ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/multipole/dielectric_new.tex
Revision: 4399
Committed: Thu Mar 24 20:05:30 2016 UTC (8 years, 3 months ago) by gezelter
Content type: application/x-tex
File size: 69031 byte(s)
Log Message:
latest version

File Contents

# Content
1 \documentclass[aip,jcp,amsmath,amssymb,preprint]{revtex4-1}
2
3 \usepackage{graphicx}% Include figure files
4 \usepackage{dcolumn}% Align table columns on decimal point
5 \usepackage{multirow}
6 \usepackage{bm}
7 \usepackage{natbib}
8 \usepackage{times}
9 \usepackage{mathptmx}
10 \usepackage[version=3]{mhchem} % this is a great package for formatting chemical reactions
11 \usepackage{url}
12 \usepackage{braket}
13 \usepackage{tabularx}
14
15 \newcolumntype{Y}{>{\centering\arraybackslash}X}
16
17 \begin{document}
18
19 \title{Real space electrostatics for multipoles. III. Dielectric Properties}
20
21 \author{Madan Lamichhane}
22 \affiliation{Department of Physics, University
23 of Notre Dame, Notre Dame, IN 46556}
24 \author{Thomas Parsons}
25 \affiliation{Department of Chemistry and Biochemistry, University
26 of Notre Dame, Notre Dame, IN 46556}
27 \author{Kathie E. Newman}
28 \affiliation{Department of Physics, University
29 of Notre Dame, Notre Dame, IN 46556}
30 \author{J. Daniel Gezelter}
31 \email{gezelter@nd.edu.}
32 \affiliation{Department of Chemistry and Biochemistry, University
33 of Notre Dame, Notre Dame, IN 46556}
34
35 \date{\today}% It is always \today, today,
36 % but any date may be explicitly specified
37
38 \begin{abstract}
39 In the first two papers in this series, we developed new shifted
40 potential (SP), gradient shifted force (GSF), and Taylor shifted
41 force (TSF) real-space methods for multipole interactions in
42 condensed phase simulations. Here, we discuss the dielectric
43 properties of fluids that emerge from simulations using these
44 methods. Most electrostatic methods (including the Ewald sum)
45 require correction to the conducting boundary fluctuation formula
46 for the static dielectric constants, and we discuss the derivation
47 of these corrections for the new real space methods. For quadrupolar
48 fluids, the analogous material property is the quadrupolar
49 susceptibility. As in the dipolar case, the fluctuation formula for
50 the quadrupolar susceptibility has corrections that depend on the
51 electrostatic method being utillized. One of the most important
52 effects measured by both the static dielectric and quadrupolar
53 susceptibility is the ability to screen charges embedded in the
54 fluid. We use potentials of mean force between solvated ions to
55 discuss how geometric factors can lead to distance-dependent
56 screening in both quadrupolar and dipolar fluids.
57 \end{abstract}
58
59 \maketitle
60
61 \section{Introduction}
62
63 Over the past several years, there has been increasing interest in
64 pairwise or ``real space'' methods for computing electrostatic
65 interactions in condensed phase
66 simulations.\cite{Wolf99,Zahn02,Kast03,Beckd05,Ma05,Fennell06,ZMM papers}
67 These techniques were initially developed by Wolf {\it et al.} in
68 their work towards an $\mathcal{O}(N)$ Coulombic sum.\cite{Wolf99}
69 Wolf's method of using cutoff neutralization and electrostatic damping
70 is able to obtain excellent agreement with Madelung energies in ionic
71 crystals.\cite{Wolf99}
72
73 Zahn \textit{et al.}\cite{Zahn02} and Fennell and Gezelter extended
74 this method using shifted force approximations at the cutoff distance
75 in order to conserve total energy in molecular dynamics
76 simulations.\cite{Fennell06} Other recent advances in real-space
77 methods for systems of point charges have included explicit zeroing
78 out of the net multipole moments inside the cutoff spheres.\cite{ZMM
79 papers}
80
81 In the previous two papers in this series, we developed three
82 generalized real space methods: shifted potential (SP), gradient
83 shifted force (GSF), and Taylor shifted force (TSF).\cite{PaperI,
84 PaperII} These methods evaluate electrostatic interactions for
85 charges and higher order multipoles using a finite-radius cutoff
86 sphere. The neutralization and damping of local moments within the
87 cutoff sphere is a multipolar generalization of Wolf's sum. In the
88 GSF and TSF methods, additional terms are added to the potential
89 energy so that forces and torques also vanish smoothly at the cutoff
90 radius. This ensures that the total energy is conserved in molecular
91 dynamics simulations.
92
93 One of the most difficult tests of any new electrostatic method is the
94 fidelity with which that method can reproduce the bulk-phase
95 polarizability or equivalently, the dielectric properties of a
96 fluid. Before the advent of computer simulations, Kirkwood and Onsager
97 developed fluctuation formulae for the dielectric properties of
98 dipolar fluids.\cite{Kirkwood39,Onsagar36} Along with projections of
99 the frequency-dependent dielectric to zero frequency, these
100 fluctuation formulae are now widely used to predict dielectric
101 response of simulated materials.
102
103 If we consider a system of dipolar or quadrupolar molecules under the
104 influence of an external field or field gradient, the net polarization
105 of the system will largely be proportional to the applied
106 perturbation.\cite{Chitanvis96, Stern-Feller03, SalvchovI14,
107 SalvchovII14} In simulations, the net polarization of the system is
108 also determined by the interactions \textit{between} the
109 molecules. Therefore the macroscopic polarizablity obtained from the
110 simulation depends on the details of the electrostatic interaction
111 methods that were utilized in the simulation. To determine the
112 relevant physical properties of the multipolar fluid from the system
113 fluctuations, the interactions between molecules must be incorporated
114 into the formalism for the bulk properties.
115
116 In modern simulations, bulk materials are usually treated using
117 periodic replicas of small regions, and this level of approximation
118 requires corrections to the fluctuation formulae that were derived for
119 the bulk fluids. In 1983 Neumann proposed a general formula for
120 evaluating dielectric properties of dipolar fluids using real-space
121 cutoff methods.\cite{NeumannI83} Steinhauser and Neumann used this
122 formula to evaluate the corrected dielectric constant for the
123 Stockmayer fluid using two different methods: Ewald-Kornfield (EK) and
124 reaction field (RF) methods.\cite{NeumannII83}
125
126 Zahn \textit{et al.}\cite{Zahn02} utilized this approach and evaluated
127 the correction factor for using damped shifted charge-charge
128 kernel. This was later generalized by Izvekov \textit{et
129 al.},\cite{Izvekov08} who that the expression for the
130 dielectric constant reduces to widely-used \textit{conducting
131 boundary} formula for real-space cutoff methods that have first
132 derivatives that vanish at the cutoff sphere.
133
134 One of the primary topics of this paper will be derivation of these
135 correction factors for the new real space methods. We find that the
136 correction formulae for dipolar molecules depends not only on the
137 methodology being used, but also on whether the molecular dipoles are
138 treated using point charges or point dipoles. We derive correction
139 factors for both cases.
140
141 In quadrupolar fluids, the relationship between quadrupolar
142 susceptibility and the dielectric constant is not as straightforward
143 as in the dipolar case. The effective dielectric constant depends on
144 the geometry of the external (or internal) field
145 perturbation.\cite{Ernst92} Significant efforts have been made to
146 increase our understanding the dielectric properties of these
147 fluids,\cite{JeonI03,JeonII03,Chitanvis96} although a correction
148 formula for different real-space methods has not yet been developed.
149
150 In this paper we derive general formulae for calculating the
151 dielectric properties of quadrupolar fluids. We also evaluate the
152 correction factor for SP, GSF, and TSF methods for both dipolar and
153 quadrupolar fluids interacting via point charges, point dipoles or
154 through quadrupole-quadrupole interactions.
155
156 We have also calculated the screening behavior for two ions immersed
157 in multipolar fluids to estimate the distance dependence of dielectric
158 screening in both dipolar and quadrupolar fluids. We have used three
159 different methods to compare our results with computer simulations:
160 \begin{enumerate}
161 \item responses of the fluid to external perturbations,
162 \item fluctuations of system multipole moments, and
163 \item potentials of mean force between solvated ions,
164 \end{enumerate}
165
166 Using external perturbations, the net polarization of the system is
167 observed as a linear response of the applied field perturbation, where
168 proportionality constant is determined by the electrostatic
169 interaction between the electrostatic multipoles at a given
170 temperature. The fluctuation formula observes the time average
171 fluctuation of the multipolar moment as a function of temperature. The
172 average fluctuation value of the system is determined by the
173 multipole-multipole interactions between molecules at a given
174 temperature. Since the expression of the electrostatic interaction
175 energy, force, and torque in the real space electrostatic methods are
176 different from their original definition, both fluctuation and
177 external field perturbation formula must also be modified
178 accordingly.
179
180 The potential of mean force allows calculation of an effective
181 dielectric constant or screening factor from the potential energy
182 between ions before and after dielectric material is introduced.
