ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/multipole/dielectric_new.tex
Revision: 4393
Committed: Thu Dec 3 17:29:17 2015 UTC (8 years, 6 months ago) by mlamichh
Content type: application/x-tex
File size: 62569 byte(s)
Log Message:
added result for dipole and quadrupole

File Contents

# Content
1 \documentclass[%
2 aip,
3 jcp,
4 amsmath,amssymb,
5 preprint,%
6 % reprint,%
7 %author-year,%
8 %author-numerical,%
9 ]{revtex4-1}
10
11 \usepackage{graphicx}% Include figure files
12 \usepackage{dcolumn}% Align table columns on decimal point
13 \usepackage{multirow}
14 \usepackage{bm}% bold math
15 \usepackage{natbib}
16 \usepackage{times}
17 \usepackage{mathptmx}
18 \usepackage[version=3]{mhchem} % this is a great package for formatting chemical reactions
19 \usepackage{url}
20 \usepackage{braket}
21 \usepackage{tabularx}
22 \newcolumntype{Y}{>{\centering\arraybackslash}X}
23
24 \begin{document}
25
26 \title{Real space electrostatics for multipoles. III. Dielectric Properties}
27
28 \author{Madan Lamichhane}
29 \affiliation{Department of Physics, University
30 of Notre Dame, Notre Dame, IN 46556}
31 \author{Thomas Parsons}
32 \affiliation{Department of Chemistry and Biochemistry, University
33 of Notre Dame, Notre Dame, IN 46556}
34 \author{Kathie E. Newman}
35 \affiliation{Department of Physics, University
36 of Notre Dame, Notre Dame, IN 46556}
37 \author{J. Daniel Gezelter}
38 \email{gezelter@nd.edu.}
39 \affiliation{Department of Chemistry and Biochemistry, University
40 of Notre Dame, Notre Dame, IN 46556}
41
42 \date{\today}% It is always \today, today,
43 % but any date may be explicitly specified
44
45 \begin{abstract}
46 Note: This manuscript is a work in progress.
47
48 We report on the dielectric properties of the shifted potential
49 (SP), gradient shifted force (GSF), and Taylor shifted force (TSF)
50 real-space methods for multipole interactions that were developed in
51 the first two papers in this series. We find that some subtlety is
52 required for computing dielectric properties with the real-space
53 methods, particularly when using the common fluctuation formulae.
54 Three distinct methods for computing the dielectric constant are
55 investigated, including the standard fluctuation formulae,
56 potentials of mean force between solvated ions, and direct
57 measurement of linear solvent polarization in response to applied
58 fields and field gradients.
59 \end{abstract}
60
61 \maketitle
62
63 \section{Introduction}
64
65 Over the past several years, there has been increasing interest in
66 pairwise methods for computing electrostatic interactions in
67 simulations of condensed molecular
68 systems.\cite{Wolf99,Zahn02,Kast03,Beckd.A.C._Bi0486381,Ma05,Fennell06}
69 These techniques were initially developed from the observations and efforts of
70 Wolf {\it et al.} and their work towards an $\mathcal{O}(N)$
71 Coulombic sum.\cite{Wolf99} Wolf's method of cutoff neutralization is
72 able to obtain excellent agreement with Madelung energies in ionic
73 crystals.\cite{Wolf99} Later, Zahn \textit{et al.} and Fennell and Gezelter extended this method which incorporates Wolf's electrostatic energy and modified it to conserve the total energy in molecular dynamic simulation.\cite{Zahn02, Fennell06} In the previous two papers
74 we developed three new generalized real space methods: Shifted potential (SP), Gradeint shifted force (GSF), and Taylor shifted force (TSF).\cite{PaperI, PaperII} These methods evaluate electrostatic interactions for higher order multipoles (dipoles and quadrupoles) using finite cutoff sphere with the neutralization of the electrostatic moment within the cutoff sphere. Furthermore, extra terms added to the potential energy so that force and torque vanish smoothly at the cutoff radius. This ensures that the total energy is conserved in a molecular dynamic simulation.
75
76 Zahn \textit{et al.} and Fennell and Gezelter extended this method
77 using shifted force approximations at the cutoff distance in order to
78 conserve total energy in molecular dynamics simulations.\cite{Zahn02,
79 Fennell06} In the previous two papers in this series we developed
80 three generalized real space methods: shifted potential (SP), gradient
81 shifted force (GSF), and Taylor shifted force (TSF).\cite{PaperI,
82 PaperII} These methods provide real-space electrostatic interactions
83 for higher order multipoles (e.g. dipoles and quadrupoles) using a
84 finite cutoff sphere with neutralization of the electrostatic moments
85 within the cutoff region. The multipolar generalizations of the
86 shifted force approach provide additional terms in the potential
87 energy so that force and torque vanish smoothly at the cutoff
88 radius. This ensures that the total energy is conserved in molecular
89 dynamics simulations.
