ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/multipole/dielectric_new.tex
Revision: 4401
Committed: Mon Mar 28 20:25:13 2016 UTC (8 years, 3 months ago) by gezelter
Content type: application/x-tex
File size: 73355 byte(s)
Log Message:
Latest updates

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 4396
16 mlamichh 4389 \newcolumntype{Y}{>{\centering\arraybackslash}X}
17 mlamichh 4353
18     \begin{document}
19    
20     \title{Real space electrostatics for multipoles. III. Dielectric Properties}
21    
22     \author{Madan Lamichhane}
23     \affiliation{Department of Physics, University
24     of Notre Dame, Notre Dame, IN 46556}
25     \author{Thomas Parsons}
26     \affiliation{Department of Chemistry and Biochemistry, University
27     of Notre Dame, Notre Dame, IN 46556}
28     \author{Kathie E. Newman}
29     \affiliation{Department of Physics, University
30     of Notre Dame, Notre Dame, IN 46556}
31     \author{J. Daniel Gezelter}
32     \email{gezelter@nd.edu.}
33     \affiliation{Department of Chemistry and Biochemistry, University
34     of Notre Dame, Notre Dame, IN 46556}
35    
36     \date{\today}% It is always \today, today,
37     % but any date may be explicitly specified
38    
39 mlamichh 4393 \begin{abstract}
40 gezelter 4399 In the first two papers in this series, we developed new shifted
41     potential (SP), gradient shifted force (GSF), and Taylor shifted
42     force (TSF) real-space methods for multipole interactions in
43     condensed phase simulations. Here, we discuss the dielectric
44     properties of fluids that emerge from simulations using these
45     methods. Most electrostatic methods (including the Ewald sum)
46     require correction to the conducting boundary fluctuation formula
47     for the static dielectric constants, and we discuss the derivation
48     of these corrections for the new real space methods. For quadrupolar
49     fluids, the analogous material property is the quadrupolar
50     susceptibility. As in the dipolar case, the fluctuation formula for
51     the quadrupolar susceptibility has corrections that depend on the
52     electrostatic method being utillized. One of the most important
53     effects measured by both the static dielectric and quadrupolar
54     susceptibility is the ability to screen charges embedded in the
55     fluid. We use potentials of mean force between solvated ions to
56     discuss how geometric factors can lead to distance-dependent
57     screening in both quadrupolar and dipolar fluids.
58 mlamichh 4353 \end{abstract}
59    
60     \maketitle
61    
62     \section{Introduction}
63    
64     Over the past several years, there has been increasing interest in
65 gezelter 4396 pairwise or ``real space'' methods for computing electrostatic
66     interactions in condensed phase
67 gezelter 4401 simulations.\cite{Wolf99,Zahn02,Kast03,Beckd05,Ma05,Wu:2005nr,Fennell06,Fukuda:2013uq,Wang:2016kx}
68 gezelter 4396 These techniques were initially developed by Wolf {\it et al.} in
69     their work towards an $\mathcal{O}(N)$ Coulombic sum.\cite{Wolf99}
70     Wolf's method of using cutoff neutralization and electrostatic damping
71     is able to obtain excellent agreement with Madelung energies in ionic
72 gezelter 4401 crystals.\cite{Wolf99}
73 mlamichh 4353
74 gezelter 4396 Zahn \textit{et al.}\cite{Zahn02} and Fennell and Gezelter extended
75     this method using shifted force approximations at the cutoff distance
76     in order to conserve total energy in molecular dynamics
77 gezelter 4399 simulations.\cite{Fennell06} Other recent advances in real-space
78 gezelter 4400 methods for systems of point charges have included explicit
79 gezelter 4401 elimination of the net multipole moments inside the cutoff sphere
80     around each charge site.\cite{Fukuda:2013uq,Wang:2016kx}
81 mlamichh 4353
82 gezelter 4399 In the previous two papers in this series, we developed three
83     generalized real space methods: shifted potential (SP), gradient
84     shifted force (GSF), and Taylor shifted force (TSF).\cite{PaperI,
85     PaperII} These methods evaluate electrostatic interactions for
86     charges and higher order multipoles using a finite-radius cutoff
87     sphere. The neutralization and damping of local moments within the
88     cutoff sphere is a multipolar generalization of Wolf's sum. In the
89     GSF and TSF methods, additional terms are added to the potential
90     energy so that forces and torques also vanish smoothly at the cutoff
91     radius. This ensures that the total energy is conserved in molecular
92     dynamics simulations.
93    
94 gezelter 4400 One of the most stringent tests of any new electrostatic method is the
95 gezelter 4390 fidelity with which that method can reproduce the bulk-phase
96     polarizability or equivalently, the dielectric properties of a
97 gezelter 4396 fluid. Before the advent of computer simulations, Kirkwood and Onsager
98     developed fluctuation formulae for the dielectric properties of
99 gezelter 4399 dipolar fluids.\cite{Kirkwood39,Onsagar36} Along with projections of
100     the frequency-dependent dielectric to zero frequency, these
101 gezelter 4400 fluctuation formulae are now widely used to predict the static
102     dielectric constant of simulated materials.
103 gezelter 4396
104 gezelter 4399 If we consider a system of dipolar or quadrupolar molecules under the
105     influence of an external field or field gradient, the net polarization
106     of the system will largely be proportional to the applied
107     perturbation.\cite{Chitanvis96, Stern-Feller03, SalvchovI14,
108     SalvchovII14} In simulations, the net polarization of the system is
109     also determined by the interactions \textit{between} the
110 gezelter 4400 molecules. Therefore the macroscopic polarizablity obtained from a
111 gezelter 4399 simulation depends on the details of the electrostatic interaction
112 gezelter 4400 methods that were employed in the simulation. To determine the
113 gezelter 4399 relevant physical properties of the multipolar fluid from the system
114     fluctuations, the interactions between molecules must be incorporated
115     into the formalism for the bulk properties.
116 gezelter 4396
117 gezelter 4400 In most simulations, bulk materials are treated using periodic
118     replicas of small regions, and this level of approximation requires
119     corrections to the fluctuation formulae that were derived for the bulk
120     fluids. In 1983 Neumann proposed a general formula for evaluating
121     dielectric properties of dipolar fluids using both Ewald and
122     real-space cutoff methods.\cite{NeumannI83} Steinhauser and Neumann
123     used this formula to evaluate the corrected dielectric constant for
124     the Stockmayer fluid using two different methods: Ewald-Kornfield (EK)
125     and reaction field (RF) methods.\cite{NeumannII83}
126 gezelter 4390
127 gezelter 4392 Zahn \textit{et al.}\cite{Zahn02} utilized this approach and evaluated
128     the correction factor for using damped shifted charge-charge
129     kernel. This was later generalized by Izvekov \textit{et
130 gezelter 4400 al.},\cite{Izvekov08} who noted that the expression for the
131 gezelter 4401 dielectric constant reduces to widely-used conducting boundary formula
132     for real-space methods that have first derivatives that vanish at the
133     cutoff radius.
134 gezelter 4390
135 gezelter 4400 One of the primary topics of this paper is the derivation of
136     correction factors for the three new real space methods. We find that
137     the correction formulae for dipolar molecules depends not only on the
138 gezelter 4399 methodology being used, but also on whether the molecular dipoles are
139     treated using point charges or point dipoles. We derive correction
140     factors for both cases.
141    
142 gezelter 4392 In quadrupolar fluids, the relationship between quadrupolar
143     susceptibility and the dielectric constant is not as straightforward
144 gezelter 4399 as in the dipolar case. The effective dielectric constant depends on
145     the geometry of the external (or internal) field
146     perturbation.\cite{Ernst92} Significant efforts have been made to
147     increase our understanding the dielectric properties of these
148 gezelter 4400 fluids,\cite{JeonI03,JeonII03,Chitanvis96} although a general
149     correction formula has not yet been developed.
150 gezelter 4390
151 gezelter 4392 In this paper we derive general formulae for calculating the
152 gezelter 4400 quadrupolar susceptibility of quadrupolar fluids. We also evaluate the
153     correction factor for SP, GSF, and TSF methods for quadrupolar fluids
154     interacting via point charges, point dipoles or directly through
155     quadrupole-quadrupole interactions.
156 gezelter 4390
157 gezelter 4400 We also calculate the screening behavior for two ions immersed in
158     multipolar fluids to estimate the distance dependence of charge
159     screening in both dipolar and quadrupolar fluids. We use three
160     distinct methods to compare our analytical results with computer
161     simulations:
162 gezelter 4390 \begin{enumerate}
163 gezelter 4396 \item responses of the fluid to external perturbations,
164 gezelter 4399 \item fluctuations of system multipole moments, and
165 gezelter 4396 \item potentials of mean force between solvated ions,
166 gezelter 4390 \end{enumerate}
167    
168 gezelter 4400 Under the influence of weak external fields, the bulk polarization of
169     the system is primarily a linear response to the perturbation, where
170     proportionality constant depends on the electrostatic interactions
171     between the multipoles. The fluctuation formulae connect bulk
172     properties of the fluid to equilibrium fluctuations in the system
173     multipolar moments during a simulation. These fluctuations also depend
174     on the form of the electrostatic interactions between molecules.
175     Therefore, the connections between the actual bulk properties and both
176     the computed fluctuation and external field responses must be modified
177     accordingly.
178 gezelter 4390
179 gezelter 4400 The potential of mean force (PMF) allows calculation of an effective
180 gezelter 4399 dielectric constant or screening factor from the potential energy
181     between ions before and after dielectric material is introduced.
182 gezelter 4400 Computing the PMF between embedded point charges is an additional
183     check on the bulk properties computed via the other two methods.
184 mlamichh 4363
185 gezelter 4400 \section{The Real-Space Methods}
186    
187 gezelter 4399 \section{Dipolar Fluids and the Dielectric Constant}
188 mlamichh 4353
189 gezelter 4399 Dielectric properties of a fluid arise mainly from responses of the
190     fluid to either applied fields or transient fields internal to the
191     fluid. In response to an applied field, the molecules have electronic
192     polarizabilities, changes to internal bond lengths, and reorientations
193     towards the direction of the applied field. There is an added
194     complication that in the presence of external field, the perturbation
195     experienced by any single molecule is not only due to external field
196     but also to the fields produced by the all other molecules in the
197     system.
198 mlamichh 4353
199 gezelter 4399 \subsection{Response to External Perturbations}
200    
201     In the presence of uniform electric field $\mathbf{E}$, an individual
202     molecule with a permanent dipole moment $p_o$ will realign along the
203     direction of the field with an average polarization given by
204 mlamichh 4353 \begin{equation}
205 gezelter 4400 \braket{\mathbf{p}} = \epsilon_0 \alpha_p \mathbf{E},
206 mlamichh 4353 \end{equation}
207 gezelter 4400 where $\alpha_p = {p_o}^2 / 3 \epsilon_0 k_B T$ is the contribution to
208     molecular polarizability due solely to reorientation dynamics.
209     Because the applied field must overcome thermal motion, the
210     orientational polarization depends inversely on the temperature.
211 mlamichh 4353
212 gezelter 4400 Likewise, a condensed phase system of permanent dipoles will also
213     polarize along the direction of an applied field. The polarization
214     density of the system is
215 mlamichh 4353 \begin{equation}
216 gezelter 4399 \textbf{P} = \epsilon_o \alpha_{D} \mathbf{E}.
217 mlamichh 4353 \label{pertDipole}
218     \end{equation}
219 gezelter 4400 The constant $\alpha_D$ is the macroscopic polarizability, which is an
220     emergent property of the dipolar fluid. Note that the polarizability,
221     $\alpha_D$ is distinct from the dipole susceptibility, $\chi_D$,
222     which is the quantity most directly related to the static dielectric
223     constant, $\epsilon = 1 + \chi_D$.