183
184 \section{Dipolar Fluids and the Dielectric Constant}
185
186 Dielectric properties of a fluid arise mainly from responses of the
187 fluid to either applied fields or transient fields internal to the
188 fluid. In response to an applied field, the molecules have electronic
189 polarizabilities, changes to internal bond lengths, and reorientations
190 towards the direction of the applied field. There is an added
191 complication that in the presence of external field, the perturbation
192 experienced by any single molecule is not only due to external field
193 but also to the fields produced by the all other molecules in the
194 system.
195
196 \subsection{Response to External Perturbations}
197
198 In the presence of uniform electric field $\mathbf{E}$, an individual
199 molecule with a permanent dipole moment $p_o$ will realign along the
200 direction of the field with an average polarization given by
201 \begin{equation}
202 \braket{p_{mol}} = \epsilon_0 \alpha_p \mathbf{E},
203 \end{equation}
204 where $ \alpha_p = \frac{1}{3}\frac{{p_o}^2}{\epsilon_0 k_B T}$ is the
205 contribution to molecular polarizability due solely to reorientation
206 dynamics. The orientational polarization depends inversely on the
207 temperature as the applied field must overcome the thermal motion.
208
209 Likewise, a condensed phase system of dipolar molecules will also
210 polarize along the direction of an applied field. The net dipolar
211 polarization $\mathbf{P}$ of the system is,
212 \begin{equation}
213 \textbf{P} = \epsilon_o \alpha_{D} \mathbf{E}.
214 \label{pertDipole}
215 \end{equation}
216 The constant $\alpha_D$ is a macroscopic polarizability, which is an
217 emergent property of the dipolar fluid in a given density and
218 temperature and is related to the static dielectric constant.
219
220 \subsection{Fluctuation Formula}
221
222 Of particular interest is the static dielectric constant, $\epsilon$.
223 Using the Ewald sum under tin-foil boundary conditions, $\epsilon$ can
224 be calculated for systems of orientationally-polarized molecules via
225 \begin{equation}
226 \epsilon = 1 + \frac{\langle M^2\rangle}{3\epsilon_0k_\textrm{B}TV},
227 \label{eq:staticDielectric}
228 \end{equation}
229 where $\epsilon_0$ is the permittivity of free space and
230 $\langle M^2\rangle$ is the fluctuation of the system dipole
231 moment.\cite{Allen89} The numerator in the fractional term in
232 equation (\ref{eq:staticDielectric}) is identical to the quantity
233 calculated in the finite-system Kirkwood $g$ factor ($G_k$):
234 \begin{equation}
235 G_k = \frac{\langle M^2\rangle}{N\mu^2},
236 \label{eq:KirkwoodFactor}
237 \end{equation}
238 where $\mu$ is the dipole moment of a single molecule of the
239 homogeneous system.\cite{NeumannI83,NeumannII83,Neumann84,Neumann85} The
240 fluctuation term in both equation (\ref{eq:staticDielectric}) and
241 (\ref{eq:KirkwoodFactor}) is calculated as follows,
242 \begin{equation}
243 \begin{split}
244 \langle M^2\rangle &= \langle\bm{M}\cdot\bm{M}\rangle
245 - \langle\bm{M}\rangle\cdot\langle\bm{M}\rangle \\
246 &= \langle M_x^2+M_y^2+M_z^2\rangle
247 - (\langle M_x\rangle^2 + \langle M_x\rangle^2
248 + \langle M_x\rangle^2).
249 \end{split}
250 \label{eq:fluctBoxDipole}
251 \end{equation}
252 This fluctuation term can be accumulated during a simulation; however,
253 it converges rather slowly, thus requiring multi-nanosecond simulation
254 times.\cite{Horn04} In the case of tin-foil boundary conditions, the
255 dielectric/surface term of the Ewald summation is equal to zero.
256
257
258
259 For a system of molecules, the net dipole moment, $\bf{M}$ at thermal
260 equilibrium of temperature T in the presence of applied field
261 $\bf{E}$, the average dipolar polarization can be expressed in terms
262 of fluctuation of the net dipole moment,
263 \begin{equation}
264 \braket{\bf{P}} = \epsilon_o \frac{\braket{\bf{M}^2}-{\braket{\bf{M}}}^2}{3 \epsilon_o V k_B T}\bf{E}
265 \label{flucDipole}
266 \end{equation}
267 This is similar to the formula for Boltzmann average of single dipolar
268 molecule. Here $\braket{\bf{P}}$ is average polarization and
269 $ \braket{\textbf{M}^2}-{\braket{\textbf{M}}}^2$ is the net dipole
270 fluctuation at temperature T. For the limiting case
271 $\textbf{E} \rightarrow 0 $, the ensemble average of both the net
272 dipole moment $\braket{\textbf{M}}$ and dipolar polarization
273 $\braket{\bf{P}}$ tends to vanish but $\braket{\bf{M}^2}$ will still
274 be non-zero. The dipolar macroscopic polarizability can be written as,
275 \begin{equation}
276 \alpha_D = \frac{\braket{\bf{M}^2}-{\braket{\bf{M}}}^2}{3 \epsilon_o V k_B T}
277 \end{equation}
278 This is a macroscopic property of dipolar material which is true even
279 if applied field $ \textbf{E} \rightarrow 0 $.
280 \subsection{Correction Factors}
281 In the presence of an external field $ \textbf{E}$ polarization $\textbf{E}$ will be induced in a dipolar system. The total electrostatic field (or Maxwell electric field) at point $\bf{r}$ in a system is,\cite{NeumannI83}
282 \begin{equation}
283 \textbf{E}(\textbf{r}) = \textbf{E}^o(\textbf{r}) + \frac{1}{4\pi\epsilon_o} \int d^3r' \textbf{T}(\textbf{r}-\textbf{r}')\cdot {\textbf{P}(\textbf{r}')}.
284 \end{equation}
285
286 We can consider the cases of Stockmayer (dipolar) soft spheres that are represented either by two closely-spaced point charges or by a single point dipole (see Fig. \ref{fig:stockmayer}).
287 \begin{figure}
288 \includegraphics[width=3in]{DielectricFigure}
289 \caption{With the real-space electrostatic methods, the effective
290 dipole tensor, $\mathbf{T}$, governing interactions between
291 molecular dipoles is not the same for charge-charge interactions as
292 for point dipoles.}
293 \label{fig:stockmayer}
294 \end{figure}
295 In the case where point charges are interacting via an electrostatic
296 kernel, $v(r)$, the effective {\it molecular} dipole tensor,
297 $\mathbf{T}$ is obtained from two successive applications of the
298 gradient operator to the electrostatic kernel,
299 \begin{equation}
300 \mathbf{T}_{\alpha \beta}(r) = \nabla_\alpha \nabla_\beta \left(v(r)\right) = \delta_{\alpha \beta}
301 \left(\frac{1}{r} v^\prime(r) \right) + \frac{r_{\alpha}
302 r_{\beta}}{r^2} \left( v^{\prime \prime}(r) - \frac{1}{r}
303 v^{\prime}(r) \right)
304 \label{dipole-chargeTensor}
305 \end{equation}
306 where $v(r)$ may be either the bare kernel ($1/r$) or one of the
307 modified (Wolf or DSF) kernels. This tensor describes the effective
308 interaction between molecular dipoles ($\mathbf{D}$) in Gaussian
309 units as $-\mathbf{D} \cdot \mathbf{T} \cdot \mathbf{D}$.
310
311 When utilizing the new real-space methods for point dipoles, the
312 tensor is explicitly constructed,
313 \begin{equation}
314 \mathbf{T}_{\alpha \beta}(r) = \delta_{\alpha \beta} v_{21}(r) +
315 \frac{r_{\alpha} r_{\beta}}{r^2} v_{22}(r)
316 \label{dipole-diopleTensor}
317 \end{equation}
318 where the functions $v_{21}(r)$ and $v_{22}(r)$ depend on the level of
319 the approximation. Although the Taylor-shifted (TSF) and
320 gradient-shifted (GSF) models produce to the same $v(r)$ function for
321 point charges, they have distinct forms for the dipole-dipole
322 interactions.
323
324 Using constitutive relation, the polarization density $\textbf{P}(\textbf{r})$ is given by,
325 \begin{equation}
326 \textbf{P}(\textbf{r}) = \epsilon_o\; \chi^*_D \left(\textbf{E}^o(\textbf{r}) + \frac{1}{4\pi\epsilon_o} \int d^3r' \textbf{T}(\textbf{r}-\textbf{r}')\cdot {\textbf{P}(\textbf{r}')}\right).
327 \label{constDipole}
328 \end{equation}
329 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}
330 \begin{equation}
331 \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}
332 \label{singular}
333 \end{equation}
334 Substituting equation (\ref{singular}) in the equation (\ref{constDipole}) we get,
335 \begin{equation}
336 \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).
337 \end{equation}
338 For both polarization and electric field homogeneous, this can be easily solved using Fourier transformation,
339 \begin{equation}
340 \textbf{P}(\kappa) = 3 \epsilon_o\; \frac{\chi^*_D}{\chi^*_D + 3} \left(1- \frac{3}{4\pi}\;\frac{\chi^*_D}{\chi^*_D + 3}\; \textbf{T}({\kappa})\right)^{-1}\textbf{E}^o({\kappa}).