90
91 One of the most difficult tests of any new electrostatic method is the
92 fidelity with which that method can reproduce the bulk-phase
93 polarizability or equivalently, the dielectric properties of a
94 fluid. Since dielectric properties are macroscopic properties, all
95 interactions between molecules in the system contribute. Before the
96 advent of computer simulations, Kirkwood and Onsager developed
97 fluctuation formulae for the dielectric properties of dipolar
98 fluids.\cite{Kirkwood39,Onsagar36} Similar developments were made by
99 Logan \textit{et al.} for the bulk polarizability of quadrupolar
100 fluids.\cite{LoganI81,LoganII82,LoganIII82}
101
102 % While calculating dielectric properties, the formula should be
103 % modified in such a way so that it can accommodate behaviour of
104 % electrostatic neutrality and smoothness of energy, force and torque
105 % at the cutoff radius. Previously many studies have been conducted to
106 % calculate dipolar and quadrupolar dielectric properties using
107 % computer simulations. \cite{Kirkwood39, Onsagar36,LoganI81,
108 % LoganII82, LoganIII82}
109
110 In modern simulations, bulk materials are usually treated using
111 periodic replicas of small regions, and this level of approximation
112 requires corrections to the fluctuation formulae that were derived for
113 the bulk fluids. In 1983 Neumann proposed a general formula for
114 evaluating dielectric properties of dipolar fluids using real-space
115 cutoff methods.\cite{Neumann83} Steinhauser and Neumann used this
116 formula to evaluate the corrected dielectric constant for the
117 Stockmayer fluid using two different methods: Ewald-Kornfield (EK) and
118 reaction field (RF) methods.\cite{Neumann-Steinhauser83}
119
120 % But these methods do not specifically take account of the cutoff
121 % behavior common in real-space electrosatic methods. In 1983 Neumann
122 % proposed a general formula for evaluating dielectric properties for
123 % dipolar fluid using real-space cutoff methods. \cite{Neumann83} In the
124 % same year Steinhauser and Neumann used this formula to evaluate the
125 % correct dielectric constant for the Stockmayer fluid using two
126 % different methods: Ewald-Kornfield (EK) and reaction field (RF)
127 % methods. \cite{Neumann-Steinhauser83} This formula contains a
128 % correction factor which is equal to $\frac{3}{4 \pi} $ times volume
129 % integral of the dipole-dipole interactions for a given electrostatic
130 % cutoff method (See equation
131 % \ref{dipole-diopleTensor}).\cite{Neumann83}
132
133 Zahn \textit{et al.}\cite{Zahn02} utilized this approach and evaluated
134 the correction factor for using damped shifted charge-charge
135 kernel. This was later generalized by Izvekov \textit{et
136 al.},\cite{Izvekov:2008wo} who showed that the expression for the
137 dielectric constant reduces to widely-used \textit{conducting
138 boundary} formula for real-space cutoff methods that have first
139 derivatives that vanish at the cutoff sphere.
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 dielectric constant depends on the
144 geometry of the external field perturbation.\cite{Ernst92} Many
145 studies have also been conducted to understand solvation theory using
146 dielectric properties of these
147 fluids,\cite{JeonI03,JeonII03,Chitanvis96} although a correction
148 formula for different cutoff 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 charge-charge, dipole-dipole or
154 quadrupole-quadrupole interactions.
155
156 We have also calculated the geometrical factor for two ions immersed
157 quadrupolar system to evaluate dielectric constant from the
158 quadrupolar susceptibility. We have used three different methods to
159 compare our results with computer simulations:
160 \begin{enumerate}
161 \item external field and field gradient perturbations,
162 \item fluctuation formulae, and
163 \item the potential of mean force,
164 \end{enumerate}
165 to study dielectric properties of the dipolar and quadrupolar
166 systems.
167
168 In the external field perturbation, the net polarization of the system
169 is observed as a linear response of the applied field perturbation,
170 where proportionality constant is determined by the electrostatic
171 interaction between the electrostatic multipoles at a given
172 temperature. The fluctuation formula observes the time average
173 fluctuation of the multipolar moment as a function of temperature. The
174 average fluctuation value of the system is determined by the
175 multipole-multipole interactions between molecules at a given
176 temperature. Since the expression of the electrostatic interaction
177 energy, force, and torque in the real space electrostatic methods are
178 different from their original definition, both fluctuation and
179 external field perturbation formula should also be modified
180 accordingly. The potential of mean force method calculates dielectric
181 constant from the potential energy between ions before and after
182 dielectric material is introduced. All of these different methods for
183 calculating dielectric properties will be discussed in detail in the
184 following sections: \ref{subsec:perturbation},
185 \ref{subsec:fluctuation}, and \ref{sec:PMF}.
186
187 \section{Boltzmann average for orientational polarization}
188 The dielectric properties of the system is mainly arise from two different ways: i) the applied field distort the charge distributions so it produces an induced multipolar moment in each molecule; and ii) the applied field tends to line up originally randomly oriented molecular moment towards the direction of the applied field. In this study, we basically focus on the orientational contribution in the dielectric properties. If we consider a system of molecules in the presence of external field perturbation, the perturbation experienced by any molecule will not be only due to external field or field gradient but also due to the field or field gradient produced by the all other molecules in the system. In the following subsections \ref{subsec:boltzAverage-Dipole} and \ref{subsec:boltzAverage-Quad}, we will discuss about the molecular polarization only due to external field perturbation. The contribution of the field or field gradient due to all other molecules will be taken into account while calculating correction factor in the section \ref{sec:corrFactor}.
189
190 \subsection{Dipole}
191 \label{subsec:boltzAverage-Dipole}
192 Consider a system of molecules with permenent dipole moment $p_o$. In the absense of external field, thermal agitation makes dipole randomly oriented therefore there is no net dipole moment. But external field tends them to line up in the direction of applied field. Here we have considered net field acting due to all other molecules is considered to be zero. Therefore the total Hamiltonian of the molecule is,\cite{Jackson98}
193
194 \begin{equation}
195 H = H_o - \bf{p_o} .\bf{E},
196 \end{equation}
197 where $H_o$ is a function of the internal coordinates of the molecule. Now Boltzmann average of the dipole moment is given by,
198 \begin{equation}
199 \braket{p_{mol}} = \frac{\displaystyle\int d\Omega\; p_o\; cos\theta\; e^{\frac{p_oE\; cos\theta}{k_B T}}}{\displaystyle\int d\Omega\; e^{\frac{p_oE\;cos\theta}{k_B T}}},
200 \end{equation}
201 where $\bf{E}$ is selected along z-axis. If we consider applied field is small i.e. $\frac{p_oE\; cos\theta}{k_B T} << 1$ then we get,
202
203 \begin{equation}
204 \braket{p_{mol}} \approx \frac{1}{3}\frac{{p_o}^2}{k_B T}E,
205 \end{equation}
206 where $ \alpha_p = \frac{1}{3}\frac{{p_o}^2}{k_B T}$ is a molecular polarizability. The orientational polarization depends inversely on the temperature and applied field must overcome the thermal agitation.
