of the damping parameter ($\alpha$) and the cutoff radius
936     ($r_c$). The units of the correction factor are
937     $ \mathrm{length}^{-2}$ for quadrupolar fluids.}
938 mlamichh 4353 \label{tab:B}
939 gezelter 4416 \begin{tabular}{l|c|c|c} \toprule
940     \multirow{2}{*}{Method} & \multicolumn{3}{c}{Molecular Representation} \\
941     & charges & dipoles & quadrupoles \\\colrule
942 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}}$ \\
943 mlamichh 4412 Gradient-shifted (GSF) & $- \frac{8 \alpha^5 {r_c}^3e^{-\alpha^2 r_c^2}}{15\sqrt{\pi}} $ & 0 & $-\frac{4{\alpha}^7{r_c}^5 e^{-\alpha^2 r_c^2}(-1+2\alpha ^2 r_c^2)}{45\sqrt{\pi}}$\\
944     Taylor-shifted (TSF) & $ -\frac{8 \alpha^5 {r_c}^3e^{-\alpha^2 r_c^2}}{15\sqrt{\pi}} $ & $\frac{4\;\mathrm{erfc(\alpha r_c)}}{{5r_c}^2} + \frac{8\alpha e^{-\alpha^2{r_c}^2}\left(3+ 2\alpha^2 {r_c}^2 +\alpha^4{r_c}^4 \right)}{15\sqrt{\pi}r_c}$ & $\frac{2\;\mathrm{erfc}(\alpha r_c )}{{r_c}^2} + \frac{4{\alpha}e^{-\alpha^2 r_c^2}\left(45 + 30\alpha ^2 {r_c}^2 + 12\alpha^4 {r_c}^4 + 3\alpha^6 {r_c}^6 + 2 \alpha^8 {r_c}^8\right)}{45\sqrt{\pi}{r_c}}$ \\
945 gezelter 4408 \colrule
946 gezelter 4432 %Spherical Cutoff (SC) & \multicolumn{3}{c}{$ -\frac{8 \alpha^5
947     % {r_c}^3e^{-\alpha^2 r_c^2}}{15\sqrt{\pi}} $}\\
948 gezelter 4416 Ewald-Kornfeld (EK) & \multicolumn{3}{c}{$ -\frac{8 \kappa^5 {r_c}^3e^{-\kappa^2 r_c^2}}{15\sqrt{\pi}}$} \\
949 gezelter 4408 \botrule
950 gezelter 4400 \end{tabular}
951     \end{sidewaystable}
952 gezelter 4399
953     \section{Screening of Charges by Multipolar Fluids}
954 gezelter 4401 \label{sec:PMF}
955     In a dipolar fluid, the static dielectric constant is also a measure
956     of the ability of the fluid to screen charges from one another. A set
957     of point charges creates an inhomogeneous field in the fluid, and the
958     fluid responds to this field as if it was created externally or via
959     local polarization fluctuations. For this reason, the dielectric
960     constant can be used to estimate an effective potential between two
961     point charges ($C_i$ and $C_j$) embedded in the fluid,
962 gezelter 4399 \begin{equation}
963 gezelter 4401 U_\mathrm{effective} = \frac{C_i C_j}{4 \pi \epsilon_0 \epsilon
964     r_{ij}}.
965     \label{eq:effectivePot}
966 gezelter 4399 \end{equation}
967 gezelter 4401
968     The same set of point charges can also create an inhomogeneous field
969     \textit{gradient}, and this will cause a response in a quadrupolar
970     fluid that will also cause an effective screening. As discussed above,
971 gezelter 4431 however, the relevant physical property in quadrupolar fluids is the
972 gezelter 4401 susceptibility, $\chi_Q$. The screening dielectric associated with
973     the quadrupolar susceptibility is defined as,\cite{Ernst92}
974 gezelter 4399 \begin{equation}
975 gezelter 4401 \epsilon = 1 + \chi_Q G = 1 + G \frac{\alpha_Q}{1 + \alpha_Q B}
976     \label{eq:dielectricFromQuadrupoles}
977 gezelter 4399 \end{equation}
978 gezelter 4401 where $G$ is a geometrical factor that depends on the geometry of the
979     field perturbation,
980     \begin{equation}
981     G = \frac{\int_V \left| \nabla \mathbf{E}^\circ \right|^2 d\mathbf{r}}
982     {\int_V \left|\mathbf{E}^\circ\right|^2 d\mathbf{r}}
983     \end{equation}
984     integrated over the interaction volume. Note that this geometrical
985     factor is also required to compute effective dielectric constants even
986     when the field gradient is homogeneous over the entire sample.
987 gezelter 4399
988 gezelter 4401 To measure effective screening in a multipolar fluid, we compute an
989     effective interaction potential, the potential of mean force (PMF),
990 gezelter 4432 between positively and negatively charged ions when they are screened
991     by the intervening fluid. The PMF is obtained from a sequence of
992 gezelter 4401 simulations in which two ions are constrained to a fixed distance, and
993     the average constraint force to hold them at a fixed distance $r$ is
994     collected during a long simulation,\cite{Wilfred07}
995 gezelter 4400 \begin{equation}
996 gezelter 4401 w(r) = \int_{r_o}^{r}\left\langle \frac{\partial f}{\partial r^\prime}
997     \right\rangle_{r^\prime} dr^\prime + 2k_BT \ln\left(\frac{r}{r_o}\right) + w(r_o),
998     \label{eq:pmf}
999 gezelter 4400 \end{equation}
1000 gezelter 4401 where $\braket{\partial f/\partial r^\prime}_{r^\prime}$ is the mean
1001     constraint force required to hold the ions at distance $r^\prime$,
1002     $2k_BT \log(r/r_o)$ is the Fixman factor,\cite{Fixman:1974fk} and
1003     $r_o$ is a reference position (usually taken as a large separation
1004 gezelter 4431 between the ions). If the dielectric constant is a good measure of the
1005 gezelter 4401 screening at all inter-ion separations, we would expect $w(r)$ to have
1006     the form in Eq. (\ref{eq:effectivePot}). Because real fluids are not
1007     continuum dielectrics, the effective dielectric constant is a function
1008     of the interionic separation,
1009 gezelter 4400 \begin{equation}
1010 gezelter 4425 \epsilon(r) = \frac{u_\mathrm{raw}(r) - u_\mathrm{raw}(r_o) }{w(r) - w(r_o)}
1011 gezelter 4400 \end{equation}
1012 gezelter 4401 where $u_\mathrm{raw}(r)$ is the direct charge-charge interaction
1013     potential that is in use during the simulation. $\epsilon(r)$ may
1014     vary considerably from the bulk estimates at short distances, although
1015     it should converge to the bulk value as the separation between the
1016     ions increases.