341 \end{equation}
342 For homogeneous applied field Fourier component is non-zero only if $\kappa = 0$. Therefore,
343 \begin{equation}
344 \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}).
345 \label{fourierDipole}
346 \end{equation}
347 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,
348 \begin{equation}
349 \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}
350 \end{equation}
351 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,
352 \begin{equation}
353 \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}
354 \label{correctionFormula}
355 \end{equation}
356 where $\epsilon_{CB}$ is dielectric constant obtained from conducting boundary condition. Equation (\ref{correctionFormula}) calculates actual dielectric constant from the dielectric constant obtained from the conducting boundary condition (which can be obtained directly from the simulation) using correction factor. The correction factor is different for different real-space cutoff methods. The expression for correction factor assuming a single point dipole or two closely spaced point charges for SP, GSF, and TSF method is listed in Table \ref{tab:A}.
357 \begin{table}
358 \caption{Expressions for the dipolar correction factor ($A$) for the real-space electrostatic methods in terms of the damping parameter
359 ($\alpha$) and the cutoff radius ($r_c$). The Ewald-Kornfeld result
360 derived in Refs. \onlinecite{Adams80,Adams81,NeumannI83} is shown for comparison using the Ewald
361 convergence parameter ($\kappa$) and the real-space cutoff value ($r_c$). }
362 \label{tab:A}
363 {%
364 \begin{tabular}{l|c|c|c|}
365
366 Method & $A_\mathrm{charges}$ & $A_\mathrm{dipoles}$ \\
367 \hline
368 Spherical Cutoff (SC) & $\mathrm{erf}(r_c \alpha) - \frac{2 \alpha r_c}{\sqrt{\pi}} e^{-\alpha^2 r_c^2}$ & $\mathrm{erf}(r_c \alpha) - \frac{2 \alpha r_c}{\sqrt{\pi}} e^{-\alpha^2 r_c^2}$ \\
369 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} $\\
370 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} $ \\
371 Taylor-shifted (TSF) & 1 & 1 \\
372 Ewald-Kornfeld (EK) & $\mathrm{erf}(r_c \kappa) - \frac{2 \kappa r_c}{\sqrt{\pi}} e^{-\kappa^2 r_c^2}$ & $\mathrm{erf}(r_c \kappa) - \frac{2 \kappa r_c}{\sqrt{\pi}} e^{-\kappa^2 r_c^2}$ \\\hline
373 \end{tabular}%
374 }
375 \end{table}
376
377
378 \section{Quadrupolar Fluids and the Quadrupolar Susceptibility}
379 \subsection{Response to External Perturbations}
380
381 In analogy to the dipolar case, a molecule with a permanent
382 quadrupole, $q_{\alpha\beta}$ will also align in the presence of a
383 uniform electric field gradient $\vec{\nabla}\mathbf{E}$. The average
384 polarization is given by
385 \begin{equation}
386 \braket{q_{\alpha\beta}} = \frac{1}{3} Tr(q^*) \delta_{\alpha\beta} + \alpha_q \partial_\alpha E_\beta,
387 \end{equation}
388 where $ \alpha_q = \frac{{\bar{q_o}}^2}{15k_BT} $ is a molecular
389 quadrupolarizablity and
390 ${\bar{q_o}}^2=
391 3{q}^*_{\alpha'\beta'}{q}^*_{\beta'\alpha'}-{q}^*_{\alpha'\alpha'}{q}^*_{\beta'\beta'}$
392 is a square of the net quadrupole moment of a molecule.
393
394 Similarly, in the presence of and external field gradient, a system of
395 quadrupolar molecule polarizes along the direction of applied field
396 gradient, giving a system quadrupolar polarization,
397 be given by,
398 \begin{equation}
399 \begin{split}
400 & {Q}_{\alpha\beta} = \frac{1}{3} Tr({Q}) \delta_{\alpha\beta} + \epsilon_o \alpha_Q \partial_{\alpha} E^o_{\beta}
401 \\ & or \\
402 & \frac{1}{3}\;\Theta_{\alpha\beta} = \epsilon_o\; \alpha_Q \; \partial_{\alpha} E^o_{\beta}
403 \end{split}
404 \label{pertQuad}
405 \end{equation}
406 where $Q_{\alpha\beta}$ is a tensor component of the traced quadrupolar moment of the system, $ \alpha_Q$ is a macroscopic quadrupolarizability has a dimension of $length^{-2}$, and $\Theta_{\alpha\beta} = 3Q_{\alpha\beta}-Tr(Q) $ is the traceless component of the quadrupole moment.
407 \subsection{Fluctuation Formula}
408
409 Fluctuation formulae for quadrupolar fluids were developed by Logan
410 \textit{et al.}\cite{LoganI81,LoganII82,LoganIII82}
411
412 In analogy with the dipolar case, a fluctuation formula can be
413 written for a system comprising quadrupolar molecules,
414 \begin{equation}
415 \braket{Q_{\alpha\beta}} = \frac{1}{3} Tr(\textbf{Q}) \delta_{\alpha\beta} + \epsilon_o \frac{\braket{\textbf{Q}^2}-{\braket{\textbf{Q}}}^2}{15 \epsilon_o V k_B T}{\partial_\alpha E^o_\beta}
416 \label{flucQuad}
417 \end{equation}
418 where $Q_{\alpha\beta}$ is a component of system quadrupole moment, $\bf{Q}$ is net quadrupolar moment which can be expressed as $\textbf{Q}^2 =3Q_{\alpha\beta}Q_{\alpha\beta}-(Tr\textbf{Q})^2 $. The macroscopic quadrupolarizability is given by,
419 \begin{equation}
420 \alpha_Q = \frac{\braket{\textbf{Q}^2}-{\braket{\textbf{Q}}}^2}{15 \epsilon_o V k_B T}
421 \label{propConstQuad}
422 \end{equation}
423
424 \subsection{Correction Factors}
425 In the presence of the field gradient $\partial_\alpha {E}_\beta $, a
426 non-vanishing quadrupolar polarization (quadrupole moment per unit
427 volume) $\bar{Q}_{\alpha\beta}$ will be induced in the entire volume
428 of a sample. The total field at any point $\vec{r}$ in the sample is
429 given by,
430 \begin{equation}
431 \partial_\alpha E_\beta(\textbf{r}) = \partial_\alpha {E^o}_\beta(\textbf{r}) + \frac{1}{4\pi \epsilon_o}\int T_{\alpha\beta\gamma\delta}(|{\textbf{r}-\textbf{r}'}|)\;{Q}_{\gamma\delta}(\textbf{r}')\; d^3r'
432 \label{gradMaxwell}
433 \end{equation}
434 where $\partial_\alpha {E^o}_\beta$ is the applied field gradient and $ T_{\alpha\beta\gamma\delta}$ is the quadrupole-quadrupole interaction tensor. We can represent quadrupole as a group of four closely spaced charges, two closely spaced point dipoles or single point quadrupole (see Fig. \ref{fig:quadrupolarFluid}). The quadrupole-quadrupole interaction tensor from the charge representation can obtained from the application of the four successive gradient operator to the electrostatic kernel $v(r)$.