207
208
209 \subsection{Quadrupole}
210 \label{subsec:boltzAverage-Quad}
211 Consider a system of molecules with permanent quadrupole moment $q_{\alpha\beta} $. The average quadrupole moment at temperature T in the presence of uniform applied field gradient is given by,\cite{AduGyamfi78, AduGyamfi81}
212 \begin{equation}
213 \braket{q_{\alpha\beta}} \;=\; \frac{\displaystyle\int d\Omega\; e^{-\frac{H}{k_B T}}q_{\alpha\beta}}{\displaystyle\int d\Omega\; e^{-\frac{H}{k_B T}}} \;=\; \frac{\displaystyle\int d\Omega\; e^{\frac{q_{\mu\nu}\;\partial_\nu E_\mu}{k_B T}}q_{\alpha\beta}}{\displaystyle\int d\Omega\; e^{\frac{q_{\mu\nu}\;\partial_\nu E_\mu}{k_B T}}},
214 \label{boltzQuad}
215 \end{equation}
216 where $\int d\Omega = \int_0^{2\pi} \int_0^\pi \int_0^{2\pi}
217 sin\theta\; d\theta\ d\phi\ d\psi$ is the integration over Euler
218 angles, $ H = H_o -q_{\mu\nu}\;\partial_\nu E_\mu $ is the energy of
219 a quadrupole in the gradient of the
220 applied field and $ H_o$ is a function of internal coordinates of the molecule. The energy and quadrupole moment can be transformed into body frame using following relation,
221 \begin{equation}
222 \begin{split}
223 &q_{\alpha\beta} = \eta_{\alpha\alpha'}\;\eta_{\beta\beta'}\;{q}^* _{\alpha'\beta'} \\
224 &H = H_o - q:\vec{\nabla}\vec{E} = H_o - q_{\mu\nu}\;\partial_\nu E_\mu = H_o -\eta_{\mu\mu'}\;\eta_{\nu\nu'}\;{q}^*_{\mu'\nu'}\;\partial_\nu E_\mu.
225 \end{split}
226 \label{energyQuad}
227 \end{equation}
228 Here the starred tensors are the components in the body fixed
229 frame. Substituting equation (\ref{energyQuad}) in the equation (\ref{boltzQuad})
230 and taking linear terms in the expansion we get,
231 \begin{equation}
232 \braket{q_{\alpha\beta}} = \frac{ \int d\Omega \left(1 + \frac{\eta_{\mu\mu'}\;\eta_{\nu\nu'}\;{q}^*_{\mu'\nu'}\;\partial_\nu E_\mu }{k_B T}\right)q_{\alpha\beta}}{ \int d\Omega \left(1 + \frac{\eta_{\mu\mu'}\;\eta_{\nu\nu'}\;{q}^*_{\mu'\nu'}\;\partial_\nu E_\mu }{k_B T}\right)},
233 \end{equation}
234 where $\eta_{\alpha\alpha'}$ is the inverse of the rotation matrix that transforms
235 the body fixed co-ordinates to the space co-ordinates,
236 \[\eta_{\alpha\alpha'}
237 = \left(\begin{array}{ccc}
238 cos\phi\; cos\psi - cos\theta\; sin\phi\; sin\psi & -cos\theta\; cos\psi\; sin\phi - cos\phi\; sin\psi & sin\theta\; sin\phi \\
239 cos\psi\; sin\phi + cos\theta\; cos\phi \; sin\psi & cos\theta\; cos\phi\; cos\psi - sin\phi\; sin\psi & -cos\phi\; sin\theta \\
240 sin\theta\; sin\psi & -cos\psi\; sin\theta & cos\theta
241 \end{array} \right).\]
242 Integration of 1st and 2nd terms in the denominator gives $8 \pi^2$
243 and $8 \pi^2 /3\;\vec{\nabla}.\vec{E}\; Tr(q^*) $ respectively. The
244 second term vanishes for charge free space
245 (i.e. $\vec{\nabla}.\vec{E} \; = \; 0)$. Similarly integration of the
246 1st term in the numerator produces
247 $8 \pi^2 /3\; Tr(q^*)\delta_{\alpha\beta}$ and the 2nd term produces
248 $8 \pi^2 /15k_B T (3{q}^*_{\alpha'\beta'}{q}^*_{\beta'\alpha'} -
249 {q}^*_{\alpha'\alpha'}{q}^*_{\beta'\beta'})\partial_\alpha E_\beta$,
250 if $\vec{\nabla}.\vec{E} \; = \; 0$,
251 $ \partial_\alpha E_\beta = \partial_\beta E_\alpha$ and
252 ${q}^*_{\alpha'\beta'}= {q}^*_{\beta'\alpha'}$. Therefore the
253 Boltzmann average of a quadrupole moment can be written as,
254
255 \begin{equation}
256 \braket{q_{\alpha\beta}}\; = \; \frac{1}{3} Tr(q^*)\;\delta_{\alpha\beta} + \frac{{\bar{q_o}}^2}{15k_BT}\;\partial_\alpha E_\beta,
257 \end{equation}
258 where $ \alpha_q = \frac{{\bar{q_o}}^2}{15k_BT} $ is a molecular quadrupolarizablity and ${\bar{q_o}}^2=
259 3{q}^*_{\alpha'\beta'}{q}^*_{\beta'\alpha'}-{q}^*_{\alpha'\alpha'}{q}^*_{\beta'\beta'}$ is a square of the net quadrupole moment of a molecule.
260
261 \section{Macroscopic Polarizability}
262 \label{sec:MacPolarizablity}
263
264 If we consider a system of dipolar or quadrupolar fluid in the external field perturbation, the net polarization of the system will still be proportional to the applied field perturbation.\cite{Chitanvis96, Stern-Feller03, Salvchov14, Salvchov14_2} In simulation the net polarization of the system is determined by the interaction of molecule with all other molecules as well as external field perturbation. Therefore the macroscopic polarizablity obtained from the simulation always varies with nature of real-space electrostatic interaction methods implemented in the simulation. To determine a susceptibility or dielectric constant of the material (which is a actual physical property of the dipolar or quadrupolar fluid) from the macroscopic polarizablity, we need to incorporate the interaction between molecules which has been discussed in detail in section \ref{sec:corrFactor}. In this section we discuss about the two different methods of calculating macroscopic polarizablity for both dipolar and quadrupolar fluid.