1017 gezelter 4400
1018 gezelter 4399 \section{Simulation Methodology}
1019 mlamichh 4422
1020 gezelter 4401 To test the formalism developed in the preceding sections, we have
1021     carried out computer simulations using three different techniques: i)
1022     simulations in the presence of external fields, ii) equilibrium
1023     calculations of box moment fluctuations, and iii) potentials of mean
1024     force (PMF) between embedded ions. In all cases, the fluids were
1025     composed of point multipoles protected by a Lennard-Jones potential.
1026     The parameters used in the test systems are given in table
1027     \ref{Tab:C}.
1028 mlamichh 4422
1029     \begin{sidewaystable}
1030 gezelter 4401 \caption{\label{Tab:C}The parameters used in simulations to evaluate
1031     the dielectric response of the new real-space methods.}
1032 mlamichh 4389 \begin{tabularx}{\textwidth}{r|cc|YYccc|Yccc} \hline
1033     & \multicolumn{2}{c|}{LJ parameters} &
1034     \multicolumn{5}{c|}{Electrostatic moments} & & & & \\
1035     Test system & $\sigma$& $\epsilon$ & $C$ & $D$ &
1036     $Q_{xx}$ & $Q_{yy}$ & $Q_{zz}$ & mass & $I_{xx}$ & $I_{yy}$ &
1037     $I_{zz}$ \\ \cline{6-8}\cline{10-12}
1038     & (\AA) & (kcal/mol) & (e) & (debye) & \multicolumn{3}{c|}{(debye \AA)} & (amu) & \multicolumn{3}{c}{(amu
1039     \AA\textsuperscript{2})} \\ \hline
1040 gezelter 4401 Dipolar fluid & 3.41 & 0.2381 & - & 1.4026 &-&-&-& 39.948 & 11.613 & 11.613 & 0.0 \\
1041     Quadrupolar fluid & 2.985 & 0.265 & - & - & 0.0 & 0.0 &-2.139 & 18.0153 & 43.0565 & 43.0565 & 0.0 \\
1042 mlamichh 4389 \ce{q+} & 1.0 & 0.1 & +1 & - & - & - & - & 22.98 & - & - & - \\
1043     \ce{q-} & 1.0 & 0.1 & -1 & - & - & - & - & 22.98 & - & - & - \\ \hline
1044     \end{tabularx}
1045 mlamichh 4422 \end{sidewaystable}
1046 mlamichh 4363
1047 gezelter 4401 The first of the test systems consists entirely of fluids of point
1048     dipolar or quadrupolar molecules in the presence of constant field or
1049     field gradients. Since there are no isolated charges within the
1050     system, the divergence of the field will be zero, \textit{i.e.}
1051 gezelter 4432 $\nabla \cdot \mathbf{E} = 0$. This condition can be satisfied
1052 gezelter 4401 by using the relatively simple applied potential as described in the
1053 gezelter 4432 supplemental material.\cite{supplemental}
1054 mlamichh 4353
1055 gezelter 4401 When a constant electric field or field gradient is applied to the
1056     system, the molecules align along the direction of the applied field,
1057     and polarize to a degree determined both by the strength of the field
1058 gezelter 4431 and the fluid's polarizability. We have calculated ensemble averages
1059 gezelter 4401 of the box dipole and quadrupole moments as a function of the strength
1060     of the applied fields. If the fields are sufficiently weak, the
1061     response is linear in the field strength, and one can easily compute
1062     the polarizability directly from the simulations.
1063 mlamichh 4389
1064 gezelter 4401 The second set of test systems consists of equilibrium simulations of
1065     fluids of point dipolar or quadrupolar molecules simulated in the
1066     absence of any external perturbation. The fluctuation of the ensemble
1067     averages of the box multipolar moment was calculated for each of the
1068     multipolar fluids. The box multipolar moments were computed as simple
1069     sums over the instantaneous molecular moments, and fluctuations in
1070     these quantities were obtained from Eqs. (\ref{eq:flucDip}) and
1071     (\ref{eq:flucQuad}). The macroscopic polarizabilities of the system at
1072     a were derived using Eqs.(\ref{flucDipole}) and (\ref{flucQuad}).
1073 mlamichh 4389
1074 gezelter 4401 The final system consists of dipolar or quadrupolar fluids with two
1075     oppositely charged ions embedded within the fluid. These ions are
1076     constrained to be at fixed distance throughout a simulation, although
1077     they are allowed to move freely throughout the fluid while satisfying
1078     that constraint. Separate simulations were run at a range of
1079     constraint distances. A dielectric screening factor was computed using
1080     the ratio between the potential between the two ions in the absence of
1081     the fluid medium and the PMF obtained from the simulations.
1082 mlamichh 4389
1083 gezelter 4401 We carried out these simulations for all three of the new real-space
1084     electrostatic methods (SP, GSF, and TSF) that were developed in the
1085     first paper (Ref. \onlinecite{PaperI}) in the series. The radius of
1086     the cutoff sphere was taken to be 12~\AA. Each of the real space
1087     methods also depends on an adjustable damping parameter $\alpha$ (in
1088     units of $\mathrm{length}^{-1}$). We have selected ten different
1089     values of damping parameter: 0.0, 0.05, 0.1, 0.15, 0.175, 0.2, 0.225,
1090     0.25, 0.3, and 0.35~\AA$^{-1}$ in our simulations of the dipolar
1091     liquids, while four values were chosen for the quadrupolar fluids:
1092     0.0, 0.1, 0.2, and 0.3~\AA$^{-1}$.