224 mlamichh 4363
225 gezelter 4399 \subsection{Fluctuation Formula}
226 gezelter 4400 % Using the Ewald sum under tin-foil boundary conditions, $\epsilon$ can
227     % be calculated for systems of orientationally-polarized molecules,
228     % \begin{equation}
229     % \epsilon = 1 + \frac{\langle M^2\rangle}{3\epsilon_0k_\textrm{B}TV},
230     % \label{eq:staticDielectric}
231     % \end{equation}
232     % where $\epsilon_0$ is the permittivity of free space and
233     % $\langle M^2\rangle$ is the fluctuation of the system dipole
234     % moment.\cite{Allen89} The numerator in the fractional term in
235     % equation (\ref{eq:staticDielectric}) is identical to the quantity
236     % calculated in the finite-system Kirkwood $g$ factor ($G_k$):
237     % \begin{equation}
238     % G_k = \frac{\langle M^2\rangle}{N\mu^2},
239     % \label{eq:KirkwoodFactor}
240     % \end{equation}
241     % where $\mu$ is the dipole moment of a single molecule of the
242     % homogeneous system.\cite{NeumannI83,NeumannII83,Neumann84,Neumann85} The
243     % fluctuation term in both equation (\ref{eq:staticDielectric}) and
244     % (\ref{eq:KirkwoodFactor}) is calculated as follows,
245     % \begin{equation}
246     % \begin{split}
247     % \langle M^2\rangle &= \langle\bm{M}\cdot\bm{M}\rangle
248     % - \langle\bm{M}\rangle\cdot\langle\bm{M}\rangle \\
249     % &= \langle M_x^2+M_y^2+M_z^2\rangle
250     % - (\langle M_x\rangle^2 + \langle M_x\rangle^2
251     % + \langle M_x\rangle^2).
252     % \end{split}
253     % \label{eq:fluctBoxDipole}
254     % \end{equation}
255     % This fluctuation term can be accumulated during a simulation; however,
256     % it converges rather slowly, thus requiring multi-nanosecond simulation
257     % times.\cite{Horn04} In the case of tin-foil boundary conditions, the
258     % dielectric/surface term of the Ewald summation is equal to zero.
259 mlamichh 4353
260 gezelter 4400 For a system of dipolar molecules at thermal equilibrium, we can
261     define both a system dipole moment, $\mathbf{M} = \sum_i \mathbf{p}_i$
262     as well as a dipole polarization density,
263     $\mathbf{P} = \braket{\mathbf{M}}/V$.
264 mlamichh 4353
265 gezelter 4400 In the presence of applied field $\mathbf{E}$, the polarization
266     density can be expressed in terms of fluctuations in the net dipole
267     moment,
268 mlamichh 4353 \begin{equation}
269 gezelter 4400 \mathbf{P} = \epsilon_o \frac{\braket{\mathbf{M}^2}-{\braket{\mathbf{M}}}^2}{3 \epsilon_o V k_B T}\bf{E}
270 gezelter 4399 \label{flucDipole}
271 mlamichh 4353 \end{equation}
272 gezelter 4400 This has structural similarity with the Boltzmann average for the
273     polarization of a single molecule. Here
274     $ \braket{\mathbf{M}^2}-{\braket{\mathbf{M}}}^2$ measures fluctuations
275 gezelter 4401 in the net dipole moment,
276 mlamichh 4353 \begin{equation}
277 gezelter 4401 \langle \mathbf{M}^2 \rangle - \langle \mathbf{M} \rangle^2 =
278     \langle M_x^2+M_y^2+M_z^2 \rangle - \left( \langle M_x \rangle^2 +
279     \langle M_y \rangle^2 +
280     \langle M_z \rangle^2 \right).
281     \label{eq:flucDip}
282     \end{equation}
283     For the limiting case $\textbf{E} \rightarrow 0 $, the ensemble
284     average of both the net dipole moment $\braket{\mathbf{M}}$ and
285     dipolar polarization $\bf{P}$ tends to vanish but
286     $\braket{\mathbf{M}^2}$ does not. The macroscopic dipole
287     polarizability can therefore be written as,
288     \begin{equation}
289 gezelter 4400 \alpha_D = \frac{\braket{\mathbf{M}^2}-{\braket{\mathbf{M}}}^2}{3 \epsilon_o V k_B T}
290 mlamichh 4353 \end{equation}
291 gezelter 4400 This relationship between a macroscopic property of dipolar material
292     and microscopic fluctuations is true even when the applied field
293     $ \textbf{E} \rightarrow 0 $.
294    
295 gezelter 4399 \subsection{Correction Factors}
296 gezelter 4400
297     In the presence of a uniform external field $ \mathbf{E}^\circ$, the
298     total electric field at $\mathbf{r}$ depends on the polarization
299     density at all other points in the system,\cite{NeumannI83}
300 mlamichh 4353 \begin{equation}
301 gezelter 4400 \mathbf{E}(\mathbf{r}) = \mathbf{E}^\circ(\mathbf{r}) +
302     \frac{1}{4\pi\epsilon_o} \int d\mathbf{r}^\prime
303     \mathbf{T}(\mathbf{r}-\mathbf{r}^\prime)\cdot
304     {\mathbf{P}(\mathbf{r}^\prime)}.
305     \label{eq:localField}
306 mlamichh 4353 \end{equation}
307 gezelter 4400 $\mathbf{T}$ is the dipole interaction tensor connecting dipoles at
308     $\mathbf{r}^\prime$ with the point of interest ($\mathbf{r}$).
309 mlamichh 4353
310 gezelter 4400 In simulations of dipolar fluids, the molecular dipoles may be
311     represented either by closely-spaced point charges or by
312     point dipoles (see Fig. \ref{fig:stockmayer}).
313 mlamichh 4353 \begin{figure}
314     \includegraphics[width=3in]{DielectricFigure}
315     \caption{With the real-space electrostatic methods, the effective
316     dipole tensor, $\mathbf{T}$, governing interactions between
317     molecular dipoles is not the same for charge-charge interactions as
318     for point dipoles.}
319     \label{fig:stockmayer}
320     \end{figure}
321     In the case where point charges are interacting via an electrostatic
322     kernel, $v(r)$, the effective {\it molecular} dipole tensor,
323     $\mathbf{T}$ is obtained from two successive applications of the
324     gradient operator to the electrostatic kernel,
325 gezelter 4400 \begin{align}
326     \mathbf{T}_{\alpha \beta}(r) =& \nabla_\alpha \nabla_\beta
327     \left(v(r)\right) \\
328     =& \delta_{\alpha \beta}
329 mlamichh 4353 \left(\frac{1}{r} v^\prime(r) \right) + \frac{r_{\alpha}
330     r_{\beta}}{r^2} \left( v^{\prime \prime}(r) - \frac{1}{r}
331     v^{\prime}(r) \right)
332     \label{dipole-chargeTensor}
333 gezelter 4400 \end{align}
334 mlamichh 4353 where $v(r)$ may be either the bare kernel ($1/r$) or one of the
335     modified (Wolf or DSF) kernels. This tensor describes the effective
336 gezelter 4400 interaction between molecular dipoles ($\mathbf{D}$) in Gaussian units
337     as $-\mathbf{D} \cdot \mathbf{T} \cdot \mathbf{D}$.
338 mlamichh 4353
339 gezelter 4400 When utilizing any of the three new real-space methods for point
340     \textit{dipoles}, the tensor is explicitly constructed,
341 mlamichh 4353 \begin{equation}
342     \mathbf{T}_{\alpha \beta}(r) = \delta_{\alpha \beta} v_{21}(r) +
343     \frac{r_{\alpha} r_{\beta}}{r^2} v_{22}(r)
344     \label{dipole-diopleTensor}
345     \end{equation}
346     where the functions $v_{21}(r)$ and $v_{22}(r)$ depend on the level of
347 gezelter 4400 the approximation.\cite{PaperI,PaperII} Although the Taylor-shifted
348     (TSF) and gradient-shifted (GSF) models produce to the same $v(r)$
349     function for point charges, they have distinct forms for the
350     dipole-dipole interaction.
351 mlamichh 4353
352 gezelter 4400 Using a constitutive relation, the polarization density
353     $\mathbf{P}(\mathbf{r})$ is given by,
354 mlamichh 4353 \begin{equation}
355 gezelter 4400 \mathbf{P}(\mathbf{r}) = \epsilon_o \chi_D
356     \left(\mathbf{E}^\circ(\mathbf{r}) + \frac{1}{4\pi\epsilon_o} \int
357     d\mathbf{r}^\prime \mathbf{T}(\mathbf{r}-\mathbf{r}^\prime ) \cdot \mathbf{P}(\mathbf{r}^\prime)\right).
358 mlamichh 4353 \label{constDipole}
359     \end{equation}
360 gezelter 4400 Note that $\chi_D$ explicitly depends on the details of the dipole
361     interaction tensor. Neumann \textit{et al.}
362     \cite{NeumannI83,NeumannII83,Neumann84,Neumann85} derived an elegant
363     way to modify the fluctuation formula to correct for approximate
364     interaction tensors. This correction was derived using a Fourier
365     representation of the interaction tensor,
366     $\tilde{\mathbf{T}}(\mathbf{k})$, and involves the quantity,
367 mlamichh 4353 \begin{equation}
368 gezelter 4400 A = \frac{3}{4\pi}\tilde{\mathbf{T}}(0) = \frac{3}{4\pi} \int_V
369     d\mathbf{r} \mathbf{T}(\mathbf{r})
370 mlamichh 4353 \end{equation}
371 gezelter 4400 which is the $k \rightarrow 0$ limit of $\tilde{\mathbf{T}}$. Using
372     this quantity (originally called $Q$ in
373     refs. \onlinecite{NeumannI83,NeumannII83,Neumann84,Neumann85}), the
374     dielectric constant can be computed
375 mlamichh 4353 \begin{equation}
376 gezelter 4400 \epsilon = \frac{3+(A + 2)(\epsilon_{CB}-1)}{3 + (A -1)(\epsilon_{CB}-1)}
377     \label{correctionFormula}
378 mlamichh 4353 \end{equation}
379 gezelter 4400 where $\epsilon_{CB}$ is the widely-used conducting boundary
380     expression for the dielectric constant,
381 mlamichh 4353 \begin{equation}
382 gezelter 4400 \epsilon_{CB} = 1 + \frac{\braket{\bf{M}^2}-{\braket{\bf{M}}}^2}{3
383     \epsilon_o V k_B T}
384 mlamichh 4353 \label{correctionFormula}
385     \end{equation}
386 gezelter 4400 Equation (\ref{correctionFormula}) allows estimation of the static
387     dielectric constant from fluctuations easily computed directly from
388     simulations.
389    
390    
391     % 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}
392     % \begin{equation}
393     % \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}
394     % \label{singular}
395     % \end{equation}
396     % Substituting equation (\ref{singular}) in the equation (\ref{constDipole}) we get,
397     % \begin{equation}
398     % \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).
399     % \end{equation}
400     % For both polarization and electric field homogeneous, this can be easily solved using Fourier transformation,
401     % \begin{equation}
402     % \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}).
403     % \end{equation}
404     % For homogeneous applied field Fourier component is non-zero only if $\kappa = 0$. Therefore,
405     % \begin{equation}
406     % \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}).
407     % \label{fourierDipole}
408     % \end{equation}
409     % 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,
410     % \begin{equation}
411     % \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}
412     % \end{equation}
413     % 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,
414     % \begin{equation}
415     % \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}
416     % \label{correctionFormula}
417     % \end{equation}
418    
419    
420     We have utilized the Neumann \textit{et al.} approach for thre three
421     new real-space methods, and obtain method-dependent correction
422     factors. The expression for the correction factor also depends on
423     whether the simulation involves point charges or point dipoles to
424     represent the molecular dipoles. These corrections factors are
425     listed in Table \ref{tab:A}.