435
436 \begin{equation}
437 \begin{split}
438 T_{\alpha\beta\gamma\delta}(r) &=\nabla_\alpha \nabla_\beta \nabla_\gamma \nabla_\delta\;v(r)
439 \\ &= \left(\delta_{\alpha\beta}\delta_{\gamma\delta} + \delta_{\alpha\gamma}\delta_{\beta\delta}+ \delta_{\alpha\delta}\delta_{\beta\gamma}\right)\left(-\frac{v'(r)}{r^3} + \frac{v''(r)}{r^2}\right)
440 \\ &+ \left(\delta_{\alpha\beta} r_\gamma r_\delta + 5 \; permutations \right) \left(\frac{3v'(r)}{r^5}-\frac{3v''(r)}{r^4} + \frac{v'''(r)}{r^3}\right)
441 \\ &+ r_\alpha r_\beta r_\gamma r_\delta\; \left(-\frac{15v'(r)}{r^7}+\frac{15v''(r)}{r^6}-\frac{6v'''(r)}{r^5} + \frac{v''''(r)}{r^4}\right),
442 \end{split}
443 \label{quadCharge}
444 \end{equation}
445 where $v(r)$ can either be electrostatic kernel for spherical truncation or one of the modified (Wolf or DSF) method. Similarly in point dipole representation the qaudrupole-quadrupole interaction tensor can be obtained from the applications of the two successive gradient in the dipole-dipole interaction tensor,
446
447 \begin{equation}
448 \begin{split}
449 T_{\alpha\beta\gamma\delta}(r) &=\nabla_\alpha \nabla_\beta \;v_{\gamma\delta}(r)
450 \\ &= \delta_{\alpha\beta}\delta_{\gamma\delta} \frac{v'_{21}(r)}{r} + \left(\delta_{\alpha\gamma}\delta_{\beta\delta}+ \delta_{\alpha\delta}\delta_{\beta\gamma}\right)\frac{v_{22}(r)}{r^2}
451 \\ &+ \delta_{\gamma\delta} r_\alpha r_\beta \left(\frac{v''_{21}(r)}{r^2}-\frac{v'_{21}(r)}{r^3} \right)
452 \\ &+\left(\delta_{\alpha\beta} r_\gamma r_\delta + \delta_{\alpha\gamma} r_\beta r_\delta +\delta_{\alpha\delta} r_\gamma r_\beta + \delta_{\beta\gamma} r_\alpha r_\delta +\delta_{\beta\delta} r_\alpha r_\gamma \right) \left(\frac{v'_{22}(r)}{r^3}-\frac{2v_{22}(r)}{r^4}\right)
453 \\ &+ r_\alpha r_\beta r_\gamma r_\delta\; \left(\frac{v''_{22}(r)}{r^4}-\frac{5v'_{22}(r)}{r^5}+\frac{8v_{22}(r)}{r^6}\right),
454 \end{split}
455 \label{quadDip}
456 \end{equation}
457 where $v_{\gamma\delta}(r)$ is the electrostatic dipole-dipole interaction tensor, which is different for different electrostatic cut off methods. Similarly $v_{21}(r) \;and\; v_{22}(r)$ are the radial function for different real space cutoff methods defined in Paper I of the series.\cite{PaperI} Using point quadrupole representation the quadrupole-quadrupole interaction can be constructed as,
458 \begin{equation}
459 \begin{split}
460 T_{\alpha\beta\gamma\delta}(r) &= \left(\delta_{\alpha\beta}\delta_{\gamma\delta} + \delta_{\alpha\gamma}\delta_{\beta\delta}+ \delta_{\alpha\delta}\delta_{\beta\gamma}\right)v_{41}(r) + \delta_{\gamma\delta} r_\alpha r_\beta \frac{v_{42}(r)}{r^2} \\ &+ r_\alpha r_\beta r_\gamma r_\delta\; \left(\frac{v_{43}(r)}{r^4}\right),
461 \end{split}
462 \label{quadRadial}
463 \end{equation}
464 where $v_{41}(r),\; v_{42}(r), \; \text{and} \; v_{43}(r)$ are defined in Paper I of the series. \cite{PaperI} They have different functional forms for different electrostatic cutoff methods.
465 \begin{figure}
466 \includegraphics[width=3in]{QuadrupoleFigure}
467 \caption{With the real-space electrostatic methods, the effective
468 quadrupolar tensor, $\mathbf{T}_{\alpha\beta\gamma\delta}(r)$, governing interactions between molecular quadrupoles can be represented by interaction of charges, point dipoles or single point quadrupoles.}
469 \label{fig:quadrupolarFluid}
470 \end{figure}
471 The integral in equation (\ref{gradMaxwell}) can be divided into two parts, $|\textbf{r}-\textbf{r}'|\rightarrow 0 $ and $|\textbf{r}-\textbf{r}'|> 0$. Since the total
472 field gradient due to quadrupolar fluid vanishes at the singularity (see Appendix \ref{singularQuad}), equation (\ref{gradMaxwell}) can be written as,
473 \begin{equation}
474 \partial_\alpha E_\beta(\textbf{r}) = \partial_\alpha {E^o}_\beta(\textbf{r}) +
475 \frac{1}{4\pi \epsilon_o}\int\limits_{|\textbf{r}-\textbf{r}'|> 0 }
476 T_{\alpha\beta\gamma\delta}(|\textbf{r}-\textbf{r}'|)\;{Q}_{\gamma\delta}(\textbf{r}')\;
477 d^3r'.
478 \end{equation}
479 If $\textbf{r} = \textbf{r}'$ is excluded from the integration, the gradient of the electric can be expressed in terms of traceless quadrupole moment as below, \cite{LoganI81}
480 \begin{equation}
481 \partial_\alpha E_\beta(\textbf{r}) = \partial_\alpha {E^o}_\beta(\textbf{r}) + \frac{1}{12\pi \epsilon_o}\int\limits_{|\textbf{r}-\textbf{r}'|> 0 } T_{\alpha\beta\gamma\delta}(|\textbf{r}-\textbf{r}'|)\;{\Theta}_{\gamma\delta}(\textbf{r}')\; d^3r',
482 \end{equation}
483 where $\Theta_{\alpha\beta} = 3Q_{\alpha\beta} - \delta_{\alpha\beta}Tr(Q)$
484 is the traceless quadrupole moment. The total quadrupolar polarization is written as,
485 \begin{equation}
486 {Q}_{\alpha\beta}(\textbf{r}) = \frac{1}{3}\delta_{\alpha\beta}\;Tr({Q})+\epsilon_o {\chi}^*_Q\;\partial_\alpha E_\beta(\textbf{r}),
487 \label{constQaud}
488 \end{equation}
489 In the equation (\ref{constQaud}), $\partial_{\alpha}E_{\beta}$ is Maxwell field gradient and ${\chi}^*_Q$ is the actual quadrupolar susceptibility of the fluid which is different than the proportionality constant $\chi_q $ in the equation (\ref{propConstQuad}). In terms of traceless quadrupole moment, equation (\ref{constQaud}) can be written as,
490 \begin{equation}
491 \frac{1}{3}{\Theta}_{\alpha\beta}(\textbf{r}) = \epsilon_o {\chi}^*_Q \; \partial_\alpha E_\beta (\textbf{r})= \epsilon_o {\chi}^*_Q \left(\partial_\alpha {E^o}_\beta(\textbf{r}) + \frac{1}{12\pi \epsilon_o}\int\limits_{|\textbf{r}-\textbf{r}'|> 0 } T_{\alpha\beta\gamma\delta}(|\textbf{r}-\textbf{r}'|)\;{\Theta}_{\gamma\delta}(\textbf{r}')\; d^3r'\right)
492 \end{equation}
493 For toroidal boundary conditions, both $\partial_\alpha E_\beta$ and
494 ${\Theta}_{\alpha\beta}$ are uniform over the entire space. After
495 performing a Fourier transform (see the Appendix in the Neumann's Paper \cite{NeumannI83}) we get,
496 \begin{equation}
497 \frac{1}{3}{{\Theta}}_{\alpha\beta}({\kappa})=
498 \epsilon_o {\chi}^*_Q \;\left[{\partial_\alpha
499 {E^o}_\beta}({\kappa})+ \frac{1}{12\pi
500 \epsilon_o}\;{T}_{\alpha\beta\gamma\delta}({\kappa})\;
501 {{\Theta}}_{\gamma\delta}({\kappa})\right]
502 \end{equation}
503 Since the quadrupolar polarization is in the direction of the applied
504 field, we can write
505 ${{\Theta}}_{\gamma\delta}({\kappa}) =
506 {{\Theta}}_{\alpha\beta}({\kappa})$
507 and
508 ${T}_{\alpha\beta\gamma\delta}({\kappa}) =
509 {T}_{\alpha\beta\alpha\beta}({\kappa})$. Therefore we can consider each component of the interaction tensor as scalar and perform calculation.