265
266 \subsection{External field perturbation}
267 \label{subsec:perturbation}
268 In the presence of uniform electric field $\textbf{E}^o$, a system of dipolar molecules polarizes along the direction of the applied field (or field gradient). Therefore the net dipolar polarization $ \textbf{P}$ of the system is,
269 \begin{equation}
270 \textbf{P} = \epsilon_o \alpha_{D}\; \textbf{E}^o.
271 \label{pertDipole}
272 \end{equation}
273 The constant $\alpha_D$ is a macroscopic polarizability, which is a property of the dipolar fluid in a given density and temperature.
274
275 Similarly, in the presence of external field gradient the system of quadrupolar molecule polarizes along the direction of applied field gradient therefore the net quadrupolar polarization of the system can be given by,
276 \begin{equation}
277 \begin{split}
278 & {Q}_{\alpha\beta} = \frac{1}{3}\; Tr({Q})\; \delta_{\alpha\beta} + \epsilon_o\; \alpha_Q \; \partial_{\alpha} E^o_{\beta}
279 \\ & or \\
280 & \frac{1}{3}\;\Theta_{\alpha\beta} = \epsilon_o\; \alpha_Q \; \partial_{\alpha} E^o_{\beta}
281 \end{split}
282 \label{pertQuad}
283 \end{equation}
284 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.
285
286
287 \subsection{Fluctuation formula}
288 \label{subsec:fluctuation}
289 For a system of molecules with net dipolar moment $\bf{M}$ at thermal equilibrium of temperature T in the presence of applied field $\bf{E}^o$, the average dipolar polarization can be expressed in terms of fluctuation of the net dipole moment as below,\cite{Stern03}
290 \begin{equation}
291 \braket{\bf{P}} = \epsilon_o \frac{\braket{\bf{M}^2}-{\braket{\bf{M}}}^2}{3 \epsilon_o V k_B T}\bf{E}^o
292 \label{flucDipole}
293 \end{equation}
294 This is similar to the formula for boltzmann average of single dipolar molecule in the subsection \ref{subsec:boltzAverage-Dipole}. Here $\braket{\bf{P}}$ is average polarization and $ \braket{\textbf{M}^2}-{\braket{\textbf{M}}}^2$ is the net dipole fluctuation at temperature T. For the limiting case $\textbf{E}^o \rightarrow 0 $, ensemble average of both net dipole moment $\braket{\textbf{M}}$ and dipolar polarization $\braket{\bf{P}}$ tends to vanish but $\braket{\bf{M}^2}$ will still be non-zero. The dipolar macroscopic polarizability can be written as,
295 \begin{equation}
296 \alpha_D = \frac{\braket{\bf{M}^2}-{\braket{\bf{M}}}^2}{3 \epsilon_o V k_B T}
297 \end{equation}
298 This is a macroscopic property of dipolar material which is true even if applied field $ \textbf{E}^o \rightarrow 0 $.
299
300 Analogous formula can also be written for a system with quadrupolar molecules,
301 \begin{equation}
302 \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}
303 \label{flucQuad}
304 \end{equation}
305 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,
306 \begin{equation}
307 \alpha_Q = \frac{\braket{\textbf{Q}^2}-{\braket{\textbf{Q}}}^2}{15 \epsilon_o V k_B T}
308 \label{propConstQuad}
309 \end{equation}
310
311
312 \section{Potential of mean force}
313 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 as,
314 \begin{equation}
315 w(r) = \int_{\inf}^{r}\braket{\frac{\partial f}{\partial r'}}dr',
316 \end{equation}
317 where $\braket{\partial f/\partial r'}$ is the mean constraint force.
318 Since the ions have a protecting Lennard-Jones (LJ) potential,
319 \begin{equation}
320 w(r) = w_{LJ}(r) + \frac{q_iq_j}{4\pi \epsilon_o \epsilon(r)}U_{method}(r).
321 \label{eq:pmf}
322 \end{equation}
323 Here $w_{LJ}$ is the PMF calculated without electrostatic interactions and $U_{method}(r)$ is the radial function for the charge-charge interaction, which is different for various real space truncation methods.
324
325 The quadrupole molecule can only couple with the gradient of the electric field and the region between two opposite point charges has both an electric field and a gradient of the electric field present. Therefore, this methodology should be usable to determine the dielectric constant for both the dipolar and quadrupolar fluid.
326 \label{sec:PMF}
327
328 \section{Correction factor}
329 \label{sec:corrFactor}
330 Since equations (\ref{pertDipole}, \ref{pertQuad}, \ref{flucDipole}, and \ref{flucQuad}) provide relation between polarization (dipolar or quadrupolar) and applied field (uniform field or field gradient), $\chi_d$ (or $ \chi_q$) is actually a macroscopic polarizability (or quadrupolarizability), which is different than the dipolar (or quadrupolar) susceptibility of the fluid. Actual constitutive relation should have a relation between polarization and Maxwell field (or field gradient) at different point in the sample. We can obtain susceptibility of the fluid from its macroscopic polarizability using correction factor evaluated below.
331 \subsection{Dipolar system}
332 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{Neumann83}
333 \begin{equation}
334 \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}')}.
335 \end{equation}
336
337 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}).
338 \begin{figure}
339 \includegraphics[width=3in]{DielectricFigure}
340 \caption{With the real-space electrostatic methods, the effective
341 dipole tensor, $\mathbf{T}$, governing interactions between
342 molecular dipoles is not the same for charge-charge interactions as
343 for point dipoles.}
344 \label{fig:stockmayer}
345 \end{figure}
346 In the case where point charges are interacting via an electrostatic
347 kernel, $v(r)$, the effective {\it molecular} dipole tensor,
348 $\mathbf{T}$ is obtained from two successive applications of the
349 gradient operator to the electrostatic kernel,
350 \begin{equation}
351 \mathbf{T}_{\alpha \beta}(r) = \nabla_\alpha \nabla_\beta \left(v(r)\right) = \delta_{\alpha \beta}
352 \left(\frac{1}{r} v^\prime(r) \right) + \frac{r_{\alpha}
353 r_{\beta}}{r^2} \left( v^{\prime \prime}(r) - \frac{1}{r}
354 v^{\prime}(r) \right)
355 \label{dipole-chargeTensor}
356 \end{equation}
357 where $v(r)$ may be either the bare kernel ($1/r$) or one of the
358 modified (Wolf or DSF) kernels. This tensor describes the effective
359 interaction between molecular dipoles ($\mathbf{D}$) in Gaussian
360 units as $-\mathbf{D} \cdot \mathbf{T} \cdot \mathbf{D}$.