426 mlamichh 4353 \begin{table}
427 gezelter 4400 \caption{Expressions for the dipolar correction factor ($A$) for the
428     real-space electrostatic methods in terms of the damping parameter
429 mlamichh 4353 ($\alpha$) and the cutoff radius ($r_c$). The Ewald-Kornfeld result
430 gezelter 4400 derived in Refs. \onlinecite{Adams80,Adams81,NeumannI83} is shown
431     for comparison using the Ewald convergence parameter ($\kappa$)
432     and the real-space cutoff value ($r_c$). }
433 mlamichh 4353 \label{tab:A}
434     {%
435 gezelter 4400 \begin{tabular}{l|c|c|c}
436 mlamichh 4353
437 gezelter 4400 Method & point charges & point dipoles \\
438 mlamichh 4353 \hline
439     Spherical Cutoff (SC) & $\mathrm{erf}(r_c \alpha) - \frac{2 \alpha r_c}{\sqrt{\pi}} e^{-\alpha^2 r_c^2}$ & $\mathrm{erf}(r_c \alpha) - \frac{2 \alpha r_c}{\sqrt{\pi}} e^{-\alpha^2 r_c^2}$ \\
440     Shifted Potental (SP) & $ \mathrm{erf}(r_c \alpha) - \frac{2 \alpha r_c}{\sqrt{\pi}} e^{-\alpha^2 r_c^2}$ & $\mathrm{erf}(r_c \alpha) -\frac{2 \alpha r_c}{\sqrt{\pi}}\left(1+\frac{2\alpha^2 {r_c}^2}{3} \right)e^{-\alpha^2{r_c}^2} $\\
441     Gradient-shifted (GSF) & 1 & $\mathrm{erf}(\alpha r_c)-\frac{2 \alpha r_c}{\sqrt{\pi}} \left(1 + \frac{2 \alpha^2 r_c^2}{3} + \frac{\alpha^4 r_c^4}{3}\right)e^{-\alpha^2 r_c^2} $ \\
442     Taylor-shifted (TSF) & 1 & 1 \\
443 gezelter 4400 Ewald-Kornfeld (EK) & $\mathrm{erf}(r_c \kappa) - \frac{2 \kappa
444     r_c}{\sqrt{\pi}} e^{-\kappa^2 r_c^2}$ & \\
445 mlamichh 4353 \end{tabular}%
446     }
447     \end{table}
448 gezelter 4400 Note that for point charges, the GSF and TSF methods produce estimates
449     of the dielectric that need no correction, and the TSF method likewise
450     needs no correction for point dipoles.
451 gezelter 4399
452     \section{Quadrupolar Fluids and the Quadrupolar Susceptibility}
453     \subsection{Response to External Perturbations}
454    
455 gezelter 4400 A molecule with a permanent quadrupole, $\mathsf{q}$, will align in
456     the presence of an electric field gradient $\nabla\mathbf{E}$. The
457     anisotropic polarization of the quadrupole is given
458     by,\cite{AduGyamfi78,AduGyamfi81}
459 gezelter 4399 \begin{equation}
460 gezelter 4400 \braket{\mathsf{q}} - \frac{\mathbf{I}}{3}
461     \mathrm{Tr}(\mathsf{q}) = \epsilon_o \alpha_q \nabla\mathbf{E},
462 gezelter 4399 \end{equation}
463 gezelter 4400 where $\alpha_q = q_o^2 / 15 \epsilon_o k_B T $ is a molecular quadrupole
464     polarizablity and $q_o$ is an effective quadrupole moment for the molecule,
465     \begin{equation}
466     q_o^2 = 3 \mathsf{q}:\mathsf{q} - \mathrm{Tr}(\mathsf{q})^2.
467     \end{equation}
468 gezelter 4399
469 gezelter 4400 In the presence of an external field gradient, a system of quadrupolar
470     molecules also organizes with an anisotropic polarization,
471 gezelter 4399 \begin{equation}
472 gezelter 4400 \mathsf{Q} - \frac{\mathbf{I}}{3} \mathrm{Tr}(\mathsf{Q}) = \epsilon_o
473     \alpha_Q \nabla\mathbf{E}
474     \end{equation}
475     where $\mathsf{Q}$ is the traced quadrupole density of the system and
476     $\alpha_Q$ is a macroscopic quadrupole polarizability which has
477     dimensions of $\mathrm{length}^{-2}$. Equivalently, the traceless form
478     may be used,
479     \begin{equation}
480     \mathsf{\Theta} = 3 \epsilon_o \alpha_Q \nabla\mathbf{E},
481 gezelter 4399 \label{pertQuad}
482     \end{equation}
483 gezelter 4400 where
484     $\mathsf{\Theta} = 3\mathsf{Q} - \mathbf{I} \mathrm{Tr}(\mathsf{Q})$
485     is the traceless tensor that also describes the system quadrupole
486     density. It is this tensor that will be utilized to derive correction
487     factors below.
488    
489 gezelter 4399 \subsection{Fluctuation Formula}
490 gezelter 4400 As in the dipolar case, we may define a system quadrupole moment,
491     $\mathsf{M}_Q = \sum_i \mathsf{q}_i$ and the traced quadrupolar
492     density, $\mathsf{Q} = \mathsf{M}_Q / V$. A fluctuation formula can
493     be written for a system comprising quadrupolar
494     molecules,\cite{LoganI81,LoganII82,LoganIII82}
495 gezelter 4399 \begin{equation}
496 gezelter 4400 \mathsf{Q} - \frac{\mathbf{I}}{3} \mathrm{Tr}(\mathsf{Q}) = \epsilon_o
497     \frac{\braket{\mathsf{M}_Q^2}-{\braket{\mathsf{M}_Q}}^2}{15 \epsilon_o
498     V k_B T} \cdot \nabla\mathbf{E}.
499 gezelter 4399 \label{flucQuad}
500     \end{equation}
501 gezelter 4400 Some care is needed in the definitions of the averaged quantities. These
502     refer to the effective quadrupole moment of the system, and they are
503     computed as follows,
504     \begin{align}
505     \braket{\mathsf{M}_Q^2} &= \braket{3 \mathsf{M}_Q:\mathsf{M}_Q -
506     \mathrm{Tr}(\mathsf{M}_Q)^2 }\\
507     \braket{\mathsf{M}_Q}^2 &= 3 \braket{\mathsf{M}_Q}:\braket{\mathsf{M}_Q} -
508     \mathrm{Tr}(\braket{\mathsf{M}_Q})^2
509 gezelter 4401 \label{eq:flucQuad}
510 gezelter 4400 \end{align}
511     The macroscopic quadrupolarizability is given by,
512 gezelter 4399 \begin{equation}
513 gezelter 4400 \alpha_Q = \frac{\braket{\mathsf{M}_Q^2}-{\braket{\mathsf{M}_Q}}^2}{15 \epsilon_o V k_B T}
514 gezelter 4399 \label{propConstQuad}
515     \end{equation}
516 gezelter 4400 This relationship between a macroscopic property of a quadrupolar
517     fluid and microscopic fluctuations should be true even when the
518     applied field gradient $\nabla \mathbf{E} \rightarrow 0$.
519 gezelter 4399
520     \subsection{Correction Factors}
521 gezelter 4400 In this section we generalize the treatment of Neumann \textit{et al.}
522     for quadrupolar fluids. Interactions involving multiple quadrupoles
523     are rank 4 tensors, and we therefore describe quantities in this
524     section using Einstein notation.
525    
526     In the presence of a uniform external field gradient,
527     $\partial_\alpha {E}^\circ_\beta $, the total field gradient at
528     $\mathbf{r}$ depends on the quadrupole polarization density at all
529     other points in the system,
530 mlamichh 4353 \begin{equation}
531 gezelter 4400 \partial_\alpha E_\beta(\mathbf{r}) = \partial_\alpha
532     {E}^\circ_\beta(\mathbf{r}) + \frac{1}{4\pi \epsilon_o}\int
533     T_{\alpha\beta\gamma\delta}({\mathbf{r}-\mathbf{r}^\prime})
534     Q_{\gamma\delta}(\mathbf{r}^\prime) d\mathbf{r}^\prime
535 mlamichh 4353 \label{gradMaxwell}
536     \end{equation}
537 gezelter 4400 where and $T_{\alpha\beta\gamma\delta}$ is the quadrupole interaction
538     tensor connecting quadrupoles at $\mathbf{r}^\prime$ with the point of
539     interest ($\mathbf{r}$).
540 mlamichh 4353
541 gezelter 4400 \begin{figure}
542     \includegraphics[width=3in]{QuadrupoleFigure}
543     \caption{With the real-space electrostatic methods, the molecular
544     quadrupole tensor, $\mathbf{T}_{\alpha\beta\gamma\delta}(r)$,
545     governing interactions between molecules is not the same for
546     quadrupoles represented via sets of charges, point dipoles, or by
547     single point quadrupoles.}
548     \label{fig:quadrupolarFluid}
549     \end{figure}
550    
551     In simulations of quadrupolar fluids, the molecular quadrupoles may be
552     represented by closely-spaced point charges, by multiple point
553     dipoles, or by a single point quadrupole (see
554     Fig. \ref{fig:quadrupolarFluid}). In the case where point charges are
555     interacting via an electrostatic kernel, $v(r)$, the effective
556     molecular quadrupole tensor can obtained from four successive
557     applications of the gradient operator to the electrostatic kernel,
558     \begin{eqnarray}
559     T_{\alpha\beta\gamma\delta}(\mathbf{r}) &=& \nabla_\alpha \nabla_\beta
560     \nabla_\gamma \nabla_\delta v(r) \\
561     &=& \left(\delta_{\alpha\beta}\delta_{\gamma\delta} +
562     \delta_{\alpha\gamma}\delta_{\beta\delta}+
563     \delta_{\alpha\delta}\delta_{\beta\gamma}\right)\left(-\frac{v^\prime(r)}{r^3}
564     + \frac{v^{\prime \prime}(r)}{r^2}\right) \nonumber \\
565     & &+ \left(\delta_{\alpha\beta} r_\gamma r_\delta + 5 \mathrm{~permutations}
566     \right) \left(\frac{3v^\prime(r)}{r^5}-\frac{3v^{\prime \prime}(r)}{r^4} +
567     \frac{v^{\prime \prime \prime}(r)}{r^3}\right) \nonumber \\
568     & &+ r_\alpha r_\beta r_\gamma r_\delta
569     \left(-\frac{15v^\prime(r)}{r^7}+\frac{15v^{\prime \prime}(r)}{r^6}-\frac{6v^{\prime
570     \prime \prime}(r)}{r^5} + \frac{v^{\prime \prime \prime \prime}(r)}{r^4}\right),
571 mlamichh 4353 \label{quadCharge}
572 gezelter 4400 \end{eqnarray}
573     where $v(r)$ can either be the electrostatic kernel ($1/r$) or one of
574     the modified (Wolf or DSF) kernels.