510 \begin{equation}
511 \begin{split}
512 \frac{1}{3}{{\Theta}}_{\alpha\beta}({\kappa}) &= \epsilon_o {\chi}^*_Q \left[{\partial_\alpha E^o_\beta}({\kappa})+ \frac{1}{12\pi \epsilon_o}{T}_{\alpha\beta\alpha\beta}({\kappa})\;{{\Theta}}_{\alpha\beta}({\kappa})\right] \\
513 &= \epsilon_o {\chi}^*_Q\;\left(1-\frac{1}{4\pi} {\chi}^*_Q\;
514 {T}_{\alpha\beta\alpha\beta}({\kappa})\right)^{-1}
515 {\partial_\alpha E^o_\beta}({\kappa})
516 \end{split}
517 \label{fourierQuad}
518 \end{equation}
519 If the field gradient is homogeneous over the
520 entire volume, ${\partial_ \alpha E_\beta}({\kappa}) = 0 $ except at
521 $ {\kappa} = 0$, hence it is sufficient to know
522 ${T}_{\alpha\beta\alpha\beta}({\kappa})$ at $ {\kappa} =
523 0$. Therefore equation (\ref{fourierQuad}) can be written as,
524 \begin{equation}
525 \begin{split}
526 \frac{1}{3}{{\Theta}}_{\alpha\beta}({0}) &= \epsilon_o {\chi}^*_Q\; \left(1-\frac{1}{4\pi} {\chi}^*_Q\;{T}_{\alpha\beta\alpha\beta}({0})\right)^{-1} \partial_\alpha E^o_\beta({0})
527 \end{split}
528 \label{fourierQuad2}
529 \end{equation}
530 where $ {T}_{\alpha\beta\alpha\beta}({0})$ can be evaluated as,
531 \begin{equation}
532 {T}_{\alpha\beta\alpha\beta}({0}) = \int {T}_{\alpha\beta\alpha\beta}\;(\textbf{r})d^3r
533 \label{realTensorQaud}
534 \end{equation}
535
536 In terms of traced quadrupole moment equation (\ref{fourierQuad2}) can be written as,
537 \begin{equation}
538 {{Q}}_{\alpha\beta} = \frac{1}{3}\delta_{\alpha\beta}\;Tr({Q}) + \epsilon_o\; {\chi}^*_Q\left(1-\frac{1}{4\pi} {\chi}^*_Q\;{T}_{\alpha\beta\alpha\beta}({0})\right)^{-1}\; \partial_\alpha E^o_\beta
539 \label{tracedConstQuad}
540 \end{equation}
541 Comparing (\ref{tracedConstQuad}) and (\ref{flucQuad}) we get,
542 \begin{equation}
543 \begin{split}
544 &\frac{\braket{{Q^2}} - \braket{Q}^2}{15 \epsilon_o Vk_BT}\; =\; {\chi}^*_Q\;\left(1-\frac{1}{4\pi} {\chi}^*_Q\;{T}_{\alpha\beta\alpha\beta}({0})\right)^{-1}, \\
545 &{\chi}^*_Q \;=\; \frac{\braket{{Q^2}} - \braket{Q}^2}{15 \epsilon_o Vk_BT}\left(1 + \frac{1}{4\pi} \frac{\braket{{Q^2}} - \braket{Q}^2}{15 \epsilon_o Vk_BT}\;{T}_{\alpha\beta\alpha\beta}({0})\right)^{-1}
546 \end{split}
547 \end{equation}
548 Finally the quadrupolar susceptibility cab be written in terms of quadrupolar correction factor ($A_{quad}$) as below,
549 \begin{equation}
550 {\chi}^*_Q \;=\; \frac{\braket{{Q^2}} - \braket{Q}^2}{15 \epsilon_o Vk_BT}\left(1 + \frac{\braket{{Q^2}} - \braket{Q}^2}{15 \epsilon_o Vk_BT}\; A_{quad}\right)^{-1} = \alpha_Q\left(1 + \alpha_Q\; A_{quad}\right)^{-1}
551 \label{eq:quadrupolarSusceptiblity}
552 \end{equation}
553 where $A_{quad} = \frac{1}{4\pi}\int {T}_{\alpha\beta\alpha\beta}\;(\textbf{r})d^3r $ has dimension of the $length^{-2}$ is different for different cutoff methods which is listed in Table \ref{tab:B}. The dielectric constant associated with the quadrupolar susceptibility is defined as,\cite{Ernst92}
554
555 \begin{equation}
556 \epsilon = 1 + \chi^*_Q\; G = 1 + G \; \alpha_Q\left(1 + \alpha_Q\; A_{quad}\right)^{-1}
557 \label{eq:dielectricFromQuadrupoles}
558 \end{equation}
559 where $G = \frac{\displaystyle\int_V |\partial_\alpha E^o_\beta|^2 d^3r}{\displaystyle\int_V{|E^o|}^2 d^3r}$ is a geometrical factor depends on the nature of the external field perturbation. This is true when the quadrupolar fluid is homogeneous over the sample. Since quadrupolar molecule couple with the gradient of the field, the distribution of the quadrupoles is inhomogeneous for varying field gradient. Hence the distribution function should also be taken into account to calculate actual geometrical factor in the presence of non-uniform gradient field. Therefore,
560 \begin{equation}
561 G = \frac{\displaystyle\int_V\; g(r, \theta, \phi)\; |\partial_\alpha E^o_\beta|^2 d^3r}{\displaystyle\int_V{|E^o|}^2 d^3r}
562 \label{eq:geometricalFactor}
563 \end{equation}
564 where $g(r,\theta, \phi)$ is a distribution function of the quadrupoles in with respect to origin at the center of line joining two probe charges.
565 \begin{table}
566 \caption{Expressions for the quadrupolar correction factor ($A$) for the real-space electrostatic methods in terms of the damping parameter
567 ($\alpha$) and the cutoff radius ($r_c$). The dimension of the correction factor is $ length^{-2}$ in case of quadrupolar fluid.}
568 \label{tab:B}
569 \centering
570 \resizebox{\columnwidth}{!}{%
571
572 \begin{tabular}{l|c|c|c|c|}
573
574 Method & $A_\mathrm{charges}$ & $A_\mathrm{dipoles}$ &$A_\mathrm{quadrupoles}$ \\\hline
575 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}$ \\
576 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}$ \\
577 Gradient-shifted (GSF) & $- \frac{8 \alpha^5 {r_c}^3}{3\sqrt{\pi}} e^{-\alpha^2 r_c^2}$ & 0 & $-\frac{4{\alpha}^7{r_c}^5 }{9\sqrt{\pi}}e^{-\alpha^2 r_c^2}(-1+2\alpha ^2 r_c^2)$\\
578 Taylor-shifted (TSF) & $ -\frac{8 \alpha^5 {r_c}^3}{3\sqrt{\pi}} e^{-\alpha^2 r_c^2}$ & $\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) $ & $\frac{10\;\mathrm{erfc}(\alpha r_c )}{{r_c}^2} + \frac{4{\alpha}}{9\sqrt{\pi}{r_c}}e^{-\alpha^2 r_c^2}\left(45 + 30\alpha ^2 {r_c}^2 + 12\alpha^4 {r_c}^4 + 3\alpha^6 {r_c}^6 + 2 \alpha^8 {r_c}^8\right)$ \\\hline
579 \end{tabular}%
580 }
581 \end{table}
582
583 \section{Screening of Charges by Multipolar Fluids}
584 In this method, we will measure the interaction between a positive and negative charge at varying distances after introducing a dipolar (or quadrupolar) material between them. The potential of mean force (PMF) between two ions in a liquid is obtained by constraining their distance and measuring the mean constraint force required to hold them at a fixed distance $r.$ The PMF is obtained from a sequence of simulations can be expressed as \cite{Wilfred07},
585 \begin{equation}
586 w(r) = \int_{r_o}^{r}\braket{\frac{\partial f}{\partial r'}}dr' + 2kT \log(r/r_o) + w(r_o),
587 \label{eq:pmf}
588 \end{equation}
589 where $\braket{\partial f/\partial r'}$ is the mean constraint force, $2kT \log(r/r_o)$ is the Fixman factor, and $r_o$ is the reference position. The potential energy between two charges at separation $r$ is given by,
590 \begin{equation}
591 w(r) = \frac{1}{4\pi\epsilon}\frac{q_1q_2}{r},
592 \end{equation}
593 where $\epsilon$ is the dielectric constant for the fluid.
594
595 The quadrupole molecule can couple with the gradient of the electric field and the two opposite point charges produces non-zero value of both electric field as well as gradient of the field. Therefore, this simulation set up can be used to determine the dielectric constant for both dipolar and quadrupolar fluid.
596 \label{sec:PMF}
597 \section{Simulation Methodology}
598 We have used three different simulation methods: i) external field perturbation, ii) fluctuation formula, and iii) potential of mean force (PMF), to calculate dielectric properties for dipolar and quadrupolar fluid. For the dipolar system we calculated macroscopic polarzability using first two methods and derived the dielectric constant using polarizability and correction factor (see equation \ref{correctionFormula}). Similarly we used equation (\ref{eq:pmf}) to calculate screening factor from dipolar fluid using PMF method. For quadrupolar fluid, we have calculated quadrupolarizablity using fluctuation formula and external field perturbation and derived quadrupolar susceptibility using quadrupolarizability and correction factor (equation \ref{eq:quadrupolarSusceptiblity}). Since dielectric constant due to quadrupolar fluid depends on the nature of gradient of the field applied in the system, we have used geometrical factor (in equation \ref{eq:geometricalFactor}) and quadrupolar susceptibility to evaluate dielectric constant for two ions dissolved quadrupolar fluid (see equation \ref{eq:dielectricFromQuadrupoles}) . The the dielectric constant evaluated using equation (\ref{eq:dielectricFromQuadrupoles}) has been compared with the result evaluated from PMF method (i.e. equation \ref{eq:pmf}). We have also used three different test systems for both dipolar and quadrupolar fluids to calculate dielectric properties. The parameters used in the test systems are given in table \ref{Tab:C}.