1093 mlamichh 4389
1094 gezelter 4401 For each of the methods and systems listed above, a reference
1095     simulation was carried out using a multipolar implementation of the
1096     Ewald sum.\cite{Smith82,Smith98} A default tolerance of
1097     $1 \times 10^{-8}$~kcal/mol was used in all Ewald calculations,
1098     resulting in Ewald coefficient 0.3119~\AA$^{-1}$ for a cutoff radius
1099     of 12~\AA. All of the electrostatics and constraint methods were
1100     implemented in our group's open source molecular simulation program,
1101     OpenMD,\cite{Meineke05,openmd} which was used for all calculations in
1102     this work.
1103    
1104     Dipolar systems contained 2048 Lennard-Jones-protected point dipolar
1105     (Stockmayer) molecules with reduced density $\rho^* = 0.822$,
1106     temperature $T^* = 1.15$, moment of inertia $I^* = 0.025$, and dipole
1107     moment $\mu^* = \sqrt{3.0}$. These systems were equilibrated for
1108     0.5~ns in the canonical (NVT) ensemble. Data collection was carried
1109     out over a 1~ns simulation in the microcanonical (NVE) ensemble. Box
1110     dipole moments were sampled every fs. For simulations with external
1111 gezelter 4432 perturbations, field strengths ranging from
1112     $0 - 10 \times 10^{-4}$~V/\AA\ with increments of $ 10^{-4}$~V/\AA\
1113     were carried out for each system. For dipolar systems the interaction
1114     potential between molecules $i$ and $j$,
1115     \begin{equation}
1116     u_{ij}(\mathbf{r}_{ij}, \mathbf{D}_i, \mathbf{D}_j) = 4 \epsilon \left(
1117     \left(\frac{\sigma}{r_{ij}} \right)^{12} - \left(\frac{\sigma}{r_{ij}}
1118     \right)^{6} \right) - \mathbf{D}_i \cdot
1119     \mathbf{T}(\mathbf{r}_{ij}) \cdot \mathbf{D}_j
1120     \end{equation}
1121     where the dipole interaction tensor, $\mathbf{T}(\mathbf{r})$, is
1122     given in Eq. (\ref{dipole-diopleTensor}).
1123 gezelter 4401
1124     Quadrupolar systems contained 4000 linear point quadrupoles with a
1125 gezelter 4408 density $2.338 \mathrm{~g/cm}^3$ at a temperature of 500~K. These
1126 gezelter 4401 systems were equilibrated for 200~ps in a canonical (NVT) ensemble.
1127     Data collection was carried out over a 500~ps simulation in the
1128 gezelter 4408 microcanonical (NVE) ensemble. Components of box quadrupole moments
1129     were sampled every 100 fs. For quadrupolar simulations with external
1130 gezelter 4401 field gradients, field strengths ranging from
1131 gezelter 4408 $0 - 9 \times 10^{-2}$~V/\AA$^2$ with increments of
1132 gezelter 4432 $10^{-2}$~V/\AA$^2$ were carried out for each system. For quadrupolar
1133     systems the interaction potential between molecules $i$ and $j$,
1134     \begin{equation}
1135     u_{ij}(\mathbf{r}_{ij}, \mathsf{Q}_i, \mathsf{Q}_j) = 4 \epsilon \left(
1136     \left(\frac{\sigma}{r_{ij}} \right)^{12} - \left(\frac{\sigma}{r_{ij}}
1137     \right)^{6} \right) + \mathsf{Q}_i \colon
1138     \mathbf{T}(\mathbf{r}_{ij}) \colon \mathsf{Q}_j
1139     \end{equation}
1140     where the quadrupole interaction tensor is given in
1141     Eq. (\ref{quadRadial}).
1142 gezelter 4401
1143     To carry out the PMF simulations, two of the multipolar molecules in
1144     the test system were converted into \ce{q+} and \ce{q-} ions and
1145     constrained to remain at a fixed distance for the duration of the
1146 gezelter 4408 simulation. The constrained distance was then varied from 5--12~\AA.
1147     In the PMF calculations, all simulations were equilibrated for 500 ps
1148     in the NVT ensemble and run for 5 ns in the microcanonical (NVE)
1149     ensemble. Constraint forces were sampled every 20~fs.
1150 mlamichh 4389
1151     \section{Results}
1152 mlamichh 4397 \subsection{Dipolar fluid}
1153     \begin{figure}
1154 gezelter 4428 \includegraphics[width=5in]{dielectricFinal_Dipole.eps}
1155 gezelter 4401 \caption{The polarizability ($\alpha_D$), correction factor ($A$), and
1156     dielectric constant ($\epsilon$) for the test dipolar fluid. The
1157     left panels were computed using external fields, and those on the
1158     right are the result of equilibrium fluctuations. In the GSF and SP
1159 gezelter 4432 methods, the corrections are large for small values of $\alpha$, and
1160     an optimal damping coefficient is evident around 0.25 \AA$^{-1}$.
1161 gezelter 4401 The dashed lines in the upper panel indicate back-calculation of the
1162 gezelter 4404 polarizability using the Ewald estimate (Refs. \onlinecite{Adams81}
1163     and \onlinecite{NeumannI83}) for the dielectric constant.}
1164 mlamichh 4397 \label{fig:dielectricDipole}
1165     \end{figure}
1166 gezelter 4432 The bulk polarizability ($\alpha_D$) for the dipolar fluid is
1167 gezelter 4408 shown in the upper panels in Fig. \ref{fig:dielectricDipole}. The
1168 gezelter 4401 polarizability obtained from the both perturbation and fluctuation
1169     approaches are in excellent agreement with each other. The data also
1170 gezelter 4431 show a strong dependence on the damping parameter for both the Shifted
1171 gezelter 4401 Potential (SP) and Gradient Shifted force (GSF) methods, while Taylor
1172     shifted force (TSF) is largely independent of the damping
1173     parameter.