575 mlamichh 4353
576 gezelter 4400 Similarly, when representing quadrupolar molecules with multiple point
577     dipoles, the molecular quadrupole interaction tensor can be obtained
578     using two successive applications of the gradient operator to the
579     dipole interaction tensor,
580     \begin{eqnarray}
581     T_{\alpha\beta\gamma\delta}(\mathbf{r}) &=& \nabla_\alpha \nabla_\beta
582     v_{\gamma\delta}(\mathbf{r}) \\
583     & = & \delta_{\alpha\beta}\delta_{\gamma\delta} \frac{v^\prime_{21}(r)}{r} +
584     \left(\delta_{\alpha\gamma}\delta_{\beta\delta}+
585     \delta_{\alpha\delta}\delta_{\beta\gamma}\right)\frac{v_{22}(r)}{r^2}
586     \nonumber\\
587     & &+ \delta_{\gamma\delta} r_\alpha r_\beta
588     \left(\frac{v^{\prime \prime}_{21}(r)}{r^2}-\frac{v^\prime_{21}(r)}{r^3} \right)\nonumber \\
589     & &+\left(\delta_{\alpha\beta} r_\gamma r_\delta +
590     \delta_{\alpha\gamma} r_\beta r_\delta +\delta_{\alpha\delta}
591     r_\gamma r_\beta + \delta_{\beta\gamma} r_\alpha r_\delta
592     +\delta_{\beta\delta} r_\alpha r_\gamma \right)
593     \left(\frac{v^\prime_{22}(r)}{r^3}-\frac{2v_{22}(r)}{r^4}\right)
594     \nonumber \\
595     & &+ r_\alpha r_\beta r_\gamma r_\delta
596     \left(\frac{v^{\prime
597     \prime}_{22}(r)}{r^4}-\frac{5v^\prime_{22}(r)}{r^5}+\frac{8v_{22}(r)}{r^6}\right),
598 mlamichh 4353 \label{quadDip}
599 gezelter 4400 \end{eqnarray}
600     where $v_{\gamma\delta}(\mathbf{r})$ is a dipole-dipole interaction
601     tensor that depends on the level of the
602     approximation.\cite{PaperI,PaperII} Similarly $v_{21}(r)$ and
603     $v_{22}(r)$ are the radial function for different real space cutoff
604     methods defined in the first paper in this series.\cite{PaperI}
605    
606     For quadrupolar liquids modeled using point quadrupoles, the
607     interaction tensor can be constructed as,
608     \begin{eqnarray}
609     T_{\alpha\beta\gamma\delta}(\mathbf{r}) & = &
610     \left(\delta_{\alpha\beta}\delta_{\gamma\delta}
611     +
612     \delta_{\alpha\gamma}\delta_{\beta\delta}+
613     \delta_{\alpha\delta}\delta_{\beta\gamma}\right)v_{41}(r)
614     + \delta_{\gamma\delta}
615     r_\alpha r_\beta
616     \frac{v_{42}(r)}{r^2}
617     \nonumber \\
618     & & + r_\alpha r_\beta r_\gamma r_\delta \left(\frac{v_{43}(r)}{r^4}\right),
619 mlamichh 4353 \label{quadRadial}
620 gezelter 4400 \end{eqnarray}
621     where again $v_{41}(r)$, $v_{42}(r)$, and $v_{43}(r)$ are radial
622     functions defined in Paper I of the series. \cite{PaperI} Note that
623     these radial functions have different functional forms depending on
624     the level of approximation being employed.
625    
626     The integral in equation (\ref{gradMaxwell}) can be divided into two
627     parts, $|\mathbf{r}-\mathbf{r}^\prime|\rightarrow 0 $ and
628     $|\mathbf{r}-\mathbf{r}^\prime|> 0$. Since the self-contribution to
629     the field gradient vanishes at the singularity (see Appendix
630     \ref{singularQuad}), equation (\ref{gradMaxwell}) can be written as,
631 mlamichh 4353 \begin{equation}
632 gezelter 4400 \partial_\alpha E_\beta(\mathbf{r}) = \partial_\alpha {E}^\circ_\beta(\mathbf{r}) +
633     \frac{1}{4\pi \epsilon_o}\int\limits_{|\mathbf{r}-\mathbf{r}^\prime|> 0 }
634     T_{\alpha\beta\gamma\delta}(\mathbf{r}-\mathbf{r}^\prime)
635     {Q}_{\gamma\delta}(\mathbf{r}^\prime) d\mathbf{r}^\prime.
636 mlamichh 4353 \end{equation}
637 gezelter 4400 If $\mathbf{r} = \mathbf{r}^\prime$ is excluded from the integration,
638     the total gradient can be most easily expressed in terms of
639     traceless quadrupole density as below,\cite{LoganI81}
640 mlamichh 4353 \begin{equation}
641 gezelter 4400 \partial_\alpha E_\beta(\mathbf{r}) = \partial_\alpha
642     {E}^\circ_\beta(\mathbf{r}) + \frac{1}{12\pi
643     \epsilon_o}\int\limits_{|\mathbf{r}-\mathbf{r}^\prime|> 0 }
644     T_{\alpha\beta\gamma\delta}(\mathbf{r}-\mathbf{r}^\prime) \Theta_{\gamma\delta}(\mathbf{r}') d\mathbf{r}^\prime,
645 mlamichh 4353 \end{equation}
646 gezelter 4400 where
647     $\Theta_{\alpha\beta} = 3Q_{\alpha\beta} - \delta_{\alpha\beta}Tr(Q)$
648     is the traceless quadrupole density. In analogy to
649     Eq. (\ref{pertQuad}) above, the quadrupolarization density may now be
650     related to the actual quadrupolar susceptibility, $\chi_Q$,
651 mlamichh 4353 \begin{equation}
652 gezelter 4400 \frac{1}{3}{\Theta}_{\alpha\beta}(\mathbf{r}) = \epsilon_o {\chi}_Q
653     \left(\partial_\alpha {E}^\circ_\beta(\mathbf{r}) + \frac{1}{12\pi
654     \epsilon_o}\int\limits_{|\mathbf{r}-\mathbf{r}^\prime|> 0 }
655     T_{\alpha\beta\gamma\delta}(\mathbf{r}-\mathbf{r}^\prime)
656     \Theta_{\gamma\delta}(\mathbf{r}^\prime) d\mathbf{r}^\prime \right).
657 mlamichh 4353 \end{equation}
658 gezelter 4400 For toroidal boundaries and with a uniform imposed
659     $\partial_\alpha E^\circ_\beta$, the quadrupole density
660     ${\Theta}_{\alpha\beta}$ will be uniform over the entire space. After
661     performing a Fourier transform (see the Appendix in
662     ref. \onlinecite{NeumannI83}) we obtain,
663 mlamichh 4353 \begin{equation}
664 gezelter 4400 \frac{1}{3}\tilde{\Theta}_{\alpha\beta}(\mathbf{k})=
665     \epsilon_o {\chi}_Q \left[{\partial_\alpha
666     \tilde{E}^\circ_\beta}(\mathbf{k})+ \frac{1}{12\pi
667     \epsilon_o} \tilde{T}_{\alpha\beta\gamma\delta}(\mathbf{k})
668     \tilde{\Theta}_{\gamma\delta}(\mathbf{k})\right].
669 mlamichh 4353 \end{equation}
670 gezelter 4400
671     The only surviving components of the quadrupole polarization will be
672     in the direction of the applied field gradient, so we may simplify
673     this expression considerably. That is, the only relevant polarization
674     term is $\tilde{\Theta}_{\alpha\beta}(\mathbf{k})$ and the only relevant
675     tensor contributions will have doubly-repeated indices, e.g.
676     $\tilde{T}_{\alpha\beta\alpha\beta}(\mathbf{k})$. Therefore we may
677     consider each component of the interaction tensor separately in a
678     scalar expression,
679     \begin{eqnarray}
680     \frac{1}{3}\tilde{\Theta}_{\alpha\beta}(\mathbf{k}) &=& \epsilon_o
681     {\chi}_Q
682     \left[{\partial_\alpha
683     \tilde{E}^\circ_\beta}(\mathbf{k})+
684     \frac{1}{12\pi
685     \epsilon_o}\tilde{T}_{\alpha\beta\alpha\beta}(\mathbf{k})
686     {\tilde{\Theta}}_{\alpha\beta}(\mathbf{k})\right]
687     \nonumber \\
688     &=& \epsilon_o {\chi}_Q \left(1-\frac{1}{4\pi} {\chi}_Q \tilde{T}_{\alpha\beta\alpha\beta}(\mathbf{k})\right)^{-1}
689     {\partial_\alpha \tilde{E}^\circ_\beta}(\mathbf{k})
690 mlamichh 4353 \label{fourierQuad}
691 gezelter 4400 \end{eqnarray}
692     If the applied field gradient is homogeneous over the
693     entire volume, ${\partial_ \alpha \tilde{E}^\circ_\beta}(\mathbf{k}) = 0 $ except at
694     $ \mathbf{k} = 0$. As in the dipolar case, the only relevant
695     contribution of the interaction tensor will also be when $\mathbf{k} = 0$.
696     Therefore equation (\ref{fourierQuad}) can be written as,
697 mlamichh 4353 \begin{equation}
698 gezelter 4400 \frac{1}{3}{\tilde{\Theta}}_{\alpha\beta}({0}) = \epsilon_o {\chi}_Q
699     \left(1-\frac{1}{4\pi} {\chi}_Q \tilde{T}_{\alpha\beta\alpha\beta}({0})\right)^{-1} \partial_\alpha \tilde{E}^\circ_\beta({0})
700 mlamichh 4353 \label{fourierQuad2}
701     \end{equation}
702 gezelter 4400 where $\tilde{T}_{\alpha\beta\alpha\beta}({0})$ can be evaluated as,
703 mlamichh 4353 \begin{equation}
704 gezelter 4400 \tilde{T}_{\alpha\beta\alpha\beta}({0}) = \int_V {T}_{\alpha\beta\alpha\beta}(\mathbf{r})d\mathbf{r}
705 mlamichh 4353 \label{realTensorQaud}
706     \end{equation}
707 gezelter 4400 We may now define a quantity to represent the contribution from the
708     $\mathbf{k} = 0$ portion of the interaction tensor,
709     \begin{equation}
710     B = \frac{1}{4 \pi} \int_V T_{\alpha \beta \alpha \beta}(\mathbf{r})
711     d\mathbf{r}
712     \end{equation}
713     which has been integrated over the interaction volume $V$ and has
714     units of $\mathrm{length}^{-2}$.
715 mlamichh 4353
716 gezelter 4400 In terms of traced quadrupole moment, equation (\ref{fourierQuad2})
717     can be written as,
718 mlamichh 4353 \begin{equation}
719 gezelter 4400 \mathsf{Q} - \frac{\mathbf{I}}{3} \mathrm{Tr}(\mathsf{Q})
720     = \epsilon_o {\chi}_Q\left(1- {\chi}_Q B \right)^{-1} \nabla \mathbf{E}^\circ
721 mlamichh 4353 \label{tracedConstQuad}
722     \end{equation}
723 gezelter 4400 Comparing (\ref{tracedConstQuad}) and (\ref{flucQuad}) we obtain,
724 mlamichh 4353 \begin{equation}
725 gezelter 4400 \frac{\braket{\mathsf{M}_Q^2}-{\braket{\mathsf{M}_Q}}^2}{15 \epsilon_o
726     V k_B T} = \chi_Q \left(1 - \chi_Q B \right)^{-1},
727 mlamichh 4353 \end{equation}
728 gezelter 4400 or equivalently,
729 mlamichh 4353 \begin{equation}
730 gezelter 4400 \chi_Q = \frac{\braket{\mathsf{M}_Q^2}-{\braket{\mathsf{M}_Q}}^2}{15 \epsilon_o
731     V k_B T} \left(1 + B \frac{\braket{\mathsf{M}_Q^2}-{\braket{\mathsf{M}_Q}}^2}{15 \epsilon_o
732     V k_B T} \right)^{-1}
733     \label{eq:finalForm}
734 mlamichh 4353 \end{equation}
735 gezelter 4400 Eq. (\ref{eq:finalForm}) now expresses a bulk property (the
736     quadrupolar susceptibility, $\chi_Q$) in terms of a straightforward
737     fluctuation in the system quadrupole moment and a quadrupolar
738     correction factor ($B$). The correction factors depend on the cutoff
739     method being employed in the simulation, and these are listed in Table
740     \ref{tab:B}.
741 mlamichh 4353
742 gezelter 4400 In terms of the macroscopic quadrupole polarizability, $\alpha_Q$,
743     which may be thought of as the ``conducting boundary'' version of the
744     susceptibility,
745 mlamichh 4353 \begin{equation}
746 gezelter 4400 \chi_Q = \frac{\alpha_Q}{1 + B \alpha_Q}.
747 mlamichh 4353 \end{equation}
748 gezelter 4400 If an electrostatic method produces $B \rightarrow 0$, the computed
749     quadrupole polarizability and quadrupole susceptibility converge to
750     the same value.