599
600 \begin{table}
601 \caption{\label{Tab:C}}
602 \begin{tabularx}{\textwidth}{r|cc|YYccc|Yccc} \hline
603 & \multicolumn{2}{c|}{LJ parameters} &
604 \multicolumn{5}{c|}{Electrostatic moments} & & & & \\
605 Test system & $\sigma$& $\epsilon$ & $C$ & $D$ &
606 $Q_{xx}$ & $Q_{yy}$ & $Q_{zz}$ & mass & $I_{xx}$ & $I_{yy}$ &
607 $I_{zz}$ \\ \cline{6-8}\cline{10-12}
608 & (\AA) & (kcal/mol) & (e) & (debye) & \multicolumn{3}{c|}{(debye \AA)} & (amu) & \multicolumn{3}{c}{(amu
609 \AA\textsuperscript{2})} \\ \hline
610 Stockmayer fluid & 3.41 & 0.2381 & - & 1.4026 &-&-&-& 39.948 & 11.613 & 11.613 & 0.0 \\
611 Quadrupolar fluid & 2.985 & 0.265 & - & - & 0.0 & 0.0 &-2.139 & 18.0153 & 43.0565 & 43.0565 & 0.0 \\
612 \ce{q+} & 1.0 & 0.1 & +1 & - & - & - & - & 22.98 & - & - & - \\
613 \ce{q-} & 1.0 & 0.1 & -1 & - & - & - & - & 22.98 & - & - & - \\ \hline
614 \end{tabularx}
615 \end{table}
616
617 First test system consists of point dipolar or quadrupolar molecules in the presence of constant field or gradient field. Since there is no isolated charge within the system, the divergence of the field should be zero $ i.e. \vec{\nabla} .\vec{E} = 0$. This condition is satisfied by selecting applied potential as described in Appendix \ref{Ap:fieldOrGradient}. When constant electric field or field gradient applied to the system, the molecules align along the direction of the applied field. We have calculated ensemble average of the box dipole or quadrupole moment as a response field or field gradient. Similarly the macroscopic polarizability of the system is derived using ratio between system multipolar moment and applied field or field gradient. This method works properly when the system is at the linear response region of field or field gradient.
618
619 Second test system consists of box of point dipolar or quadrupolar molecules is simulated for 1 ns in NVE ensemble after equilibration in the absence of any external perturbation. The fluctuation of the ensemble average of the box multipolar moment, $\braket{A^2} - \braket{A}^2 $ where A is box dipolar or quadrupolar moment, is measured at the fixed temperature and density for a given multipolar fluid. Finally the macroscopic polarizability of the system at a particular density is derived using equation (\ref{flucQuad}).
620
621 Final system consists of dipolar or quadrupolar fluids with two oppositely charged ions immersed in it. These ions are constraint to be at fixed distance throughout the simulation. We run separate simulations for the different constraint distances. We calculated the screening factor using ratio between the force between the two ions in the absence of medium and the average constraint force during the simulation. Since the constraint force is pretty noisy we run each simulation for long time ($\sim 5$ ns) to reduce simulation error.
622
623 \subsection{Implementation}
624 We have used real-space electrostatic methods implemented in OpenMD \cite{openmd2.3} software to evaluate electrostatic interactions between the molecules. In our simulations we used all three different real-space electrostatic methods: SP, GSF, and TSF developed in the previous paper \cite{PaperI} in the series. The radius of the cutoff sphere is taken to be $12 \r{A}$. Each real space method can be tuned using different values of damping parameter. We have selected ten different values of damping parameter (unit-${\r{A}}^{-1}$); 0.0, 0.05, 0.1, 0.15, 0.175, 0.2, 0.225, 0.25, 0.3, and 0.35 in our simulations for dipolar system. Since quadrupolar interactions are less sensitive to damping parameter as compared to dipole we selected only four values of dampig alpha 0.0, 0.1 0.2, and 0.3 $\AA-1$ in our simulation. The short range interaction in the simulations is incorporated using 6-12 Lennard Jones interactions.
625
626 To derive the box multipolar (dipolar or quadrupolar) moment, we added the component each individual molecule and taken ensemble average of each snapshot in the entire simulation. The first component of the fluctuation of the dipolar moment is derived by using relation $\braket{M^2} = \braket{{M_x}^2 + {M_y}^2 + {M_z}^2}$, where $M_x$, $M_y$, and $M_z $ are x, y and z components of the box dipolar moment. Similarly the first term in the quadrupolar system is derived using relation $ \braket{Q^2} = \braket{3 Q:Q - \mathrm{Tr}Q^2} $, where $ Q $ is the box quadrupole moment, double dot represent the outer product of the quadrupolar matrices, and $TrQ$ is the trace of the box quadrupolar moment. The second component of the fluctuation formula has been derived using square of the ensemble average of the box dipole moment. The applied constant field or field gradient in the test systems has been taken in the form described in the Appendix \ref{Ap:fieldOrGradient}.
627 \subsection{Model systems}
628 To evaluate dielectric properties for dipolar systems using perturbation and fluctuation formula methods, we have taken system of 2048 Stockmayer molecules with reduced density $ \rho^* = 0.822$, temperature $T^* = 1.15 $, moment of inertia $I^* = 0.025 $, and dipole moment $ \mu^* = \sqrt{3.0} $. Test systems are equilibrated for 500 ps and run for $1\; ns$ and components of box dipole moment are obtained at every femtosecond. The systems have been run in the presence of constant external field from $ 0 - 10\; \times\; 10^{-4}\;V/{\r{A}}$ in the step of $ 10 ^{-4}\; V/\r{A}$ for each simulation. For pmf method, Two dipolar molecules in the above system are converted into $q+$ and $q-$ ions and constrained to remain in fixed distance in simulation. The constrained distance is varied from $5\;\r{A} - 12\; \r{A} $ for different simulations. In pmf method all simulations are equilibrated for 500 ps in NVT ensemble and run for 5 ns in NVE ensemble to print constraint force at an interval of 20 fs.
629
630 Quadrupolar systems consists 4000 linear point quadrupoles with density $ 2.338\; g/cm^3$ at temperature $ 500\; ^oK $. For both perturbation and fluctuation methods, test systems are equalibrated for 200 ps in NVT ensemble and run for 500 ps in NVE ensemble. To find the ensemble average of the box quadrupole moment and fluctuation, the components of box quadrupole moments are printed every 100 fs. Each simulations repeated at different values of applied gradients from $ 0 - 9 \times 10^{-2}\; V/\AA^2 $. To calculate dielectric constant using pmf method, two ions in the systems are converted into $q+$ and $q-$ ions and constrained to remain at fixed distance in the simulation. These constraint distances are varied from $5\;\AA - 12\; \AA $ at the step of $0.1\; \AA $ for different simulations. For calculating dielectric constant, the test systems are run for 500 ps to equlibrate and run for 5 ns to print constraint force at a time interval of 20 fs.
631
632 \section{Results}
633 \subsection{Dipolar fluid}
634 The macroscopic polarizability ($\alpha_D$) for Stockmayer fluid evaluated using perturbation (left) and fluctuation (right) methods as shown in figure \ref{fig:dielectricDipole} . The polarizability obtained from the both perturbation and fluctuation methods show excellent agreement with each other. The result shows that polarizability hugely depends on the values of the damping parameter in the case of Shifted Potential (SP) and Gradient Shifted force (GSF) methods. But the results obtained using Taylor Shifted Force (TSF) show slight dependence on the damping alpha. To get correct dielectric properties from the polarizability we used a correction factor as we discussed in section \ref{sec:corrFactor}. The correction factor for dipolar fluids ($\mathrm{A}_\mathrm{dipole}$ ) for the different values of damping parameter is also plotted in figure \ref{fig:dielectricDipole}. For TSF method, $ \mathrm{A}_\mathrm{dipole} = 1 $ for all the values of damping alpha. Therefore we do not need to perform any correction to obtain dielectric constant from the polarizability. The values of $\mathrm{A}_\mathrm{dipole} $ varies with the damping parameter in the case of SP and GSF methods therefore correction factor should be incorporated to calculate the dielectric constant ($\epsilon$) from the polarzability. The dielectric constant obtained from the polarizability and correction factor is also plotted in the figure \ref{fig:dielectricDipole}. The dielectric constant evaluated using SP and GSF method shows good agreement with the previous simulation result (straight dotted line in the lower panels)\cite{NeumannI83} for the damping parameter from $0.25\; \mathrm{to}\; 0.3$ $\AA-1$. For lower value of the damping alpha the dielectric constant is much deviated away from the previous simulation result. To investigate furthermore we performed inverse transformation of expected dielectric constant \cite{NeumannI83}, using correction factor and calculated expected polarizability for the simulation, which is represented by the dotted line in the upper-left and upper-right panels. In the plot we can clearly see that expected ploarizability for the SP and GSF methods are very close to results obtained from the simulations. Sometime even they are within the range of the error-bar but when the dielectric constant is calculated using simulation result and correction formula, it is found to be deviated much away from the expected result for the lower values of the damping alpha. This tell us that the correction formula (equation \ref{correctionFormula}) for the dipolar fluid is very sensitive when the value of $\mathrm{A}_\mathrm{dipole}$ is deviated much away from 1 for the small values of damping parameter. Therefore it is always better to select damping alpha from 0.25 to 0.3 $\AA^{-1}$ to evaluate dielectric constant using perturbation and fluctuation formula while using SP and GSF methods in simulation.