1174 mlamichh 4397
1175 gezelter 4401 The calculated correction factors discussed in section
1176 gezelter 4408 \ref{sec:corrFactor} are shown in the middle panels. Because the TSF
1177 gezelter 4401 method has $A = 1$ for all values of the damping parameter, the
1178     computed polarizabilities need no correction for the dielectric
1179     calculation. The value of $A$ varies with the damping parameter in
1180     both the SP and GSF methods, and inclusion of the correction yields
1181     dielectric estimates (shown in the lower panel) that are generally too
1182 gezelter 4408 large until the damping reaches $\sim$~0.25~\AA$^{-1}$. Above this
1183 gezelter 4433 value, the dielectric constants are in reasonable agreement with
1184 gezelter 4408 previous simulation results.\cite{NeumannI83}
1185 mlamichh 4397
1186 gezelter 4401 Figure \ref{fig:dielectricDipole} also contains back-calculations of
1187     the polarizability using the reference (Ewald) simulation
1188     results.\cite{NeumannI83} These are indicated with dashed lines in the
1189     upper panels. It is clear that the expected polarizability for the SP
1190     and GSF methods are quite close to results obtained from the
1191     simulations. This indicates that the correction formula for the
1192 gezelter 4433 dipolar fluid (Eq. \ref{correctionFormula}) is extraordinarily
1193     sensitive when the value of $\mathrm{A}$ deviates significantly from
1194     unity. It is also apparent that Gaussian damping is essential for
1195     capturing the field effects from other
1196     dipoles. Eq. (\ref{correctionFormula}) works well when real-space
1197     methods employ moderate damping, but is not capable of providing
1198     adequate correction for undamped or weakly-damped multipoles.
1199 gezelter 4401
1200 gezelter 4432 Because the dielectric correction in Eq. (\ref{correctionFormula}) is
1201     so sensitive to $A$ values away from unity, the entries in table
1202     \ref{tab:A} can provide an effective minimum on the values of $\alpha$
1203     that should be used. With a minimum $A=0.995$ and a cutoff radius of
1204     12~\AA, the minimum $\alpha$ values are 0.241~\AA$^{-1}$ (SP) or
1205 gezelter 4433 0.268~\AA$^{-1}$ (GSF). The TSF method is not sensitive to the choice
1206 gezelter 4432 of damping parameter.
1207 gezelter 4401
1208 mlamichh 4397 \begin{figure}
1209 gezelter 4428 \includegraphics[width=4in]{screen_pmf.eps}
1210 gezelter 4401 \caption{The distance-dependent screening factor, $\epsilon(r)$,
1211     between two ions immersed in the dipolar fluid. The new methods are
1212     shown in separate panels, and different values of the damping
1213     parameter ($\alpha$) are indicated with different symbols. All of
1214     the methods appear to be converging to the bulk dielectric constant
1215 gezelter 4432 ($\sim 65$) for higher values of $\alpha$ and at large ion
1216     separations.}
1217 mlamichh 4397 \label{fig:ScreeningFactor_Dipole}
1218     \end{figure}
1219 gezelter 4401 We have also evaluated the distance-dependent screening factor,
1220     $\epsilon(r)$, between two oppositely charged ions when they are
1221     placed in the dipolar fluid. These results were computed using
1222     Eq. \ref{eq:pmf} and are shown in Fig.
1223     \ref{fig:ScreeningFactor_Dipole}.
1224 mlamichh 4397
1225 gezelter 4401 The screening factor is similar to the dielectric constant, but
1226     measures a local property of the ions in the fluid and depends on both
1227     ion-dipole and dipole-dipole interactions. These interactions depend
1228     on the distance between ions as well as the electrostatic interaction
1229     methods utilized in the simulations. The screening should converge to
1230     the dielectric constant when the field due to ions is small. This
1231     occurs when the ions are separated (or when the damping parameter is
1232     large). In Fig. \ref{fig:ScreeningFactor_Dipole} we observe that for
1233     the higher value of damping alpha \textit{i.e.}
1234 gezelter 4408 $\alpha = 0.2$~\AA$^{-1}$ and $0.3$~\AA$^{-1}$ and large separation
1235 gezelter 4401 between ions, the screening factor does indeed approach the correct
1236     dielectric constant.
1237 mlamichh 4397
1238 gezelter 4401 It is also notable that the TSF method again displays smaller
1239     perturbations away from the correct dielectric screening behavior. We
1240     also observe that for TSF method yields high dielectric screening even
1241 gezelter 4428 for lower values of $\alpha$.
1242 gezelter 4401
1243     At short distances, the presence of the ions creates a strong local
1244     field that acts to align nearby dipoles nearly perfectly in opposition
1245     to the field from the ions. This has the effect of increasing the
1246     effective screening when the ions are brought close to one another.
1247 gezelter 4428 This effect is present even in the full Ewald treatment, and indicates
1248     that the local ordering behavior is being captured by all of the
1249     moderately-damped real-space methods.
1250 gezelter 4401
1251 gezelter 4432 \subsubsection{Distance-dependent Kirkwood factors}
1252     One of the most sensitive measures of dipolar ordering in a liquid is
1253     the disance dependent Kirkwood factor,
1254     \begin{equation}
1255     G_K(r) = \left< \frac{1}{N} \sum_{i} \sum_{\substack{j \\ r_{ij} < r}}
1256     \frac{\mathbf{D}_i \cdot \mathbf{D}_j}{\left| D_i \right| \left| D_j
1257     \right|} \right>
1258     \label{eq:kirkwood}
1259     \end{equation}
1260     which measures the net orientational (cosine) ordering of dipoles
1261     inside a sphere of radius $r$. The outer brackets denote a
1262     configurational average. Figure \ref{fig:kirkwood} shows $G_K(r)$ for
1263     the three real space methods with $r_c = 3.52 \sigma = 12$~\AA~ and for
1264     the Ewald sum. For SP and GSF, the underdamped cases exhibit the
1265     ``hole'' at $r_c$ that is sometimes seen in cutoff-based method
1266     simulations of liquid water,\cite{Mark:2002rt,Fukuda:2012fr} but for
1267     values of $\alpha > 0.225$ \AA$^{-1}$, agreement with the Ewald
1268     results is quite good.