751    
752     \begin{sidewaystable}
753     \caption{Expressions for the quadrupolar correction factor ($B$) for
754     the real-space electrostatic methods in terms of the damping
755     parameter ($\alpha$) and the cutoff radius ($r_c$). The units
756     of the correction factor are $ \mathrm{length}^{-2}$ for quadrupolar
757     fluids.}
758 mlamichh 4353 \label{tab:B}
759 gezelter 4400 \begin{tabular}{l|c|c|c|c}
760 mlamichh 4353
761 gezelter 4400 Method & charges & dipoles & quadrupoles \\\hline
762 mlamichh 4353 Spherical Cutoff (SC) & $ -\frac{8 \alpha^5 {r_c}^3}{3\sqrt{\pi}} e^{-\alpha^2 r_c^2}$ & $ -\frac{8 \alpha^5 {r_c}^3}{3\sqrt{\pi}} e^{-\alpha^2 r_c^2}$ & $ -\frac{8 {\alpha}^5 {r_c}^3}{3\sqrt{\pi}} e^{-\alpha^2 r_c^2}$ \\
763     Shifted Potental (SP) & $ -\frac{8 \alpha^5 {r_c}^3}{3\sqrt{\pi}} e^{-\alpha^2 r_c^2}$ & $- \frac{8 \alpha^5 {r_c}^3}{3\sqrt{\pi}} e^{-\alpha^2 r_c^2}$& $ -\frac{16 \alpha^7 {r_c}^5}{9\sqrt{\pi}} e^{-\alpha^2 r_c^2}$ \\
764     Gradient-shifted (GSF) & $- \frac{8 \alpha^5 {r_c}^3}{3\sqrt{\pi}} e^{-\alpha^2 r_c^2}$ & 0 & $-\frac{4{\alpha}^7{r_c}^5 }{9\sqrt{\pi}}e^{-\alpha^2 r_c^2}(-1+2\alpha ^2 r_c^2)$\\
765 gezelter 4400 Taylor-shifted (TSF) & $ -\frac{8 \alpha^5 {r_c}^3}{3\sqrt{\pi}}
766     e^{-\alpha^2 r_c^2}$ &
767     $\frac{4\;\mathrm{erfc(\alpha r_c)}}{{r_c}^2} + \frac{8 \alpha}{3\sqrt{\pi}r_c}e^{-\alpha^2 {r_c}^2}\left(3+ 2 \alpha^2 {r_c}^2 + \alpha^4 {r_c}^4\right) $ & $\frac{10\;\mathrm{erfc}(\alpha r_c )}{{r_c}^2} + \frac{4{\alpha}}{9\sqrt{\pi}{r_c}}e^{-\alpha^2 r_c^2}\left(45 + 30\alpha ^2 {r_c}^2 + 12\alpha^4 {r_c}^4 + 3\alpha^6 {r_c}^6 + 2 \alpha^8 {r_c}^8\right)$ \\
768     Ewald-Kornfeld (EK) & & &
769     \end{tabular}
770     \end{sidewaystable}
771 gezelter 4399
772     \section{Screening of Charges by Multipolar Fluids}
773 gezelter 4401 \label{sec:PMF}
774     In a dipolar fluid, the static dielectric constant is also a measure
775     of the ability of the fluid to screen charges from one another. A set
776     of point charges creates an inhomogeneous field in the fluid, and the
777     fluid responds to this field as if it was created externally or via
778     local polarization fluctuations. For this reason, the dielectric
779     constant can be used to estimate an effective potential between two
780     point charges ($C_i$ and $C_j$) embedded in the fluid,
781 gezelter 4399 \begin{equation}
782 gezelter 4401 U_\mathrm{effective} = \frac{C_i C_j}{4 \pi \epsilon_0 \epsilon
783     r_{ij}}.
784     \label{eq:effectivePot}
785 gezelter 4399 \end{equation}
786 gezelter 4401
787     The same set of point charges can also create an inhomogeneous field
788     \textit{gradient}, and this will cause a response in a quadrupolar
789     fluid that will also cause an effective screening. As discussed above,
790     however, the relevant phyiscal property in quadrupolar fluids is the
791     susceptibility, $\chi_Q$. The screening dielectric associated with
792     the quadrupolar susceptibility is defined as,\cite{Ernst92}
793 gezelter 4399 \begin{equation}
794 gezelter 4401 \epsilon = 1 + \chi_Q G = 1 + G \frac{\alpha_Q}{1 + \alpha_Q B}
795     \label{eq:dielectricFromQuadrupoles}
796 gezelter 4399 \end{equation}
797 gezelter 4401 where $G$ is a geometrical factor that depends on the geometry of the
798     field perturbation,
799     \begin{equation}
800     G = \frac{\int_V \left| \nabla \mathbf{E}^\circ \right|^2 d\mathbf{r}}
801     {\int_V \left|\mathbf{E}^\circ\right|^2 d\mathbf{r}}
802     \end{equation}
803     integrated over the interaction volume. Note that this geometrical
804     factor is also required to compute effective dielectric constants even
805     when the field gradient is homogeneous over the entire sample.
806 gezelter 4399
807 gezelter 4401 To measure effective screening in a multipolar fluid, we compute an
808     effective interaction potential, the potential of mean force (PMF),
809     between positively and negatively charged ions when they screened by
810     the intervening fluid. The PMF is obtained from a sequence of
811     simulations in which two ions are constrained to a fixed distance, and
812     the average constraint force to hold them at a fixed distance $r$ is
813     collected during a long simulation,\cite{Wilfred07}
814 gezelter 4400 \begin{equation}
815 gezelter 4401 w(r) = \int_{r_o}^{r}\left\langle \frac{\partial f}{\partial r^\prime}
816     \right\rangle_{r^\prime} dr^\prime + 2k_BT \ln\left(\frac{r}{r_o}\right) + w(r_o),
817     \label{eq:pmf}
818 gezelter 4400 \end{equation}
819 gezelter 4401 where $\braket{\partial f/\partial r^\prime}_{r^\prime}$ is the mean
820     constraint force required to hold the ions at distance $r^\prime$,
821     $2k_BT \log(r/r_o)$ is the Fixman factor,\cite{Fixman:1974fk} and
822     $r_o$ is a reference position (usually taken as a large separation
823     betwen the ions). If the dielectric constant is a good measure of the
824     screening at all inter-ion separations, we would expect $w(r)$ to have
825     the form in Eq. (\ref{eq:effectivePot}). Because real fluids are not
826     continuum dielectrics, the effective dielectric constant is a function
827     of the interionic separation,
828 gezelter 4400 \begin{equation}
829 gezelter 4401 \epsilon(r) = \frac{u_\mathrm{raw}(r) }{w(r)}
830 gezelter 4400 \end{equation}
831 gezelter 4401 where $u_\mathrm{raw}(r)$ is the direct charge-charge interaction
832     potential that is in use during the simulation. $\epsilon(r)$ may
833     vary considerably from the bulk estimates at short distances, although
834     it should converge to the bulk value as the separation between the
835     ions increases.
836 gezelter 4400
837 gezelter 4399 \section{Simulation Methodology}
838 gezelter 4401 To test the formalism developed in the preceding sections, we have
839     carried out computer simulations using three different techniques: i)
840     simulations in the presence of external fields, ii) equilibrium
841     calculations of box moment fluctuations, and iii) potentials of mean
842     force (PMF) between embedded ions. In all cases, the fluids were
843     composed of point multipoles protected by a Lennard-Jones potential.
844     The parameters used in the test systems are given in table
845     \ref{Tab:C}.
846 mlamichh 4353
847 mlamichh 4389 \begin{table}
848 gezelter 4401 \caption{\label{Tab:C}The parameters used in simulations to evaluate
849     the dielectric response of the new real-space methods.}
850 mlamichh 4389 \begin{tabularx}{\textwidth}{r|cc|YYccc|Yccc} \hline
851     & \multicolumn{2}{c|}{LJ parameters} &
852     \multicolumn{5}{c|}{Electrostatic moments} & & & & \\
853     Test system & $\sigma$& $\epsilon$ & $C$ & $D$ &
854     $Q_{xx}$ & $Q_{yy}$ & $Q_{zz}$ & mass & $I_{xx}$ & $I_{yy}$ &
855     $I_{zz}$ \\ \cline{6-8}\cline{10-12}
856     & (\AA) & (kcal/mol) & (e) & (debye) & \multicolumn{3}{c|}{(debye \AA)} & (amu) & \multicolumn{3}{c}{(amu
857     \AA\textsuperscript{2})} \\ \hline
858 gezelter 4401 Dipolar fluid & 3.41 & 0.2381 & - & 1.4026 &-&-&-& 39.948 & 11.613 & 11.613 & 0.0 \\
859     Quadrupolar fluid & 2.985 & 0.265 & - & - & 0.0 & 0.0 &-2.139 & 18.0153 & 43.0565 & 43.0565 & 0.0 \\
860 mlamichh 4389 \ce{q+} & 1.0 & 0.1 & +1 & - & - & - & - & 22.98 & - & - & - \\
861     \ce{q-} & 1.0 & 0.1 & -1 & - & - & - & - & 22.98 & - & - & - \\ \hline
862     \end{tabularx}
863     \end{table}
864 mlamichh 4363
865 gezelter 4401 The first of the test systems consists entirely of fluids of point
866     dipolar or quadrupolar molecules in the presence of constant field or
867     field gradients. Since there are no isolated charges within the
868     system, the divergence of the field will be zero, \textit{i.e.}
869     $\vec{\nabla} \cdot \mathbf{E} = 0$. This condition can be satisfied
870     by using the relatively simple applied potential as described in the
871     supporting information.
872 mlamichh 4353
873 gezelter 4401 When a constant electric field or field gradient is applied to the
874     system, the molecules align along the direction of the applied field,
875     and polarize to a degree determined both by the strength of the field
876     and the fluid's polarizibility. We have calculated ensemble averages
877     of the box dipole and quadrupole moments as a function of the strength
878     of the applied fields. If the fields are sufficiently weak, the
879     response is linear in the field strength, and one can easily compute
880     the polarizability directly from the simulations.
881 mlamichh 4389
882 gezelter 4401 The second set of test systems consists of equilibrium simulations of
883     fluids of point dipolar or quadrupolar molecules simulated in the
884     absence of any external perturbation. The fluctuation of the ensemble
885     averages of the box multipolar moment was calculated for each of the
886     multipolar fluids. The box multipolar moments were computed as simple
887     sums over the instantaneous molecular moments, and fluctuations in
888     these quantities were obtained from Eqs. (\ref{eq:flucDip}) and
889     (\ref{eq:flucQuad}). The macroscopic polarizabilities of the system at
890     a were derived using Eqs.(\ref{flucDipole}) and (\ref{flucQuad}).
891 mlamichh 4389
892 gezelter 4401 The final system consists of dipolar or quadrupolar fluids with two
893     oppositely charged ions embedded within the fluid. These ions are
894     constrained to be at fixed distance throughout a simulation, although
895     they are allowed to move freely throughout the fluid while satisfying
896     that constraint. Separate simulations were run at a range of
897     constraint distances. A dielectric screening factor was computed using
898     the ratio between the potential between the two ions in the absence of
899     the fluid medium and the PMF obtained from the simulations.
900 mlamichh 4389
901 gezelter 4401 We carried out these simulations for all three of the new real-space
902     electrostatic methods (SP, GSF, and TSF) that were developed in the
903     first paper (Ref. \onlinecite{PaperI}) in the series. The radius of
904     the cutoff sphere was taken to be 12~\AA. Each of the real space
905     methods also depends on an adjustable damping parameter $\alpha$ (in
906     units of $\mathrm{length}^{-1}$). We have selected ten different
907     values of damping parameter: 0.0, 0.05, 0.1, 0.15, 0.175, 0.2, 0.225,
908     0.25, 0.3, and 0.35~\AA$^{-1}$ in our simulations of the dipolar
909     liquids, while four values were chosen for the quadrupolar fluids:
910     0.0, 0.1, 0.2, and 0.3~\AA$^{-1}$.