635
636 \begin{figure}
637 \includegraphics[width=3in]{dielectricFinal_Dipole.pdf}
638 \caption{In the figure, $\alpha_D$, $\mathrm{A}_\mathrm{dipole}$, and $\epsilon$ are polarizability, correction factor, and dielectric constant for Stockmayer fluid. Plots in the left panel show results for (a) perturbation method and right (b) fluctuation method.}
639 \label{fig:dielectricDipole}
640 \end{figure}
641
642 We have also evaluated the screening factor between two oppositely charged ions when they are placed in the Stockmayer fluid as shown in figure \ref{fig:ScreeningFactor_Dipole}. The screening factor have been evaluated using equation \ref{eq:pmf}. This screening factor is analogous to dielectric screening but there is subtle differences between them. Actually screening factor measures a local property of the ions in the fluid and depends on ion-dipole and dipole-dipole interactions. These interactions depends on the distance between ions and electrostatic interaction methods implemented in the simulations. But we can expect screening to be close to dielectric constant when the field due to ions is very gentle in perturbation. It is possible when distance between ions are far apart and values of damping parameter relatively higher. In the figure \ref{fig:ScreeningFactor_Dipole} we can observe that for the higher value of damping alpha i.e $\alpha = 0.2 \AA^{-1}$ and $0.3 \AA^{-1}$ and large separation between ions, the screening factor approaches to the dielectric constant. We can also observe that for TSF method has got higher screening factor as compared SP and GSF method even for lower damping alpha. It is because of dipole-dipole interactions do not need any correction factor in the case of TSF as we discussed earlier. But presence of ion-dipole interaction causing local effect even in the case of TSF method for small ions separation.
643
644 \begin{figure}
645 \includegraphics[width=3in]{ScreeningFactor_Dipole.pdf}
646 \caption{Figure shows screening factor between two oppositely charged ions immersed in Stockmayer fluid as a function of ions-separation for the different values damping alpha and electrostatic methods.}
647 \label{fig:ScreeningFactor_Dipole}
648 \end{figure}
649
650
651 \subsection{Quadrupolar fluid}
652 The polarizability ($\alpha_Q$), correction factor ($\mathrm{A_{quad}}$), and susceptibility ($\chi$) for the quadrupolar fluid is plotted against damping parameter in the figure \ref{fig:dielectricQuad}. Different than the dipolar fluid, the polarizability and susceptibility for the quadrupolar fluid have a dimension of $\AA ^2$. Although susceptibility has got dimension, it is a measure of macroscopic quadrupolar properties at a given temperature and density \cite{JeonI03, JeonII03}. The left panel in the plot showing results obtained from the perturbation whereas the results from the fluctuation formula are plotted in the right panel. In the figure, we can observe that the polarizablity evaluated from the perturbation (top-left) shows excellent agreement with the result obtained from the fluctuation formula (top-right). The susceptibility for the quadrupolar fluid is obtained from quadrupolarizability and correction factor using equation \ref{eq:quadrupolarSusceptiblity} which is plotted at the bottom-left and bottom-right in the figure \ref{fig:dielectricQuad}. All three methods: SP, GSF, and TSF tends to produce same value of susceptibility for the given value damping alpha. This shows that susceptibility derived using correction factor and polarizablity is tends to be independent of the electrostatic method implemented in the simulation for quadrupolar fluid.
653 \begin{figure}
654 \includegraphics[width=3in]{polarizabilityFinal_Quad.pdf}
655 \caption{In the figure, $\alpha_Q$, $\mathrm{A}_\mathrm{quad}$, and $\chi$ are polarizability, correction factor, and susceptibility for the quadrupolar fluid. Plots in the left panel show results for (a) perturbation method and right (b) fluctuation method.}
656 \label{fig:dielectricQuad}
657 \end{figure}
658
659 The actual test of the susceptibility can be made by comparing the results obtained from the perturbation and fluctuation with the result calculated from more direct, potential of mean force (PMF), method. For this purpose, we have calculated the expected dielectric constant from the susceptibility and geometrical factor using equation \ref{eq:dielectricFromQuadrupoles}. Since the dielectric constant for quadrupolar fluid is geometry-dependent, it is not a actual measure of the property of the quadrupolar fluid. This geometrical factor varies with the variation in ions separation as well as damping parameter therefore the dielectric constant can be refereed as screening factor. The screening factor plotted against ions separation, evaluated from the perturbation and fluctuation method are shown in right and central panel of the figure \ref{fig:screenQuad}. Here the susceptibility is calculated from the simulation and the geometrical factor is evaluated analytically, using field and field-gradient produced by ions. The right hand panel shows screening factor plotted against ions separation, for the different values of the damping alpha, obtained from the PMF method. The figure clearly shows that the screening factor obtained from the perturbation and fluctuation formula shows good agreement with the screening factor calculated using PMF method. Since there is no huge differences in quadrupole-quadruople interactions for various real-space methods, we do not observe large differences in the screen factors for SP, GSF, and TSF methods.
660
661 \begin{figure}
662 \includegraphics[width=3in]{ScreeningFactor_Quad.pdf}
663 \caption{}
664 \label{fig:screenQuad}
665 \end{figure}
666
667 \section{Discussion}
668 We have used the perturbation and fluctuation formula to evaluate dielectric properties for dipolar and quadrupolar fluids implementing SP, GSF, and TSF methods in the simulation. Similarly the correction factors for the both dipolar and quadrupolar systems have also been calculated for all SP, GSF, and TSF methods. We have also derived screening factor between two ions immersed in the dipolar and quadrupolar fluids using PMF method. The result from the perturbation and fluctuation formula are compared with PMF method to test the accuracy of the simulation result.
669
670 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 $0.3 \AA^{-1}$.
671
672 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.
673
674
675 \appendix
676 \section{Point-multipolar interactions with a spatially-varying electric field}
677
678 We can treat objects $a$, $b$, and $c$ containing embedded collections
679 of charges. When we define the primitive moments, we sum over that
680 collections of charges using a local coordinate system within each
681 object. The point charge, dipole, and quadrupole for object $a$ are
682 given by $C_a$, $\mathbf{D}_a$, and $\mathsf{Q}_a$, respectively.
683 These are the primitive multipoles which can be expressed as a
684 distribution of charges,
685 \begin{align}
686 C_a =&\sum_{k \, \text{in }a} q_k , \label{eq:charge} \\
687 D_{a\alpha} =&\sum_{k \, \text{in }a} q_k r_{k\alpha}, \label{eq:dipole}\\
688 Q_{a\alpha\beta} =& \frac{1}{2} \sum_{k \, \text{in } a} q_k
689 r_{k\alpha} r_{k\beta} . \label{eq:quadrupole}
690 \end{align}
691 Note that the definition of the primitive quadrupole here differs from
692 the standard traceless form, and contains an additional Taylor-series
693 based factor of $1/2$. In Paper 1, we derived the forces and torques
694 each object exerts on the others.
695
696 Here we must also consider an external electric field that varies in
697 space: $\mathbf E(\mathbf r)$. Each of the local charges $q_k$ in
698 object $a$ will then experience a slightly different field. This
699 electric field can be expanded in a Taylor series around the local
700 origin of each object. A different Taylor series expansion is carried
701 out for each object.
702
703 For a particular charge $q_k$, the electric field at that site's
704 position is given by:
705 \begin{equation}
706 E_\gamma + \nabla_\delta E_\gamma r_{k \delta}
707 + \frac {1}{2} \nabla_\delta \nabla_\varepsilon E_\gamma r_{k \delta}
708 r_{k \varepsilon} + ...
709 \end{equation}
710 Note that the electric field is always evaluated at the origin of the
711 objects, and treating each object using point multipoles simplifies
712 this greatly.
713
714 To find the force exerted on object $a$ by the electric field, one
715 takes the electric field expression, and multiplies it by $q_k$, and
716 then sum over all charges in $a$:
717
718 \begin{align}
719 F_\gamma &= \sum_{k \textrm{~in~} a} q_k \lbrace E_\gamma + \nabla_\delta E_\gamma r_{k \delta}
720 + \frac {1}{2} \nabla_\delta \nabla_\varepsilon E_\gamma r_{k \delta}
721 r_{k \varepsilon} + ... \rbrace \\
722 &= C_a E_\gamma + D_{a \delta} \nabla_\delta E_\gamma
723 + Q_{a \delta \varepsilon} \nabla_\delta \nabla_\varepsilon E_\gamma +
724 ...
725 \end{align}
726
727 Similarly, the torque exerted by the field on $a$ can be expressed as
728 \begin{align}
729 \tau_\alpha &= \sum_{k \textrm{~in~} a} (\mathbf r_k \times q_k \mathbf E)_\alpha \\
730 & = \sum_{k \textrm{~in~} a} \epsilon_{\alpha \beta \gamma} q_k
731 r_{k\beta} E_\gamma(\mathbf r_k) \\
732 & = \epsilon_{\alpha \beta \gamma} D_\beta E_\gamma
733 + 2 \epsilon_{\alpha \beta \gamma} Q_{\beta \delta} \nabla_\delta
734 E_\gamma + ...
735 \end{align}
736
737 The last term is essentially identical with form derived by Torres del
738 Castillo and M\'{e}ndez Garrido,\cite{Torres-del-Castillo:2006uo} although their derivation
739 utilized a traceless form of the quadrupole that is different than the
740 primitive definition in use here. We note that the Levi-Civita symbol
741 can be eliminated by utilizing the matrix cross product in an
742 identical form as in Ref. \onlinecite{Smith98}:
743 \begin{equation}
744 \left[\mathsf{A} \times \mathsf{B}\right]_\alpha = \sum_\beta
745 \left[\mathsf{A}_{\alpha+1,\beta} \mathsf{B}_{\alpha+2,\beta}
746 -\mathsf{A}_{\alpha+2,\beta} \mathsf{B}_{\alpha+1,\beta}
747 \right]
748 \label{eq:matrixCross}
749 \end{equation}
750 where $\alpha+1$ and $\alpha+2$ are regarded as cyclic permuations of
751 the matrix indices. In table \ref{tab:UFT} we give compact
752 expressions for how the multipole sites interact with an external
753 field that has exhibits spatial variations.