361
362 When utilizing the new real-space methods for point dipoles, the
363 tensor is explicitly constructed,
364 \begin{equation}
365 \mathbf{T}_{\alpha \beta}(r) = \delta_{\alpha \beta} v_{21}(r) +
366 \frac{r_{\alpha} r_{\beta}}{r^2} v_{22}(r)
367 \label{dipole-diopleTensor}
368 \end{equation}
369 where the functions $v_{21}(r)$ and $v_{22}(r)$ depend on the level of
370 the approximation. Although the Taylor-shifted (TSF) and
371 gradient-shifted (GSF) models produce to the same $v(r)$ function for
372 point charges, they have distinct forms for the dipole-dipole
373 interactions.
374
375 Using constitutive relation, the polarization density $\textbf{P}(\textbf{r})$ is given by,
376 \begin{equation}
377 \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).
378 \label{constDipole}
379 \end{equation}
380 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{sec:perturbation} and \ref{sec: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{Neumann83, Jackson98}
381 \begin{equation}
382 \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}
383 \label{singular}
384 \end{equation}
385 Substituting equation (\ref{singular}) in the equation (\ref{constDipole}) we get,
386 \begin{equation}
387 \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).
388 \end{equation}
389 For both polarization and electric field homogeneous, this can be easily solved using Fourier transformation,
390 \begin{equation}
391 \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}).
392 \end{equation}
393 For homogeneous applied field Fourier component is non-zero only if $\kappa = 0$. Therefore,
394 \begin{equation}
395 \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}).
396 \label{fourierDipole}
397 \end{equation}
398 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,
399 \begin{equation}
400 \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}
401 \end{equation}
402 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,
403 \begin{equation}
404 \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}
405 \label{correctionFormula}
406 \end{equation}
407 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}.
408 \begin{table}
409 \caption{Expressions for the dipolar correction factor ($A$) for the real-space electrostatic methods in terms of the damping parameter
410 ($\alpha$) and the cutoff radius ($r_c$). The Ewald-Kornfeld result
411 derived in Refs. \onlinecite{Adams:1980rt,Adams:1981fr,Neumann83} is shown for comparison using the Ewald
412 convergence parameter ($\kappa$) and the real-space cutoff value ($r_c$). }
413 \label{tab:A}
414 {%
415 \begin{tabular}{l|c|c|c|}
416
417 Method & $A_\mathrm{charges}$ & $A_\mathrm{dipoles}$ \\
418 \hline
419 Spherical Cutoff (SC) & $\mathrm{erf}(r_c \alpha) - \frac{2 \alpha r_c}{\sqrt{\pi}} e^{-\alpha^2 r_c^2}$ & $\mathrm{erf}(r_c \alpha) - \frac{2 \alpha r_c}{\sqrt{\pi}} e^{-\alpha^2 r_c^2}$ \\
420 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} $\\
421 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} $ \\
422 Taylor-shifted (TSF) & 1 & 1 \\
423 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
424 \end{tabular}%
425 }
426 \end{table}
427 \subsection{Quadrupolar system}
428 In the presence of the field gradient $\partial_\alpha {E}_\beta $, a
429 non-vanishing quadrupolar polarization (quadrupole moment per unit
430 volume) $\bar{Q}_{\alpha\beta}$ will be induced in the entire volume
431 of a sample. The total field at any point $\vec{r}$ in the sample is
432 given by,
433 \begin{equation}
434 \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'
435 \label{gradMaxwell}
436 \end{equation}
437 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 $v(r)$.
438
439 \begin{equation}
440 \begin{split}
441 T_{\alpha\beta\gamma\delta}(r) &=\nabla_\alpha \nabla_\beta \nabla_\gamma \nabla_\delta\;v(r)
442 \\ &= \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)
443 \\ &+ \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)
444 \\ &+ 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),
445 \end{split}
446 \label{quadCharge}
447 \end{equation}
448 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,
449
450 \begin{equation}
451 \begin{split}
452 T_{\alpha\beta\gamma\delta}(r) &=\nabla_\alpha \nabla_\beta \;v_{\gamma\delta}(r)
453 \\ &= \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}
454 \\ &+ \delta_{\gamma\delta} r_\alpha r_\beta \left(\frac{v''_{21}(r)}{r^2}-\frac{v'_{21}(r)}{r^3} \right)
455 \\ &+\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)
456 \\ &+ 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),
457 \end{split}
458 \label{quadDip}
459 \end{equation}
460 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,
461 \begin{equation}
462 \begin{split}
463 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),
464 \end{split}
465 \label{quadRadial}
466 \end{equation}
467 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.
468 \begin{figure}
469 \includegraphics[width=3in]{QuadrupoleFigure}
470 \caption{With the real-space electrostatic methods, the effective
471 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.}
472 \label{fig:quadrupolarFluid}
473 \end{figure}
474 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
475 field gradient due to quadrupolar fluid vanishes at the singularity (see Appendix \ref{singularQuad}), equation (\ref{gradMaxwell}) can be written as,
476 \begin{equation}
477 \partial_\alpha E_\beta(\textbf{r}) = \partial_\alpha {E^o}_\beta(\textbf{r}) +
478 \frac{1}{4\pi \epsilon_o}\int\limits_{|\textbf{r}-\textbf{r}'|> 0 }
479 T_{\alpha\beta\gamma\delta}(|\textbf{r}-\textbf{r}'|)\;{Q}_{\gamma\delta}(\textbf{r}')\;
480 d^3r'.