1269     \begin{figure}
1270     \includegraphics[width=\linewidth]{kirkwood.eps}
1271     \caption{The distance-dependent Kirkwood factors of the dipolar system
1272     for the three real space methods at a range of Gaussian damping
1273     parameters ($\alpha$) with a cutoff $r_c = 3.52 \sigma$.}
1274     \label{fig:kirkwood}
1275     \end{figure}
1276     Note that like the dielectric constant, $G_K(r)$ can also be corrected
1277     using the expressions for $A$ in table \ref{tab:A}. This is discussed
1278     in more detail in the supplemental material.\cite{supplemental}
1279    
1280 mlamichh 4397 \subsection{Quadrupolar fluid}
1281     \begin{figure}
1282 gezelter 4428 \includegraphics[width=4in]{polarizabilityFinal_Quad.eps}
1283 gezelter 4401 \caption{The quadrupole polarizability ($\alpha_Q$), correction factor
1284     ($B$), and susceptibility ($\chi_Q$) for the test quadrupolar
1285     fluid. The left panels were computed using external field gradients,
1286     and those on the right are the result of equilibrium fluctuations.
1287     The GSF and SP methods allow nearly unmodified use of the
1288     ``conducting boundary'' or polarizability results in place of the
1289 gezelter 4404 bulk susceptibility.}
1290 mlamichh 4397 \label{fig:dielectricQuad}
1291     \end{figure}
1292 gezelter 4401 The polarizability ($\alpha_Q$), correction factor ($B$), and
1293     susceptibility ($\chi_Q$) for the quadrupolar fluid is plotted against
1294     damping parameter Fig. \ref{fig:dielectricQuad}. In quadrupolar
1295     fluids, both the polarizability and susceptibility have units of
1296 gezelter 4419 $\mathrm{length}^2$. Although the susceptibility has dimensionality, it
1297 gezelter 4401 is the relevant measure of macroscopic quadrupolar
1298 gezelter 4419 properties.\cite{JeonI03, JeonII03} The left panel in
1299 gezelter 4401 Fig. \ref{fig:dielectricQuad} shows results obtained from the applied
1300     field gradient simulations whereas the results from the equilibrium
1301     fluctuation formula are plotted in the right panels.
1302 mlamichh 4397
1303 gezelter 4401 The susceptibility for the quadrupolar fluid is obtained from
1304     quadrupolarizability and a correction factor using
1305     Eq. (\ref{eq:quadrupolarSusceptiblity}). The susceptibilities are
1306     shown in the bottom panels of Fig. \ref{fig:dielectricQuad}. All three
1307 gezelter 4432 methods: (SP, GSF, and TSF) produce small correction factors, $B$, so
1308     all show similar susceptibilities over the range of damping
1309     parameters. This shows that susceptibility derived using the
1310     quadrupolarizability and the correction factors are essentially
1311     independent of the electrostatic method utilized in the simulation.
1312 mlamichh 4397
1313 gezelter 4433 There is a notable difference in the dependence on $\alpha$ for the
1314     quadrupolar correction compared with the dipolar correction. This is
1315     due to the reduced range of the quadrupole-quadrupole interaction when
1316     compared with dipolar interactions. The effects of the Gaussian
1317     damping for dipoles are significant near the cutoff radius, which can
1318     be observed in Fig. \ref{fig:kirkwood}, while for quadrupoles, most of
1319     the interaction is naturally diminished by that point. Overdamping
1320     obscures orientational preferences, and so quadrupolar fluids can be
1321     safely simulated with smaller values of $\alpha$ than a similar
1322     dipolar fluid.
1323    
1324 mlamichh 4397 \begin{figure}
1325 gezelter 4428 \includegraphics[width=\linewidth]{screen_all.eps}
1326 gezelter 4401 \caption{\label{fig:screenQuad} The distance-dependent screening
1327     factor, $\epsilon(r)$, between two ions immersed in the quadrupolar
1328     fluid. Results from the perturbation and fluctuation methods are
1329 gezelter 4432 shown in left and central panels. Here the susceptibility is
1330 gezelter 4401 calculated from the simulation and the geometrical factor is
1331     evaluated analytically, using the field and field-gradient produced
1332     by ions. The right hand panel shows the screening factor obtained
1333     from the PMF calculations.}
1334 mlamichh 4397 \end{figure}
1335 gezelter 4401 A more difficult test of the quadrupolar susceptibility is made by
1336     comparing with direct calculation of the electrostatic screening using
1337     the potential of mean force (PMF). Since the effective dielectric
1338     constant for a quadrupolar fluid depends on the geometry of the field
1339     and field gradient, this is not a physical property of the quadrupolar
1340     fluid.
1341 mlamichh 4397
1342 gezelter 4401 The geometrical factor for embedded ions changes with the ion
1343     separation distance. It is therefore reasonable to treat the
1344     dielectric constant as a distance-dependent screening factor. Since
1345     the quadrupolar molecules couple with the gradient of the field, the
1346     distribution of the quadrupoles will be inhomogeneously distributed
1347     around the point charges. Hence the distribution of quadrupolar
1348     molecules should be taken into account when computing the geometrical
1349     factors in the presence of this perturbation,
1350     \begin{eqnarray}
1351     G &=& \frac{\int_V g(\mathbf{r}) \left|\nabla \mathbf{E}^\circ
1352     \right|^2 d\mathbf{r}}{\int_V\left|\mathbf{E}^\circ\right|^2
1353     d\mathbf{r}} \nonumber \\ \nonumber \\
1354     &=& \frac{ 2 \pi \int_{-1}^{1} \int_{0}^{R} r^2 g(r,
1355 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}}
1356 gezelter 4401 \label{eq:geometricalFactor}
1357     \end{eqnarray}
1358     where $g(r,\cos\theta)$ is a distribution function for the quadrupoles
1359     with respect to an origin at midpoint of a line joining the two probe
1360     charges.