754
755 \begin{table}
756 \caption{Potential energy $(U)$, force $(\mathbf{F})$, and torque
757 $(\mathbf{\tau})$ expressions for a multipolar site embedded in an
758 electric field with spatial variations, $\mathbf{E}(\mathbf{r})$.
759 \label{tab:UFT}}
760 \begin{tabular}{r|ccc}
761 & Charge & Dipole & Quadrupole \\ \hline
762 $U$ & $C \phi(\mathbf{r})$ & $-\mathbf{D} \cdot \mathbf{E}(\mathbf{r})$ & $- \mathsf{Q}:\nabla \mathbf{E}(\mathbf{r})$ \\
763 $\mathbf{F}$ & $C \mathbf{E}(\mathbf{r})$ & $+\mathbf{D} \cdot \nabla \mathbf{E}(\mathbf{r})$ & $+\mathsf{Q} : \nabla\nabla\mathbf{E}(\mathbf{r})$ \\
764 $\mathbf{\tau}$ & & $\mathbf{D} \times \mathbf{E}(\mathbf{r})$ & $+2 \mathsf{Q} \times \nabla \mathbf{E}(\mathbf{r})$
765 \end{tabular}
766 \end{table}
767 \section{Gradient of the field due to quadrupolar polarization}
768 \label{singularQuad}
769 In this section, we will discuss the gradient of the field produced by
770 quadrupolar polarization. For this purpose, we consider a distribution
771 of charge ${\rho}(r)$ which gives rise to an electric field
772 $\vec{E}(r)$ and gradient of the field $\vec{\nabla} \vec{E}(r)$
773 throughout space. The total gradient of the electric field over volume
774 due to the all charges within the sphere of radius $R$ is given by
775 (cf. Jackson equation 4.14):
776 \begin{equation}
777 \int_{r<R} \vec{\nabla}\vec{E}\;d^3r = -\int_{r=R} R^2 \vec{E}\;\hat{n}\; d\Omega
778 \label{eq:8}
779 \end{equation}
780 where $d\Omega$ is the solid angle and $\hat{n}$ is the normal vector
781 of the surface of the sphere which is equal to
782 $sin[\theta]cos[\phi]\hat{x} + sin[\theta]sin[\phi]\hat{y} +
783 cos[\theta]\hat{z}$
784 in spherical coordinates. For the charge density ${\rho}(r')$, the
785 total gradient of the electric field can be written as (cf. Jackson
786 equation 4.16),
787 \begin{equation}
788 \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
789 \label{eq:9}
790 \end{equation}
791 The radial function in the equation (\ref{eq:9}) can be expressed in
792 terms of spherical harmonics as (cf. Jackson equation 3.70),
793 \begin{equation}
794 \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)
795 \label{eq:10}
796 \end{equation}
797 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,
798 \begin{equation}
799 \begin{split}
800 \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 \\
801 &= -\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
802 '
803 \end{split}
804 \label{eq:11}
805 \end{equation}
806 The gradient of the product of radial function and spherical harmonics
807 is given by (cf. Arfken, p.811 eq. 16.94):
808 \begin{equation}
809 \begin{split}
810 \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
811 {\partial}{\partial r}+\frac{l}{r} \right]f(r)\; Y_{l, l-1, m}(\theta, \phi).
812 \end{split}
813 \label{eq:12}
814 \end{equation}
815 Using equation (\ref{eq:12}) we get,
816 \begin{equation}
817 \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}},
818 \label{eq:13}
819 \end{equation}
820 where $ Y_{l,l+1,m}(\theta, \phi)$ is the vector spherical harmonics
821 which can be expressed in terms of spherical harmonics as shown in
822 below (cf. Arfkan p.811),
823 \begin{equation}
824 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},
825 \label{eq:14}
826 \end{equation}
827 where $C(l+1,1,l|m_1,m_2,m)$ is a Clebsch-Gordan coefficient and
828 $\hat{e}_{m_2}$ is a spherical tensor of rank 1 which can be expressed
829 in terms of Cartesian coordinates,
830 \begin{equation}
831 {\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}}
832 \label{eq:15}
833 \end{equation}
834 The normal vector $\hat{n} $ can be expressed in terms of spherical tensor of rank 1 as shown in below,
835 \begin{equation}
836 \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)
837 \label{eq:16}
838 \end{equation}
839 The surface integral of the product of $\hat{n}$ and
840 ${Y_{l+1}}^{m_1}(\theta, \phi)$ gives,
841 \begin{equation}
842 \begin{split}
843 \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 \\
844 &= \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 \\
845 &= \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),
846 \end{split}
847 \label{eq:17}
848 \end{equation}
849 where ${Y_{l}}^{-m} = (-1)^m\;{{Y_{l}}^{m}}^* $ and
850 $ \int {{Y_{l}}^{m}}^*\;{Y_{l'}}^{m'}\;d\Omega =
851 \delta_{ll'}\delta_{mm'} $.
852 Non-vanishing values of equation \ref{eq:17} require $l = 0$,
853 therefore the value of $ m = 0 $. Since the values of $ m_1$ are -1,
854 1, and 0 then $m_2$ takes the values 1, -1, and 0, respectively
855 provided that $m = m_1 + m_2$. Equation \ref{eq:11} can therefore be
856 modified,
857 \begin{equation}
858 \begin{split}
859 \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(
860 1, 1, 0|0,0,0)\;{\hat{e}_{0}}{\hat{e}_{0}} ]\; d^3r'.
861 \end{split}
862 \label{eq:18}
863 \end{equation}
864 After substituting ${Y^*}_{00} = \frac{1}{\sqrt{4\pi}} $ and using the
865 values of the Clebsch-Gorden coefficients: $ C(1, 1, 0|-1,1,0) =
866 \frac{1}{\sqrt{3}}, \; C(1, 1, 0|-1,1,0)= \frac{1}{\sqrt{3}}$ and $
867 C(1, 1, 0|0,0,0) = -\frac{1}{\sqrt{3}}$ in equation \ref{eq:18} we
868 obtain,
869 \begin{equation}
870 \begin{split}
871 \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)\\
872 &= - \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).
873 \end{split}
874 \label{eq:19}
875 \end{equation}
876 Equation (\ref{eq:19}) gives the total gradient of the field over a
877 sphere due to the distribution of the charges. For quadrupolar fluids
878 the total charge within a sphere is zero, therefore
879 $ \int_{r<R} \vec{\nabla}\vec{E}\;d^3r = 0 $. Hence the quadrupolar
880 polarization produces zero net gradient of the field inside the
881 sphere.
882
883 \section{Applied field or field gradient}
884 \label{Ap:fieldOrGradient}
885
886 To satisfy the condition $ \nabla . E = 0 $, within the box of molecules we have taken electrostatic potential in the following form
887 \begin{equation}
888 \begin{split}
889 \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. \\
890 & \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),
891 \end{split}
892 \label{eq:appliedPotential}
893 \end{equation}
894 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,
895 \[\bf{E}
896 =\frac{g_o}{2} \left(\begin{array}{ccc}
897 2(a_1\; b_1 - \frac{cos\psi}{3})\;x \;+ (a_1\; b_2 \;+ a_2\; b_1)\;y + (a_1\; b_3 \;+ a_3\; b_1)\;z \\
898 (a_2\; b_1 \;+ a_1\; b_2)\;x + 2(a_2\; b_2 \;- \frac{cos\psi}{3})\;y + (a_2\; b_3 \;+ a_3\; b_3)\;z \\
899 (a_3\; b_1 \;+ a_3\; b_2)\;x + (a_3\; b_2 \;+ a_2\; b_3)y + 2(a_3\; b_3 \;- \frac{cos\psi}{3})\;z
900 \end{array} \right).\]
901 The gradient of the applied field derived from the potential can be written in the following form,
902 \[\nabla\bf{E}
903 = \frac{g_o}{2}\left(\begin{array}{ccc}
904 2(a_1\; b_1 - \frac{cos\psi}{3}) & (a_1\; b_2 \;+ a_2\; b_1) & (a_1\; b_3 \;+ a_3\; b_1)\;z \\
905 (a_2\; b_1 \;+ a_1\; b_2) & 2(a_2\; b_2 \;- \frac{cos\psi}{3}) & (a_2\; b_3 \;+ a_3\; b_3)\;z \\
906 (a_3\; b_1 \;+ a_3\; b_2) & (a_3\; b_2 \;+ a_2\; b_3) & 2(a_3\; b_3 \;- \frac{cos\psi}{3})\;z
907 \end{array} \right).\]
908 \newpage
909
910 \bibliography{dielectric_new}
911
912 \end{document}