481 \end{equation}
482 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}
483 \begin{equation}
484 \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',
485 \end{equation}
486 where $\Theta_{\alpha\beta} = 3Q_{\alpha\beta} - \delta_{\alpha\beta}Tr(Q)$
487 is the traceless quadrupole moment. The total quadrupolar polarization is written as,
488 \begin{equation}
489 {Q}_{\alpha\beta}(\textbf{r}) = \frac{1}{3}\delta_{\alpha\beta}\;Tr({Q})+\epsilon_o {\chi}^*_Q\;\partial_\alpha E_\beta(\textbf{r}),
490 \label{constQaud}
491 \end{equation}
492 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,
493 \begin{equation}
494 \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)
495 \end{equation}
496 For toroidal boundary conditions, both $\partial_\alpha E_\beta$ and
497 ${\Theta}_{\alpha\beta}$ are uniform over the entire space. After
498 performing a Fourier transform (see the Appendix in the Neumann's Paper \cite{Neumann83}) we get,
499 \begin{equation}
500 \frac{1}{3}{{\Theta}}_{\alpha\beta}({\kappa})=
501 \epsilon_o {\chi}^*_Q \;\left[{\partial_\alpha
502 {E^o}_\beta}({\kappa})+ \frac{1}{12\pi
503 \epsilon_o}\;{T}_{\alpha\beta\gamma\delta}({\kappa})\;
504 {{\Theta}}_{\gamma\delta}({\kappa})\right]
505 \end{equation}
506 Since the quadrupolar polarization is in the direction of the applied
507 field, we can write
508 ${{\Theta}}_{\gamma\delta}({\kappa}) =
509 {{\Theta}}_{\alpha\beta}({\kappa})$
510 and
511 ${T}_{\alpha\beta\gamma\delta}({\kappa}) =
512 {T}_{\alpha\beta\alpha\beta}({\kappa})$. Therefore we can consider each component of the interaction tensor as scalar and perform calculation.
513 \begin{equation}
514 \begin{split}
515 \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] \\
516 &= \epsilon_o {\chi}^*_Q\;\left(1-\frac{1}{4\pi} {\chi}^*_Q\;
517 {T}_{\alpha\beta\alpha\beta}({\kappa})\right)^{-1}
518 {\partial_\alpha E^o_\beta}({\kappa})
519 \end{split}
520 \label{fourierQuad}
521 \end{equation}
522 If the field gradient is homogeneous over the
523 entire volume, ${\partial_ \alpha E_\beta}({\kappa}) = 0 $ except at
524 $ {\kappa} = 0$, hence it is sufficient to know
525 ${T}_{\alpha\beta\alpha\beta}({\kappa})$ at $ {\kappa} =
526 0$. Therefore equation (\ref{fourierQuad}) can be written as,
527 \begin{equation}
528 \begin{split}
529 \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})
530 \end{split}
531 \label{fourierQuad2}
532 \end{equation}
533 where $ {T}_{\alpha\beta\alpha\beta}({0})$ can be evaluated as,
534 \begin{equation}
535 {T}_{\alpha\beta\alpha\beta}({0}) = \int {T}_{\alpha\beta\alpha\beta}\;(\textbf{r})d^3r
536 \label{realTensorQaud}
537 \end{equation}
538
539 In terms of traced quadrupole moment equation (\ref{fourierQuad2}) can be written as,
540 \begin{equation}
541 {{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
542 \label{tracedConstQuad}
543 \end{equation}
544 Comparing (\ref{tracedConstQuad}) and (\ref{flucQuad}) we get,
545 \begin{equation}
546 \begin{split}
547 &\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}, \\
548 &{\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}
549 \end{split}
550 \end{equation}
551 Finally the quadrupolar susceptibility cab be written in terms of quadrupolar correction factor ($A_{quad}$) as below,
552 \begin{equation}
553 {\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}
554 \label{eq:quadrupolarSusceptiblity}
555 \end{equation}
556 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}
557
558 \begin{equation}
559 \epsilon = 1 + \chi^*_Q\; G = 1 + G \; \alpha_Q\left(1 + \alpha_Q\; A_{quad}\right)^{-1}
560 \label{eq:dielectricFromQuadrupoles}
561 \end{equation}
562 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,
563 \begin{equation}
564 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}
565 \label{eq:geometricalFactor}
566 \end{equation}
567 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.
568 \begin{table}
569 \caption{Expressions for the quadrupolar correction factor ($A$) for the real-space electrostatic methods in terms of the damping parameter
570 ($\alpha$) and the cutoff radius ($r_c$). The dimension of the correction factor is $ length^{-2}$ in case of quadrupolar fluid.}
571 \label{tab:B}
572 \centering
573 \resizebox{\columnwidth}{!}{%
574
575 \begin{tabular}{l|c|c|c|c|}
576
577 Method & $A_\mathrm{charges}$ & $A_\mathrm{dipoles}$ &$A_\mathrm{quadrupoles}$ \\\hline
578 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}$ \\
579 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}$ \\
580 Gradient-shifted (GSF) & $- \frac{8 \alpha^5 {r_c}^3}{3\sqrt{\pi}} e^{-\alpha^2 r_c^2}$ & 0 & $-\frac{4{\alpha}^7{r_c}^5 }{9\sqrt{\pi}}e^{-\alpha^2 r_c^2}(-1+2\alpha ^2 r_c^2)$\\
581 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
582 \end{tabular}%
583 }
584 \end{table}
585 \section{Methodology}
586 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. In case of dipolar system we calculated macroscopic polarzability using first two methods separately and derived the dielectric constant utilizing equation (\ref{correctionFormula}). Similarly we used equation (\ref{eq:pmf}) to calculate dielectric constant 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 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. The parameters used in the test systems are given in table \ref{Tab:C}.