1361    
1362     The effective screening factor is plotted against ion separation
1363     distance in Fig. \ref{fig:screenQuad}. The screening evaluated from
1364 gezelter 4433 the perturbation and fluctuation methods are shown in the left and
1365     central panels. Here the susceptibilities are calculated from bulk
1366     fluid simulations and the geometrical factors are evaluated using the
1367     field and field gradients produced by the ions. The field gradients
1368     have been weighted by the $g(r, \cos\theta)$ from the PMF calculations
1369     (Eq. (\ref{eq:geometricalFactor})). The right hand panel shows the
1370     screening factor obtained directly from the PMF calculations.
1371 gezelter 4401
1372     We note that the screening factor obtained from both the perturbation
1373 gezelter 4433 and fluctuation methods are in good agreement with each other at
1374     similar values of $\alpha$, and agree with Ewald for
1375     $\alpha = 0.2$~\AA$^{-1}$. The magnitude of these screening factors
1376     depends strongly on the $g(r, \cos\theta)$ weighting originating in
1377     the PMF calculations.
1378 gezelter 4401
1379 gezelter 4433 In Ewald-based simulations, the PMF calculations include interactions
1380     between periodic replicas of the ions, and there is a significant
1381     reduction in the screening factor because of this effect. Because the
1382     real-space methods do not include coupling to periodic replicas, both
1383     the magnitude and distance-dependent decay of the PMF are
1384     significantly larger. For moderate damping ($\alpha \sim 0.2-0.3$
1385     \AA$^{-1}$), screening factors for GSF, TSF, and SP are converging to
1386     similar values at large ion separations, and this value is the same as
1387     the large-separation estimate from the perturbation and fluctuation
1388     simulations for $\alpha \sim 0.2$ \AA$^{-1}$. The PMF calculations
1389     also show signs of coalescence of the ionic solvation shells at
1390     separations smaller than 7~\AA. At larger separations, the
1391     $\alpha = 0.2$~\AA$^{-1}$ PMF calculations appear to be reproducing
1392     the bulk screening values. These results suggest that using either
1393     TSF or GSF with moderate damping is a relatively safe way to predict
1394     screening in quadrupolar fluids.
1395    
1396 gezelter 4404 \section{Conclusions}
1397     We have used both perturbation and fluctuation approaches to evaluate
1398     dielectric properties for dipolar and quadrupolar fluids. The static
1399     dielectric constant is the relevant bulk property for dipolar fluids,
1400     while the quadrupolar susceptibility plays a similar role for
1401     quadrupoles. Corrections to both the static dielectric constant and
1402     the quadrupolar susceptibility were derived for three new real space
1403     electrostatic methods, and these corrections were tested against a
1404     third measure of dielectric screening, the potential of mean force
1405     between two ions immersed in the fluids.
1406 gezelter 4401
1407 gezelter 4404 For the dipolar fluids, we find that the polarizability evaluated
1408     using the perturbation and fluctuation methods show excellent
1409     agreement, indicating that equilibrium calculations of the dipole
1410     fluctuations are good measures of bulk polarizability. One of the
1411     findings of the second paper in this series is that the moderately
1412     damped GSF and SP methods were most suitable for molecular dynamics
1413     and Monte Carlo simulations, respectively.\cite{PaperII}
1414 mlamichh 4353
1415 gezelter 4432 Dielectic properties like $\epsilon$ and $G_K(r)$ are sensitive probes
1416     of local treatment of electrostatic damping for the new real space
1417 gezelter 4433 methods, as well as for the Ewald sum. Choosing a Gaussian damping
1418 gezelter 4432 parameter ($\alpha$) in a reasonable range is essential for obtaining
1419     agreement between the electrostatic methods. Although correction
1420     formulae can approximate bulk properties from non-optimal parameter
1421     choices, these formulae are extraordinarly sensitive to the width of
1422     the Gaussian damping approximation. The newly-derived entries in table
1423     \ref{tab:A} can provide an effective minimum on the values of $\alpha$
1424     that should be used in simulations. With a cutoff radius of 12 \AA,
1425     $\alpha = 0.241$ \AA$^{-1}$ (SP) or $0.268$ \AA$^{-1}$ (GSF) would
1426     capture dielectric screening with reasonable fidelity. The sensitivity
1427     of the dielectric screening is also observed in the effective
1428     screening of ions embedded in the fluid.
1429 mlamichh 4353
1430 gezelter 4432 With good choices of $\alpha$, the dielectric constant evaluated using
1431     the computed polarizability and correction factors agrees well with
1432     the previous Ewald-based simulation results.\cite{Adams81,NeumannI83}
1433 gezelter 4404 Although the TSF method alters many dynamic and structural features in
1434     multipolar liquids,\cite{PaperII} it is surprisingly good at computing
1435     bulk dielectric properties at nearly all ranges of the damping
1436     parameter. In fact, the correction factor, $A = 1$, for the TSF
1437     method so the conducting boundary formula is essentially correct when
1438     using this method for point dipolar fluids.
1439 mlamichh 4397
1440 gezelter 4404 As in the dipolar case, the quadpole polarizability evaluated from
1441 gezelter 4433 both perturbation and fluctuation simulations show good agreement,
1442     again confirming that equilibrium fluctuation calculations are
1443     sufficient to reproduce bulk dielectric properties in these fluids.
1444     The quadrupolar susceptibility calculated via our derived correction
1445     factors tends to produce the same result for all three real space
1446     methods. Similarly, with good choices of the damping parameter, the
1447     screening factor calculated using the susceptibility and a weighted
1448     geometric factor provides good agreement with results obtained
1449     directly via potentials of mean force. For quadrupolar fluids, the
1450     distance dependence of the electrostatic interaction is significantly
1451     reduced and the correction factors are all small. These points
1452     suggest that how an electrostatic method treats the cutoff radius
1453     become less consequential for higher order multipoles.