911 mlamichh 4389
912 gezelter 4401 For each of the methods and systems listed above, a reference
913     simulation was carried out using a multipolar implementation of the
914     Ewald sum.\cite{Smith82,Smith98} A default tolerance of
915     $1 \times 10^{-8}$~kcal/mol was used in all Ewald calculations,
916     resulting in Ewald coefficient 0.3119~\AA$^{-1}$ for a cutoff radius
917     of 12~\AA. All of the electrostatics and constraint methods were
918     implemented in our group's open source molecular simulation program,
919     OpenMD,\cite{Meineke05,openmd} which was used for all calculations in
920     this work.
921    
922     Dipolar systems contained 2048 Lennard-Jones-protected point dipolar
923     (Stockmayer) molecules with reduced density $\rho^* = 0.822$,
924     temperature $T^* = 1.15$, moment of inertia $I^* = 0.025$, and dipole
925     moment $\mu^* = \sqrt{3.0}$. These systems were equilibrated for
926     0.5~ns in the canonical (NVT) ensemble. Data collection was carried
927     out over a 1~ns simulation in the microcanonical (NVE) ensemble. Box
928     dipole moments were sampled every fs. For simulations with external
929     perturbations, field strengths ranging from
930     $0 - 10 \times 10^{-4} \mathrm{~V/\r{A}}$ with increments of
931     $ 10^{-4} \mathrm{~V/\r{A}}$ were carried out for each system.
932    
933     Quadrupolar systems contained 4000 linear point quadrupoles with a
934     density $2.338 \mathrm{~g/cm}^3$ at a temperature of 500~K. These
935     systems were equilibrated for 200~ps in a canonical (NVT) ensemble.
936     Data collection was carried out over a 500~ps simulation in the
937     microcanonical (NVE) ensemble. Components of box quadrupole moments
938     were sampled every 100 fs. For quadrupolar simulations with external
939     field gradients, field strengths ranging from
940     $0 - 9 \times 10^{-2} \mathrm{~V/\r{A}}^2$ with increments of
941     $10^{-2} \mathrm{~V/\r{A}^2}$ were carried out for each system.
942    
943     To carry out the PMF simulations, two of the multipolar molecules in
944     the test system were converted into \ce{q+} and \ce{q-} ions and
945     constrained to remain at a fixed distance for the duration of the
946     simulation. The constrained distance was then varied from
947     $5 - 12 \mathrm{~\r{A}}$. In the PMF calculations, all simulations were
948     equilibrated for 500 ps in the NVT ensemble and run for 5 ns in the
949     microcanonical (NVE) ensemble. Constraint forces were sampled every
950     20~fs.
951 mlamichh 4389
952     \section{Results}
953 mlamichh 4397 \subsection{Dipolar fluid}
954     \begin{figure}
955 gezelter 4401 \includegraphics[width=5in]{dielectricFinal_Dipole.pdf}
956     \caption{The polarizability ($\alpha_D$), correction factor ($A$), and
957     dielectric constant ($\epsilon$) for the test dipolar fluid. The
958     left panels were computed using external fields, and those on the
959     right are the result of equilibrium fluctuations. In the GSF and SP
960     methods, the corrections are large in with small values of $\alpha$,
961     and a optimal damping coefficient is evident around 0.25 \AA$^{-1}$.
962     The dashed lines in the upper panel indicate back-calculation of the
963     polarizability using the Ewald estimate
964     (Ref. \onlinecite{NeumannI83}) for the dielectric constant. NOTE:
965     Change $A_{dipole}$ to $A$.}
966 mlamichh 4397 \label{fig:dielectricDipole}
967     \end{figure}
968 gezelter 4401 The macroscopic polarizability ($\alpha_D$) for the dipolar fluid is
969     shown in the upper panels in Fig. \ref{fig:dielectricDipole}. The
970     polarizability obtained from the both perturbation and fluctuation
971     approaches are in excellent agreement with each other. The data also
972     show a stong dependence on the damping parameter for both the Shifted
973     Potential (SP) and Gradient Shifted force (GSF) methods, while Taylor
974     shifted force (TSF) is largely independent of the damping
975     parameter.
976 mlamichh 4397
977 gezelter 4401 The calculated correction factors discussed in section
978     \ref{sec:corrFactor} are shown in the middle panels. Because the TSF
979     method has $A = 1$ for all values of the damping parameter, the
980     computed polarizabilities need no correction for the dielectric
981     calculation. The value of $A$ varies with the damping parameter in
982     both the SP and GSF methods, and inclusion of the correction yields
983     dielectric estimates (shown in the lower panel) that are generally too
984     large until the damping reaches $\sim 0.25 \mathrm{~\r{A}}^{-1}$.
985     Above this value, the dielectric constants are generally in good
986     agreement with previous simulation results.\cite{NeumannI83}
987 mlamichh 4397
988 gezelter 4401 Figure \ref{fig:dielectricDipole} also contains back-calculations of
989     the polarizability using the reference (Ewald) simulation
990     results.\cite{NeumannI83} These are indicated with dashed lines in the
991     upper panels. It is clear that the expected polarizability for the SP
992     and GSF methods are quite close to results obtained from the
993     simulations. This indicates that the correction formula for the
994     dipolar fluid (Eq. \ref{correctionFormula}) is quite sensitive when
995     the value of $\mathrm{A}$ deviates significantly from unity.
996    
997     These results also suggest an optimal value for the damping parameter
998     of ($\alpha \sim 0.2-0.3 \mathrm{\r{A}}^{-1}$ when evaluating
999     dielectric constants of point dipolar fluids using either the
1000     perturbation and fluctuation approaches for the new real-space
1001     methods.
1002    
1003 mlamichh 4397 \begin{figure}
1004 gezelter 4401 \includegraphics[width=4in]{ScreeningFactor_Dipole.pdf}
1005     \caption{The distance-dependent screening factor, $\epsilon(r)$,
1006     between two ions immersed in the dipolar fluid. The new methods are
1007     shown in separate panels, and different values of the damping
1008     parameter ($\alpha$) are indicated with different symbols. All of
1009     the methods appear to be converging to the bulk dielectric constant
1010     ($\sim 65$) at large ion separations. CHECK PLOT}
1011 mlamichh 4397 \label{fig:ScreeningFactor_Dipole}
1012     \end{figure}
1013 gezelter 4401 We have also evaluated the distance-dependent screening factor,
1014     $\epsilon(r)$, between two oppositely charged ions when they are
1015     placed in the dipolar fluid. These results were computed using
1016     Eq. \ref{eq:pmf} and are shown in Fig.
1017     \ref{fig:ScreeningFactor_Dipole}.
1018 mlamichh 4397
1019 gezelter 4401 The screening factor is similar to the dielectric constant, but
1020     measures a local property of the ions in the fluid and depends on both
1021     ion-dipole and dipole-dipole interactions. These interactions depend
1022     on the distance between ions as well as the electrostatic interaction
1023     methods utilized in the simulations. The screening should converge to
1024     the dielectric constant when the field due to ions is small. This
1025     occurs when the ions are separated (or when the damping parameter is
1026     large). In Fig. \ref{fig:ScreeningFactor_Dipole} we observe that for
1027     the higher value of damping alpha \textit{i.e.}
1028     $\alpha = 0.2 \AA^{-1}$ and $0.3 \AA^{-1}$ and large separation
1029     between ions, the screening factor does indeed approach the correct
1030     dielectric constant.
1031 mlamichh 4397
1032 gezelter 4401 REVISIT THESE THREE PARAGRAPHS:
1033    
1034     It is also notable that the TSF method again displays smaller
1035     perturbations away from the correct dielectric screening behavior. We
1036     also observe that for TSF method yields high dielectric screening even
1037     for lower values of $\alpha$.
1038    
1039     At short distances, the presence of the ions creates a strong local
1040     field that acts to align nearby dipoles nearly perfectly in opposition
1041     to the field from the ions. This has the effect of increasing the
1042     effective screening when the ions are brought close to one another.
1043    
1044     This is due to the absence of corrections for dipole-dipole
1045     interactions as discussed above. However, the TSF remains quite
1046     perturbative for ion-dipole interactions, an effect that is most
1047     apparent at small ion separations.
1048    
1049 mlamichh 4397 \subsection{Quadrupolar fluid}
1050     \begin{figure}
1051 gezelter 4401 \includegraphics[width=4in]{polarizabilityFinal_Quad.pdf}
1052     \caption{The quadrupole polarizability ($\alpha_Q$), correction factor
1053     ($B$), and susceptibility ($\chi_Q$) for the test quadrupolar
1054     fluid. The left panels were computed using external field gradients,
1055     and those on the right are the result of equilibrium fluctuations.
1056     The GSF and SP methods allow nearly unmodified use of the
1057     ``conducting boundary'' or polarizability results in place of the
1058     bulk susceptibility. NOTE: Change $A_{quad}$ to $B$, change $\chi$
1059     to $\chi_Q$.}
1060 mlamichh 4397 \label{fig:dielectricQuad}
1061     \end{figure}
1062 gezelter 4401 The polarizability ($\alpha_Q$), correction factor ($B$), and
1063     susceptibility ($\chi_Q$) for the quadrupolar fluid is plotted against
1064     damping parameter Fig. \ref{fig:dielectricQuad}. In quadrupolar
1065     fluids, both the polarizability and susceptibility have units of
1066     $\mathrm{length}^2$ Although the susceptibility has dimensionality, it
1067     is the relevant measure of macroscopic quadrupolar
1068     properties.\cite{JeonI03, JeonII03}. The left panel in
1069     Fig. \ref{fig:dielectricQuad} shows results obtained from the applied
1070     field gradient simulations whereas the results from the equilibrium
1071     fluctuation formula are plotted in the right panels.
1072 mlamichh 4397
1073 gezelter 4401 The susceptibility for the quadrupolar fluid is obtained from
1074     quadrupolarizability and a correction factor using
1075     Eq. (\ref{eq:quadrupolarSusceptiblity}). The susceptibilities are
1076     shown in the bottom panels of Fig. \ref{fig:dielectricQuad}. All three
1077     methods: (SP, GSF, and TSF) produce similar susceptibilities over the
1078     range of damping parameters. This shows that susceptibility derived
1079     using the quadrupolarizability and the correction factors are
1080     essentially independent of the electrostatic method utilized in the
1081     simulation.
1082 mlamichh 4397
1083     \begin{figure}
1084 gezelter 4401 \includegraphics[width=5in]{ScreeningFactor_Quad.pdf}
1085     \caption{\label{fig:screenQuad} The distance-dependent screening
1086     factor, $\epsilon(r)$, between two ions immersed in the quadrupolar
1087     fluid. Results from the perturbation and fluctuation methods are
1088     shown in right and central panels. Here the susceptibility is
1089     calculated from the simulation and the geometrical factor is
1090     evaluated analytically, using the field and field-gradient produced
1091     by ions. The right hand panel shows the screening factor obtained
1092     from the PMF calculations.}
1093 mlamichh 4397 \end{figure}
1094 gezelter 4401 A more difficult test of the quadrupolar susceptibility is made by
1095     comparing with direct calculation of the electrostatic screening using
1096     the potential of mean force (PMF). Since the effective dielectric
1097     constant for a quadrupolar fluid depends on the geometry of the field
1098     and field gradient, this is not a physical property of the quadrupolar
1099     fluid.