587
588 \begin{table}
589 \caption{\label{Tab:C}}
590 \begin{tabularx}{\textwidth}{r|cc|YYccc|Yccc} \hline
591 & \multicolumn{2}{c|}{LJ parameters} &
592 \multicolumn{5}{c|}{Electrostatic moments} & & & & \\
593 Test system & $\sigma$& $\epsilon$ & $C$ & $D$ &
594 $Q_{xx}$ & $Q_{yy}$ & $Q_{zz}$ & mass & $I_{xx}$ & $I_{yy}$ &
595 $I_{zz}$ \\ \cline{6-8}\cline{10-12}
596 & (\AA) & (kcal/mol) & (e) & (debye) & \multicolumn{3}{c|}{(debye \AA)} & (amu) & \multicolumn{3}{c}{(amu
597 \AA\textsuperscript{2})} \\ \hline
598 Stockmayer fluid & 3.41 & 0.2381 & - & 1.4026 &-&-&-& 39.948 & 11.613 & 11.613 & 0.0 \\
599 Quadrupolar fluid & 2.985 & 0.265 & - & - & 0.0 & 0.0 &-2.139 & 18.0153 & 43.0565 & 43.0565 & 0.0 \\
600 \ce{q+} & 1.0 & 0.1 & +1 & - & - & - & - & 22.98 & - & - & - \\
601 \ce{q-} & 1.0 & 0.1 & -1 & - & - & - & - & 22.98 & - & - & - \\ \hline
602 \end{tabularx}
603 \end{table}
604
605 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 evaluate ensemble average of the box dipole or quadrupole moment as a response field or field gradient. The macroscopic polarizability of the system is derived using ratio between system multipolar moment and applied field or field gradient. This method works properly only at the linear response region of field or field gradient.
606
607 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 i.e. $\braket{A^2} - \braket{A}^2 $ is measured at the fixed temperature and density for a given multipolar fluid. Finally the macroscopic polraizability of the system at a particular density is derived using equation (\ref{flucQuad}).
608
609 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 different constraint distances. Finally we calculated dielectric constant 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 run to reduce simulation error.
610
611 \subsection{Implementation}
612 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. The short range interaction in the simulations is incorporated with 6-12 Lennard Jones interaction method.
613
614 To derive the box multipolar (dipolar or quadrupolar) moment, we added the component each individual molecule in the space frame and taken ensemble average of the snapshots of the whole 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 quadrupole moment. Similarly the first term in the quadrupolar system is derived using relation $ \braket{Q^2} = \braket{3 Q:Q - TrQ^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}.
615 \subsection{Model systems}
616 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 are 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.
617
618 Quadrupolar systems consists 4000 linear point quadrupolar molecules 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 of the quadrupole moment the components of box quadrupole moments are printed every 100 fs. Each simulations are repeated at different values of applied constant gradients from $ 0 - 9 \times 10^{-2}\; V/\r{A}^2 $. To find 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\;\r{A} - 12\; \r{A} $ at the step of $0.1\; \r{A} $ 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.
619
620 \section{Results}
621
622 \section{Conclusion}
623
624 \newpage
625
626 \appendix
627 \section{Point-multipolar interactions with a spatially-varying electric field}
628
629 We can treat objects $a$, $b$, and $c$ containing embedded collections
630 of charges. When we define the primitive moments, we sum over that
631 collections of charges using a local coordinate system within each
632 object. The point charge, dipole, and quadrupole for object $a$ are
633 given by $C_a$, $\mathbf{D}_a$, and $\mathsf{Q}_a$, respectively.
634 These are the primitive multipoles which can be expressed as a
635 distribution of charges,
636 \begin{align}
637 C_a =&\sum_{k \, \text{in }a} q_k , \label{eq:charge} \\
638 D_{a\alpha} =&\sum_{k \, \text{in }a} q_k r_{k\alpha}, \label{eq:dipole}\\
639 Q_{a\alpha\beta} =& \frac{1}{2} \sum_{k \, \text{in } a} q_k
640 r_{k\alpha} r_{k\beta} . \label{eq:quadrupole}
641 \end{align}
642 Note that the definition of the primitive quadrupole here differs from
643 the standard traceless form, and contains an additional Taylor-series
644 based factor of $1/2$. In Paper 1, we derived the forces and torques
645 each object exerts on the others.
646
647 Here we must also consider an external electric field that varies in
648 space: $\mathbf E(\mathbf r)$. Each of the local charges $q_k$ in
649 object $a$ will then experience a slightly different field. This
650 electric field can be expanded in a Taylor series around the local
651 origin of each object. A different Taylor series expansion is carried
652 out for each object.
653
654 For a particular charge $q_k$, the electric field at that site's
655 position is given by:
656 \begin{equation}
657 E_\gamma + \nabla_\delta E_\gamma r_{k \delta}
658 + \frac {1}{2} \nabla_\delta \nabla_\varepsilon E_\gamma r_{k \delta}
659 r_{k \varepsilon} + ...
660 \end{equation}
661 Note that the electric field is always evaluated at the origin of the
662 objects, and treating each object using point multipoles simplifies
663 this greatly.
664
665 To find the force exerted on object $a$ by the electric field, one
666 takes the electric field expression, and multiplies it by $q_k$, and
667 then sum over all charges in $a$:
668
669 \begin{align}
670 F_\gamma &= \sum_{k \textrm{~in~} a} q_k \lbrace E_\gamma + \nabla_\delta E_\gamma r_{k \delta}
671 + \frac {1}{2} \nabla_\delta \nabla_\varepsilon E_\gamma r_{k \delta}
672 r_{k \varepsilon} + ... \rbrace \\
673 &= C_a E_\gamma + D_{a \delta} \nabla_\delta E_\gamma
674 + Q_{a \delta \varepsilon} \nabla_\delta \nabla_\varepsilon E_\gamma +
675 ...