1454 gezelter 4404
1455     For this reason, we renew our recommendation that the
1456 gezelter 4433 moderately-damped ($\alpha = 0.24-0.25$~\AA$^{-1}$) GSF method is an
1457     excellent choice for molecular dynamics simulation where
1458     point-multipole interactions are being utilized to compute bulk
1459     properties of fluids.
1460 mlamichh 4353
1461 mlamichh 4412 \appendix
1462 gezelter 4413 \section{Contraction of the quadrupolar tensor with the traceless
1463     quadrupole moment }
1464 mlamichh 4412 \label{ap:quadContraction}
1465 gezelter 4413 For quadrupolar liquids modeled using point quadrupoles, the
1466     interaction tensor is shown in Eq. (\ref{quadRadial}). The Fourier
1467     transformation of this tensor for $ \mathbf{k} = 0$ is,
1468 mlamichh 4412 \begin{equation}
1469     \tilde{T}_{\alpha\beta\gamma\delta}(0) = \int_V T_{\alpha\beta\gamma\delta}(\mathbf{r}) d \mathbf{r}
1470     \end{equation}
1471 gezelter 4413 On the basis of symmetry, the 81 elements can be placed in four
1472     different groups: $\tilde{T}_{aaaa}$, $\tilde{T}_{aaab}$,
1473     $\tilde{T}_{aabb}$, and $\tilde{T}_{aabc}$, where $a$, $b$, and $c$,
1474 gezelter 4414 and can take on distinct values from the set $\left\{x, y, z\right\}$.
1475     The elements belonging to each of these groups can be obtained using
1476     permutations of the indices. Integration of all of the elements shows
1477     that only the groups with indices ${aaaa}$ and ${aabb}$ are non-zero.
1478 gezelter 4413
1479     We can derive values of the components of $\tilde{T}_{aaaa}$ and
1480     $\tilde{T}_{aabb}$ as follows;
1481 mlamichh 4412 \begin{eqnarray}
1482 gezelter 4413 \tilde{T}_{xxxx}(0) &=&
1483 mlamichh 4412 \int_{\textrm{V}}
1484 gezelter 4432 \left[ 3v_{41}(r)+6x^2v_{42}(r)/r^2 + x^4\,v_{43}(r)/r^4 \right] d\mathbf{r} \nonumber \\
1485 mlamichh 4412 &=&12\pi \int_0^{r_c}
1486 gezelter 4414 \left[ v_{41}(r)+\frac{2}{3} v_{42}(r) + \frac{1}{15}v_{43}(r) \right] r^2\,dr =
1487 gezelter 4413 \mathrm{12 \pi B}
1488 mlamichh 4412 \end{eqnarray}
1489     and
1490     \begin{eqnarray}
1491 gezelter 4413 \tilde{T}_{xxyy}(0)&=&
1492     \int_{\textrm{V}}
1493 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 \\
1494 gezelter 4413 &=&4\pi \int_0^{r_c}
1495 gezelter 4414 \left[ v_{41}(r)+\frac{2}{3} v_{42}(r) + \frac{1}{15}v_{43}(r) \right] r^2\,dr =
1496 gezelter 4413 \mathrm{4 \pi B}.
1497 mlamichh 4412 \end{eqnarray}
1498 gezelter 4413 These integrals yield the same values for all permutations of the
1499 gezelter 4432 indices in both tensor element groups. In Eq.
1500 gezelter 4413 \ref{fourierQuadZeroK}, for a particular value of the quadrupolar
1501     polarization $\tilde{\Theta}_{aa}$ we can contract
1502     $\tilde{T}_{aa\gamma\delta}(0)$ with $\tilde{\Theta}_{\gamma\delta}$,
1503     using the traceless properties of the quadrupolar moment,
1504 mlamichh 4412 \begin{eqnarray}
1505 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 \\
1506     &=& 12 \pi \mathrm{B}\tilde{\Theta}_{xx}(0) +
1507     4 \pi \mathrm{B}\tilde{\Theta}_{yy}(0) +
1508     4 \pi \mathrm{B}\tilde{\Theta}_{zz}(0) \nonumber \\
1509     &=& 8 \pi \mathrm{B}\tilde{\Theta}_{xx}(0) + 4 \pi
1510     \mathrm{B}\left(\tilde{\Theta}_{xx}(0)+\tilde{\Theta}_{yy}(0) +
1511     \tilde{\Theta}_{zz}(0)\right) \nonumber \\
1512     &=& 8 \pi \mathrm{B}\tilde{\Theta}_{xx}(0)
1513 mlamichh 4412 \end{eqnarray}
1514 gezelter 4413 Similarly for a quadrupolar polarization $\tilde{\Theta}_{xy}$ in
1515 gezelter 4432 Eq. \ref{fourierQuadZeroK}, we can contract
1516 gezelter 4413 $\tilde{T}_{xy\gamma\delta}(0)$ with $\tilde{\Theta}_{\gamma\delta}$,
1517     using the only surviving terms of the tensor,
1518 mlamichh 4412 \begin{eqnarray}
1519 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 \\
1520     &=& 4 \pi \mathrm{B}\tilde{\Theta}_{xy}(0) +
1521     4 \pi \mathrm{B}\tilde{\Theta}_{yx}(0) \nonumber \\
1522     &=& 8 \pi \mathrm{B}\tilde{\Theta}_{xy}(0)
1523 mlamichh 4412 \end{eqnarray}
1524 gezelter 4413 Here, we have used the symmetry of the quadrupole tensor to combine
1525     the symmetric terms. Therefore we can write matrix contraction for
1526     $\tilde{T}_{\alpha\beta\gamma\delta}(\mathrm{0})$ and
1527     $ \tilde{\Theta}_{\gamma\delta}(\mathrm{0})$ in a general form,
1528 mlamichh 4412 \begin{equation}
1529 gezelter 4413 \tilde{T}_{\alpha\beta\gamma\delta}(\mathrm{0})\tilde{\Theta}_{\gamma\delta}(\mathrm{0})
1530     = 8 \pi \mathrm{B} \tilde{\Theta}_{\alpha\beta}(\mathrm{0}),
1531 mlamichh 4412 \label{contract}
1532     \end{equation}
1533 gezelter 4413 which is the same as Eq. (\ref{quadContraction}).