1100 mlamichh 4397
1101 gezelter 4401 The geometrical factor for embedded ions changes with the ion
1102     separation distance. It is therefore reasonable to treat the
1103     dielectric constant as a distance-dependent screening factor. Since
1104     the quadrupolar molecules couple with the gradient of the field, the
1105     distribution of the quadrupoles will be inhomogeneously distributed
1106     around the point charges. Hence the distribution of quadrupolar
1107     molecules should be taken into account when computing the geometrical
1108     factors in the presence of this perturbation,
1109     \begin{eqnarray}
1110     G &=& \frac{\int_V g(\mathbf{r}) \left|\nabla \mathbf{E}^\circ
1111     \right|^2 d\mathbf{r}}{\int_V\left|\mathbf{E}^\circ\right|^2
1112     d\mathbf{r}} \nonumber \\ \nonumber \\
1113     &=& \frac{ 2 \pi \int_{-1}^{1} \int_{0}^{R} r^2 g(r,
1114     \cos\theta) \left|\nabla \mathbf{E}^\circ \right|^2 dr d\cos\theta }{\int_V\left|\mathbf{E}^\circ\right|^2 d\mathbf{r}}
1115     \label{eq:geometricalFactor}
1116     \end{eqnarray}
1117     where $g(r,\cos\theta)$ is a distribution function for the quadrupoles
1118     with respect to an origin at midpoint of a line joining the two probe
1119     charges.
1120    
1121     The effective screening factor is plotted against ion separation
1122     distance in Fig. \ref{fig:screenQuad}. The screening evaluated from
1123     the perturbation and fluctuation methods are shown in right and
1124     central panels. Here the susceptibility is calculated from the
1125     simulation and the geometrical factor is evaluated analytically, using
1126     the field and field-gradient produced by ions. The right hand panel
1127     shows the screening factor obtained from the PMF calculations.
1128    
1129     We note that the screening factor obtained from both the perturbation
1130     and fluctuation formula show good agreement with the screening factor
1131     calculated using PMF method. As there are no large differences in
1132     quadrupole-quadruople interactions for various real-space
1133     methods,\cite{PaperI,PaperII} we generally good agreement for the
1134     screening factors using any of the real space methods.
1135    
1136    
1137 gezelter 4399 \section{Discussion}
1138 gezelter 4401 We have used the perturbation and fluctuation formula to evaluate
1139     dielectric properties for dipolar and quadrupolar fluids implementing
1140     SP, GSF, and TSF methods in the simulation. Similarly the correction
1141     factors for the both dipolar and quadrupolar systems have also been
1142     calculated for all SP, GSF, and TSF methods. We have also derived
1143     screening factor between two ions immersed in the dipolar and
1144     quadrupolar fluids using PMF method. The result from the perturbation
1145     and fluctuation formula are compared with PMF method to test the
1146     accuracy of the simulation result.
1147 mlamichh 4353
1148 mlamichh 4397 For the dipolar fluid, the polarizability evaluated using the perturbation and fluctuation methods show excellent agreement with each other. The dielectric constant evaluated using polarizability and correction factor agree with the previous simulation results \cite{NeumannI83} for the damping parameter 0.25 - 0.3$\AA^{-1}$ and the electrostatic interaction methods, SP and GSF implemented in the simulation. Since the correction factor for TSF, $\mathrm{A_{dipole}} = 1$, it produces dielectric constant more closer to previous simulation results \cite{NeumannI83} for all values damping parameter. This method also produces best dielectric constant for damping parameter 0.15 - 0.25$\AA^{-1}$. We have also found that correction formula (equation ~\ref{correctionFormula}) is very sensitive, when the correction factor ($\mathrm{A_{dipole}}$) much away from the 1 for the SP and GSF methods. Therefore it is better to choose damping parameter from 0.25 - 0.3$\AA^{-1}$ to evaluate dielectric constant using SP and GSF methods. The screening factor between ions is found to be closer to dielectric constant when the distance between the ions are far apart ($\sim 9 \AA$) and damping parameter $0.2$ to $0.3 \AA^{-1}$.
1149 mlamichh 4353
1150 mlamichh 4397 The quadpolarizability evaluated from the both perturbation and fluctuation methods shows excellent agreement with each other for the case of quadrupolar fluid. The susceptibility is calculated from the quadpolarizability and correction factor tends to produce same result for all; SP, GSF, and TSF, methods. Similarly the screening factor calculated using susceptibility and geometrical factor shows excellent agreement with the result obtained from the PMF method.
1151    
1152    
1153 mlamichh 4353 \appendix
1154     \section{Point-multipolar interactions with a spatially-varying electric field}
1155    
1156     We can treat objects $a$, $b$, and $c$ containing embedded collections
1157     of charges. When we define the primitive moments, we sum over that
1158     collections of charges using a local coordinate system within each
1159     object. The point charge, dipole, and quadrupole for object $a$ are
1160     given by $C_a$, $\mathbf{D}_a$, and $\mathsf{Q}_a$, respectively.
1161     These are the primitive multipoles which can be expressed as a
1162     distribution of charges,
1163     \begin{align}
1164     C_a =&\sum_{k \, \text{in }a} q_k , \label{eq:charge} \\
1165     D_{a\alpha} =&\sum_{k \, \text{in }a} q_k r_{k\alpha}, \label{eq:dipole}\\
1166     Q_{a\alpha\beta} =& \frac{1}{2} \sum_{k \, \text{in } a} q_k
1167     r_{k\alpha} r_{k\beta} . \label{eq:quadrupole}
1168     \end{align}
1169     Note that the definition of the primitive quadrupole here differs from
1170     the standard traceless form, and contains an additional Taylor-series
1171     based factor of $1/2$. In Paper 1, we derived the forces and torques
1172     each object exerts on the others.
1173    
1174     Here we must also consider an external electric field that varies in
1175     space: $\mathbf E(\mathbf r)$. Each of the local charges $q_k$ in
1176     object $a$ will then experience a slightly different field. This
1177     electric field can be expanded in a Taylor series around the local
1178     origin of each object. A different Taylor series expansion is carried
1179     out for each object.
1180    
1181     For a particular charge $q_k$, the electric field at that site's
1182     position is given by:
1183     \begin{equation}
1184     E_\gamma + \nabla_\delta E_\gamma r_{k \delta}
1185     + \frac {1}{2} \nabla_\delta \nabla_\varepsilon E_\gamma r_{k \delta}
1186     r_{k \varepsilon} + ...
1187     \end{equation}
1188     Note that the electric field is always evaluated at the origin of the
1189     objects, and treating each object using point multipoles simplifies
1190     this greatly.
1191    
1192     To find the force exerted on object $a$ by the electric field, one
1193     takes the electric field expression, and multiplies it by $q_k$, and
1194     then sum over all charges in $a$:
1195    
1196     \begin{align}
1197     F_\gamma &= \sum_{k \textrm{~in~} a} q_k \lbrace E_\gamma + \nabla_\delta E_\gamma r_{k \delta}
1198     + \frac {1}{2} \nabla_\delta \nabla_\varepsilon E_\gamma r_{k \delta}
1199     r_{k \varepsilon} + ... \rbrace \\
1200     &= C_a E_\gamma + D_{a \delta} \nabla_\delta E_\gamma
1201     + Q_{a \delta \varepsilon} \nabla_\delta \nabla_\varepsilon E_\gamma +
1202     ...
1203     \end{align}
1204    
1205     Similarly, the torque exerted by the field on $a$ can be expressed as
1206     \begin{align}
1207     \tau_\alpha &= \sum_{k \textrm{~in~} a} (\mathbf r_k \times q_k \mathbf E)_\alpha \\
1208     & = \sum_{k \textrm{~in~} a} \epsilon_{\alpha \beta \gamma} q_k
1209     r_{k\beta} E_\gamma(\mathbf r_k) \\
1210     & = \epsilon_{\alpha \beta \gamma} D_\beta E_\gamma
1211     + 2 \epsilon_{\alpha \beta \gamma} Q_{\beta \delta} \nabla_\delta
1212     E_\gamma + ...
1213     \end{align}
1214    
1215     The last term is essentially identical with form derived by Torres del
1216     Castillo and M\'{e}ndez Garrido,\cite{Torres-del-Castillo:2006uo} although their derivation
1217     utilized a traceless form of the quadrupole that is different than the
1218     primitive definition in use here. We note that the Levi-Civita symbol
1219     can be eliminated by utilizing the matrix cross product in an
1220     identical form as in Ref. \onlinecite{Smith98}:
1221     \begin{equation}
1222     \left[\mathsf{A} \times \mathsf{B}\right]_\alpha = \sum_\beta
1223     \left[\mathsf{A}_{\alpha+1,\beta} \mathsf{B}_{\alpha+2,\beta}
1224     -\mathsf{A}_{\alpha+2,\beta} \mathsf{B}_{\alpha+1,\beta}
1225     \right]
1226     \label{eq:matrixCross}
1227     \end{equation}
1228     where $\alpha+1$ and $\alpha+2$ are regarded as cyclic permuations of
1229     the matrix indices. In table \ref{tab:UFT} we give compact
1230     expressions for how the multipole sites interact with an external
1231     field that has exhibits spatial variations.
1232    
1233     \begin{table}
1234     \caption{Potential energy $(U)$, force $(\mathbf{F})$, and torque
1235     $(\mathbf{\tau})$ expressions for a multipolar site embedded in an
1236     electric field with spatial variations, $\mathbf{E}(\mathbf{r})$.
1237     \label{tab:UFT}}
1238     \begin{tabular}{r|ccc}
1239     & Charge & Dipole & Quadrupole \\ \hline
1240     $U$ & $C \phi(\mathbf{r})$ & $-\mathbf{D} \cdot \mathbf{E}(\mathbf{r})$ & $- \mathsf{Q}:\nabla \mathbf{E}(\mathbf{r})$ \\
1241     $\mathbf{F}$ & $C \mathbf{E}(\mathbf{r})$ & $+\mathbf{D} \cdot \nabla \mathbf{E}(\mathbf{r})$ & $+\mathsf{Q} : \nabla\nabla\mathbf{E}(\mathbf{r})$ \\
1242     $\mathbf{\tau}$ & & $\mathbf{D} \times \mathbf{E}(\mathbf{r})$ & $+2 \mathsf{Q} \times \nabla \mathbf{E}(\mathbf{r})$
1243     \end{tabular}
1244     \end{table}
1245     \section{Gradient of the field due to quadrupolar polarization}
1246     \label{singularQuad}
1247     In this section, we will discuss the gradient of the field produced by
1248     quadrupolar polarization. For this purpose, we consider a distribution
1249     of charge ${\rho}(r)$ which gives rise to an electric field
1250     $\vec{E}(r)$ and gradient of the field $\vec{\nabla} \vec{E}(r)$
1251     throughout space. The total gradient of the electric field over volume
1252     due to the all charges within the sphere of radius $R$ is given by
1253     (cf. Jackson equation 4.14):
1254     \begin{equation}
1255     \int_{r<R} \vec{\nabla}\vec{E}\;d^3r = -\int_{r=R} R^2 \vec{E}\;\hat{n}\; d\Omega
1256     \label{eq:8}
1257     \end{equation}
1258     where $d\Omega$ is the solid angle and $\hat{n}$ is the normal vector
1259     of the surface of the sphere which is equal to
1260     $sin[\theta]cos[\phi]\hat{x} + sin[\theta]sin[\phi]\hat{y} +
1261     cos[\theta]\hat{z}$
1262     in spherical coordinates. For the charge density ${\rho}(r')$, the
1263     total gradient of the electric field can be written as (cf. Jackson
1264     equation 4.16),
1265     \begin{equation}
1266     \int_{r<R} \vec{\nabla}\vec{E}\; d^3r=-\int_{r=R} R^2\; \vec{\nabla}\Phi\; \hat{n}\; d\Omega =-\frac{1}{4\pi\;\epsilon_o}\int_{r=R} R^2\; \vec{\nabla}\;\left(\int \frac{\rho(r')}{|\vec{r}-\vec{r'}|}\;d^3r'\right) \hat{n}\; d\Omega
1267     \label{eq:9}
1268     \end{equation}
1269     The radial function in the equation (\ref{eq:9}) can be expressed in
1270     terms of spherical harmonics as (cf. Jackson equation 3.70),
1271     \begin{equation}
1272     \frac{1}{|\vec{r} - \vec{r'}|} = 4\pi \sum_{l=0}^{\infty}\sum_{m=-l}^{m=l}\frac{1}{2l+1}\;\frac{{r^l_<}}{{r^{l+1}_>}}\;{Y^*}_{lm}(\theta', \phi')\;Y_{lm}(\theta, \phi)
1273     \label{eq:10}
1274     \end{equation}
1275     If the sphere completely encloses the charge density then $ r_< = r'$ and $r_> = R$. Substituting equation (\ref{eq:10}) into (\ref{eq:9}) we get,
1276     \begin{equation}
1277     \begin{split}
1278     \int_{r<R} \vec{\nabla}\vec{E}\;d^3r &=-\frac{R^2}{\epsilon_o}\int_{r=R} \; \vec{\nabla}\;\left(\int \rho(r')\sum_{l=0}^{\infty}\sum_{m=-l}^{m=l}\frac{1}{2l+1}\;\frac{{r'^l}}{{R^{l+1}}}\;{Y^*}_{lm}(\theta', \phi')\;Y_{lm}(\theta, \phi)\;d^3r'\right) \hat{n}\; d\Omega \\
1279     &= -\frac{R^2}{\epsilon_o}\sum_{l=0}^{\infty}\sum_{m=-l}^{m=l}\frac{1}{2l+1}\;\int \rho(r')\;{r'^l}\;{Y^*}_{lm}(\theta', \phi')\left(\int_{r=R}\vec{\nabla}\left({R^{-(l+1)}}\;Y_{lm}(\theta, \phi)\right)\hat{n}\; d\Omega \right)d^3r
1280     '
1281     \end{split}
1282     \label{eq:11}
1283     \end{equation}
1284     The gradient of the product of radial function and spherical harmonics
1285     is given by (cf. Arfken, p.811 eq. 16.94):
1286     \begin{equation}
1287     \begin{split}
1288     \vec{\nabla}\left[ f(r)\;Y_{lm}(\theta, \phi)\right] = &-\left(\frac{l+1}{2l+1}\right)^{1/2}\; \left[\frac{\partial}{\partial r}-\frac{l}{r} \right]f(r)\; Y_{l, l+1, m}(\theta, \phi)\\ &+ \left(\frac{l}{2l+1}\right)^{1/2}\left[\frac
1289     {\partial}{\partial r}+\frac{l}{r} \right]f(r)\; Y_{l, l-1, m}(\theta, \phi).