1534 mlamichh 4412
1535 gezelter 4413 When the molecular quadrupoles are represented by point charges, the
1536     symmetry of the quadrupolar tensor is same as for point quadrupoles
1537     (see Eqs.~\ref{quadCharge} and \ref{quadRadial}). However, for
1538     molecular quadrupoles represented by point dipoles, the symmetry of
1539     the quadrupolar tensor must be handled separately (compare
1540 gezelter 4432 Eqs.~\ref{quadDip} and~\ref{quadRadial}). Although there is a
1541 gezelter 4414 difference in symmetry, the final result (Eq.~\ref{contract}) also holds
1542     true for dipolar representations.
1543 gezelter 4413
1544 gezelter 4414 \section{Quadrupolar correction factor for the Ewald-Kornfeld (EK)
1545     method}
1546     The interaction tensor between two point quadrupoles in the Ewald
1547     method may be expressed,\cite{Smith98,NeumannII83}
1548     \begin{align}
1549     {T}_{\alpha\beta\gamma\delta}(\mathbf{r}) = &\frac{4\pi}{V
1550     }\sum_{k\neq0}^{\infty}
1551     e^{-k^2 / 4
1552     \kappa^2} e^{-i\mathbf{k}\cdot
1553     \mathbf{r}} \left(\frac{r_\alpha r_\beta k_\delta k_\gamma}{k^2}\right) \nonumber \\
1554     &+ \left(\delta_{\alpha\beta}\delta_{\gamma\delta}+\delta_{\alpha\gamma}\delta_{\beta\delta}+\delta_{\alpha\delta}\delta_{\beta\gamma}\right)
1555     B_2(r) \nonumber \\
1556     &- \left(\delta_{\gamma\delta} r_\alpha r_\beta + \mathrm{ 5\; permutations}\right) B_3(r) \nonumber \\
1557     &+ \left(r_\alpha r_\beta r_\gamma r_\delta \right) B_4(r)
1558 mlamichh 4412 \label{ewaldTensor}
1559 gezelter 4414 \end{align}
1560     where $B_n(r)$ are radial functions defined in reference
1561     \onlinecite{Smith98},
1562     \begin{align}
1563     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) \\
1564     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) \\
1565     B_4(r) =& \frac{105}{r^9}\left(\frac{2r\kappa e^{-r^2
1566     \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}}
1567     + \frac{16r^7\kappa^7 e^{-r^2 \kappa^2}}{105\sqrt{\pi}} + \mathrm{erfc(\kappa r)} \right)
1568     \end{align}
1569    
1570     We can divide ${T}_{\alpha\beta\gamma\delta}(\mathbf{r})$ into three
1571     parts:
1572 mlamichh 4412 \begin{eqnarray}
1573 gezelter 4414 & & \mathbf{T}(\mathbf{r}) =
1574     \mathbf{T}^\mathrm{K}(\mathbf{r}) +
1575     \mathbf{T}^\mathrm{R1}(\mathbf{r}) +
1576     \mathbf{T}^\mathrm{R2}(\mathbf{r})
1577 mlamichh 4412 \end{eqnarray}
1578 gezelter 4414 where the first term is the reciprocal space portion. Since the
1579     quadrupolar correction factor $B = \tilde{T}_{abab}(0) / 4\pi$ and
1580 gezelter 4415 $\mathbf{k} = 0 $ is excluded from the reciprocal space sum,
1581 gezelter 4414 $\mathbf{T}^\mathrm{K}$ will not contribute.\cite{NeumannII83} The
1582     remaining terms,
1583     \begin{equation}
1584     \mathbf{T}^\mathrm{R1}(\mathbf{r}) = \mathbf{T}^\mathrm{bare}(\mathbf{r}) \left(\frac{2r\kappa e^{-r^2
1585     \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}}
1586     + \frac{16r^7\kappa^7 e^{-r^2 \kappa^2}}{105\sqrt{\pi}} + \mathrm{erfc(\kappa r)} \right)
1587     \end{equation}
1588 mlamichh 4412 and
1589     \begin{eqnarray}
1590 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 \\
1591     &-&\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)
1592 mlamichh 4412 \end{eqnarray}
1593 gezelter 4414 are contributions from the real space
1594     sum.\cite{Adams76,Adams80,Adams81} Here
1595     $\mathbf{T}^\mathrm{bare}(\mathbf{r})$ is the unmodified quadrupolar
1596     tensor (for undamped quadrupoles). Due to the angular symmetry of the
1597     unmodified tensor, the integral of
1598     $\mathbf{T}^\mathrm{R1}(\mathbf{r})$ will vanish when integrated over
1599     a spherical region. The only term contributing to the correction
1600     factor (B) is therefore
1601     $T^\mathrm{R2}_{\alpha\beta\gamma\delta}(\mathbf{r})$, which allows us
1602     to derive the correction factor for the Ewald-Kornfeld (EK) method,
1603 mlamichh 4412 \begin{eqnarray}
1604 gezelter 4414 \mathrm{B} &=& \frac{1}{4\pi} \int_V T^\mathrm{R2}_{abab}(\mathbf{r}) \nonumber \\
1605 gezelter 4432 &=& -\frac{8r_c^3 \kappa^5 e^{-\kappa^2 r_c^2}}{15\sqrt{\pi}}.
1606 mlamichh 4412 \end{eqnarray}
1607 gezelter 4428
1608     \begin{acknowledgments}
1609     Support for this project was provided by the National Science Foundation
1610     under grant CHE-1362211. Computational time was provided by the
1611 gezelter 4433 Center for Research Computing (CRC) at the University of Notre
1612     Dame. The authors would like to thank the reviewer for helpful
1613     comments and suggestions.
1614 gezelter 4428 \end{acknowledgments}
1615    
1616 mlamichh 4353 \newpage
1617 mlamichh 4397 \bibliography{dielectric_new}
1618 mlamichh 4353
1619     \end{document}