1290     \end{split}
1291     \label{eq:12}
1292     \end{equation}
1293     Using equation (\ref{eq:12}) we get,
1294     \begin{equation}
1295     \vec{\nabla}\left({R^{-(l+1)}}\;Y_{lm}(\theta, \phi)\right) = [(l+1)(2l+1)]^{1/2}\; Y_{l,l+1,m}(\theta, \phi) \; \frac{1}{R^{l+2}},
1296     \label{eq:13}
1297     \end{equation}
1298     where $ Y_{l,l+1,m}(\theta, \phi)$ is the vector spherical harmonics
1299     which can be expressed in terms of spherical harmonics as shown in
1300     below (cf. Arfkan p.811),
1301     \begin{equation}
1302     Y_{l,l+1,m}(\theta, \phi) = \sum_{m_1, m_2} C(l+1,1,l|m_1,m_2,m)\; {Y_{l+1}}^{m_1}(\theta,\phi)\; \hat{e}_{m_2},
1303     \label{eq:14}
1304     \end{equation}
1305     where $C(l+1,1,l|m_1,m_2,m)$ is a Clebsch-Gordan coefficient and
1306     $\hat{e}_{m_2}$ is a spherical tensor of rank 1 which can be expressed
1307     in terms of Cartesian coordinates,
1308     \begin{equation}
1309     {\hat{e}}_{+1} = - \frac{\hat{x}+i\hat{y}}{\sqrt{2}},\quad {\hat{e}}_{0} = \hat{z},\quad and \quad {\hat{e}}_{-1} = \frac{\hat{x}-i\hat{y}}{\sqrt{2}}
1310     \label{eq:15}
1311     \end{equation}
1312     The normal vector $\hat{n} $ can be expressed in terms of spherical tensor of rank 1 as shown in below,
1313     \begin{equation}
1314     \hat{n} = \sqrt{\frac{4\pi}{3}}\left(-{Y_1}^{-1}{\hat{e}}_1 -{Y_1}^{1}{\hat{e}}_{-1} + {Y_1}^{0}{\hat{e}}_0 \right)
1315     \label{eq:16}
1316     \end{equation}
1317     The surface integral of the product of $\hat{n}$ and
1318     ${Y_{l+1}}^{m_1}(\theta, \phi)$ gives,
1319     \begin{equation}
1320     \begin{split}
1321     \int \hat{n}\;{Y_{l+1}}^{m_1}\;d\Omega &= \int \sqrt{\frac{4\pi}{3}}\left(-{Y_1}^{-1}{\hat{e}}_1 -{Y_1}^{1}{\hat{e}}_{-1} + {Y_1}^{0}{\hat{e}}_0 \right)\;{Y_{l+1}}^{m_1}\; d\Omega \\
1322     &= \int \sqrt{\frac{4\pi}{3}}\left({{Y_1}^{1}}^* {\hat{e}}_1 +{{Y_1}^{-1}}^* {\hat{e}}_{-1} + {{Y_1}^{0}}^* {\hat{e}}_0 \right)\;{Y_{l+1}}^{m_1}\; d\Omega \\
1323     &= \sqrt{\frac{4\pi}{3}}\left({\delta}_{l+1, 1}\;{\delta}_{1, m_1}\;{\hat{e}}_1 + {\delta}_{l+1, 1}\;{\delta}_{-1, m_1}\;{\hat{e}}_{-1}+ {\delta}_{l+1, 1}\;{\delta}_{0, m_1} \;{\hat{e}}_0\right),
1324     \end{split}
1325     \label{eq:17}
1326     \end{equation}
1327     where ${Y_{l}}^{-m} = (-1)^m\;{{Y_{l}}^{m}}^* $ and
1328     $ \int {{Y_{l}}^{m}}^*\;{Y_{l'}}^{m'}\;d\Omega =
1329     \delta_{ll'}\delta_{mm'} $.
1330     Non-vanishing values of equation \ref{eq:17} require $l = 0$,
1331     therefore the value of $ m = 0 $. Since the values of $ m_1$ are -1,
1332     1, and 0 then $m_2$ takes the values 1, -1, and 0, respectively
1333     provided that $m = m_1 + m_2$. Equation \ref{eq:11} can therefore be
1334     modified,
1335     \begin{equation}
1336     \begin{split}
1337     \int_{r<R} \vec{\nabla}\vec{E}\;d^3r = &- \sqrt{\frac{4\pi}{{3}}}\;\frac{1}{\epsilon_o}\int \rho(r')\;{Y^*}_{00}(\theta', \phi')[ C(1, 1, 0|-1,1,0)\;{\hat{e}_{-1}}{\hat{e}_{1}}\\ &+ C(1, 1, 0|-1,1,0)\;{\hat{e}_{1}}{\hat{e}_{-1}}+C(
1338     1, 1, 0|0,0,0)\;{\hat{e}_{0}}{\hat{e}_{0}} ]\; d^3r'.
1339     \end{split}
1340     \label{eq:18}
1341     \end{equation}
1342     After substituting ${Y^*}_{00} = \frac{1}{\sqrt{4\pi}} $ and using the
1343     values of the Clebsch-Gorden coefficients: $ C(1, 1, 0|-1,1,0) =
1344     \frac{1}{\sqrt{3}}, \; C(1, 1, 0|-1,1,0)= \frac{1}{\sqrt{3}}$ and $
1345     C(1, 1, 0|0,0,0) = -\frac{1}{\sqrt{3}}$ in equation \ref{eq:18} we
1346     obtain,
1347     \begin{equation}
1348     \begin{split}
1349     \int_{r<R} \vec{\nabla}\vec{E}\;d^3r &= -\sqrt{\frac{4\pi}{{3}}}\;\frac{1}{\epsilon_o}\int \rho(r')\;d^3r'\left({\hat{e}_{-1}}{\hat{e}_{1}}+{\hat{e}_{1}}{\hat{e}_{-1}}-{\hat{e}_{0}}{\hat{e}_{0}}\right)\\
1350     &= - \sqrt{\frac{4\pi}{{3}}}\;\frac{1}{\epsilon_o}\;C_{total}\;\left({\hat{e}_{-1}}{\hat{e}_{1}}+{\hat{e}_{1}}{\hat{e}_{-1}}-{\hat{e}_{0}}{\hat{e}_{0}}\right).
1351     \end{split}
1352     \label{eq:19}
1353     \end{equation}
1354     Equation (\ref{eq:19}) gives the total gradient of the field over a
1355     sphere due to the distribution of the charges. For quadrupolar fluids
1356     the total charge within a sphere is zero, therefore
1357     $ \int_{r<R} \vec{\nabla}\vec{E}\;d^3r = 0 $. Hence the quadrupolar
1358     polarization produces zero net gradient of the field inside the
1359     sphere.
1360    
1361 mlamichh 4389 \section{Applied field or field gradient}
1362     \label{Ap:fieldOrGradient}
1363 mlamichh 4353
1364 mlamichh 4389 To satisfy the condition $ \nabla . E = 0 $, within the box of molecules we have taken electrostatic potential in the following form
1365     \begin{equation}
1366     \begin{split}
1367     \phi(x, y, z) =\; &-g_o \left(\frac{1}{2}(a_1\;b_1 - \frac{cos\psi}{3})\;x^2+\frac{1}{2}(a_2\;b_2 - \frac{cos\psi}{3})\;y^2 + \frac{1}{2}(a_3\;b_3 - \frac{cos\psi}{3})\;z^2 \right. \\
1368     & \left. + \frac{(a_1\;b_2 + a_2\;b_1)}{2} x\;y + \frac{(a_1\;b_3 + a_3\;b_1)}{2} x\;z + \frac{(a_2\;b_3 + a_3\;b_2)}{2} y\;z \right),
1369     \end{split}
1370     \label{eq:appliedPotential}
1371     \end{equation}
1372     where $a = (a_1, a_2, a_3)$ and $b = (b_1, b_2, b_3)$ are basis vectors determine coefficients in x, y, and z direction. And $g_o$ and $\psi$ are overall strength of the potential and angle between basis vectors respectively. The electric field derived from the above potential is,
1373     \[\bf{E}
1374     =\frac{g_o}{2} \left(\begin{array}{ccc}
1375     2(a_1\; b_1 - \frac{cos\psi}{3})\;x \;+ (a_1\; b_2 \;+ a_2\; b_1)\;y + (a_1\; b_3 \;+ a_3\; b_1)\;z \\
1376     (a_2\; b_1 \;+ a_1\; b_2)\;x + 2(a_2\; b_2 \;- \frac{cos\psi}{3})\;y + (a_2\; b_3 \;+ a_3\; b_3)\;z \\
1377     (a_3\; b_1 \;+ a_3\; b_2)\;x + (a_3\; b_2 \;+ a_2\; b_3)y + 2(a_3\; b_3 \;- \frac{cos\psi}{3})\;z
1378     \end{array} \right).\]
1379     The gradient of the applied field derived from the potential can be written in the following form,
1380     \[\nabla\bf{E}
1381     = \frac{g_o}{2}\left(\begin{array}{ccc}
1382     2(a_1\; b_1 - \frac{cos\psi}{3}) & (a_1\; b_2 \;+ a_2\; b_1) & (a_1\; b_3 \;+ a_3\; b_1)\;z \\
1383     (a_2\; b_1 \;+ a_1\; b_2) & 2(a_2\; b_2 \;- \frac{cos\psi}{3}) & (a_2\; b_3 \;+ a_3\; b_3)\;z \\
1384     (a_3\; b_1 \;+ a_3\; b_2) & (a_3\; b_2 \;+ a_2\; b_3) & 2(a_3\; b_3 \;- \frac{cos\psi}{3})\;z
1385     \end{array} \right).\]
1386 mlamichh 4353 \newpage
1387    
1388 mlamichh 4397 \bibliography{dielectric_new}
1389 mlamichh 4353
1390     \end{document}