676 \end{align}
677
678 Similarly, the torque exerted by the field on $a$ can be expressed as
679 \begin{align}
680 \tau_\alpha &= \sum_{k \textrm{~in~} a} (\mathbf r_k \times q_k \mathbf E)_\alpha \\
681 & = \sum_{k \textrm{~in~} a} \epsilon_{\alpha \beta \gamma} q_k
682 r_{k\beta} E_\gamma(\mathbf r_k) \\
683 & = \epsilon_{\alpha \beta \gamma} D_\beta E_\gamma
684 + 2 \epsilon_{\alpha \beta \gamma} Q_{\beta \delta} \nabla_\delta
685 E_\gamma + ...
686 \end{align}
687
688 The last term is essentially identical with form derived by Torres del
689 Castillo and M\'{e}ndez Garrido,\cite{Torres-del-Castillo:2006uo} although their derivation
690 utilized a traceless form of the quadrupole that is different than the
691 primitive definition in use here. We note that the Levi-Civita symbol
692 can be eliminated by utilizing the matrix cross product in an
693 identical form as in Ref. \onlinecite{Smith98}:
694 \begin{equation}
695 \left[\mathsf{A} \times \mathsf{B}\right]_\alpha = \sum_\beta
696 \left[\mathsf{A}_{\alpha+1,\beta} \mathsf{B}_{\alpha+2,\beta}
697 -\mathsf{A}_{\alpha+2,\beta} \mathsf{B}_{\alpha+1,\beta}
698 \right]
699 \label{eq:matrixCross}
700 \end{equation}
701 where $\alpha+1$ and $\alpha+2$ are regarded as cyclic permuations of
702 the matrix indices. In table \ref{tab:UFT} we give compact
703 expressions for how the multipole sites interact with an external
704 field that has exhibits spatial variations.
705
706 \begin{table}
707 \caption{Potential energy $(U)$, force $(\mathbf{F})$, and torque
708 $(\mathbf{\tau})$ expressions for a multipolar site embedded in an
709 electric field with spatial variations, $\mathbf{E}(\mathbf{r})$.
710 \label{tab:UFT}}
711 \begin{tabular}{r|ccc}
712 & Charge & Dipole & Quadrupole \\ \hline
713 $U$ & $C \phi(\mathbf{r})$ & $-\mathbf{D} \cdot \mathbf{E}(\mathbf{r})$ & $- \mathsf{Q}:\nabla \mathbf{E}(\mathbf{r})$ \\
714 $\mathbf{F}$ & $C \mathbf{E}(\mathbf{r})$ & $+\mathbf{D} \cdot \nabla \mathbf{E}(\mathbf{r})$ & $+\mathsf{Q} : \nabla\nabla\mathbf{E}(\mathbf{r})$ \\
715 $\mathbf{\tau}$ & & $\mathbf{D} \times \mathbf{E}(\mathbf{r})$ & $+2 \mathsf{Q} \times \nabla \mathbf{E}(\mathbf{r})$
716 \end{tabular}
717 \end{table}
718 \section{Gradient of the field due to quadrupolar polarization}
719 \label{singularQuad}
720 In this section, we will discuss the gradient of the field produced by
721 quadrupolar polarization. For this purpose, we consider a distribution
722 of charge ${\rho}(r)$ which gives rise to an electric field
723 $\vec{E}(r)$ and gradient of the field $\vec{\nabla} \vec{E}(r)$
724 throughout space. The total gradient of the electric field over volume
725 due to the all charges within the sphere of radius $R$ is given by
726 (cf. Jackson equation 4.14):
727 \begin{equation}
728 \int_{r<R} \vec{\nabla}\vec{E}\;d^3r = -\int_{r=R} R^2 \vec{E}\;\hat{n}\; d\Omega
729 \label{eq:8}
730 \end{equation}
731 where $d\Omega$ is the solid angle and $\hat{n}$ is the normal vector
732 of the surface of the sphere which is equal to
733 $sin[\theta]cos[\phi]\hat{x} + sin[\theta]sin[\phi]\hat{y} +
734 cos[\theta]\hat{z}$
735 in spherical coordinates. For the charge density ${\rho}(r')$, the
736 total gradient of the electric field can be written as (cf. Jackson
737 equation 4.16),
738 \begin{equation}
739 \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
740 \label{eq:9}
741 \end{equation}
742 The radial function in the equation (\ref{eq:9}) can be expressed in
743 terms of spherical harmonics as (cf. Jackson equation 3.70),
744 \begin{equation}
745 \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)
746 \label{eq:10}
747 \end{equation}
748 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,
749 \begin{equation}
750 \begin{split}
751 \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 \\
752 &= -\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
753 '
754 \end{split}
755 \label{eq:11}
756 \end{equation}
757 The gradient of the product of radial function and spherical harmonics
758 is given by (cf. Arfken, p.811 eq. 16.94):
759 \begin{equation}
760 \begin{split}
761 \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
762 {\partial}{\partial r}+\frac{l}{r} \right]f(r)\; Y_{l, l-1, m}(\theta, \phi).
763 \end{split}
764 \label{eq:12}
765 \end{equation}
766 Using equation (\ref{eq:12}) we get,
767 \begin{equation}
768 \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}},
769 \label{eq:13}
770 \end{equation}
771 where $ Y_{l,l+1,m}(\theta, \phi)$ is the vector spherical harmonics
772 which can be expressed in terms of spherical harmonics as shown in
773 below (cf. Arfkan p.811),
774 \begin{equation}
775 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},
776 \label{eq:14}
777 \end{equation}
778 where $C(l+1,1,l|m_1,m_2,m)$ is a Clebsch-Gordan coefficient and
779 $\hat{e}_{m_2}$ is a spherical tensor of rank 1 which can be expressed
780 in terms of Cartesian coordinates,
781 \begin{equation}
782 {\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}}
783 \label{eq:15}
784 \end{equation}
785 The normal vector $\hat{n} $ can be expressed in terms of spherical tensor of rank 1 as shown in below,
786 \begin{equation}
787 \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)
788 \label{eq:16}
789 \end{equation}
790 The surface integral of the product of $\hat{n}$ and
791 ${Y_{l+1}}^{m_1}(\theta, \phi)$ gives,
792 \begin{equation}
793 \begin{split}
794 \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 \\
795 &= \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 \\
796 &= \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),
797 \end{split}
798 \label{eq:17}
799 \end{equation}
800 where ${Y_{l}}^{-m} = (-1)^m\;{{Y_{l}}^{m}}^* $ and
801 $ \int {{Y_{l}}^{m}}^*\;{Y_{l'}}^{m'}\;d\Omega =
802 \delta_{ll'}\delta_{mm'} $.
803 Non-vanishing values of equation \ref{eq:17} require $l = 0$,
804 therefore the value of $ m = 0 $. Since the values of $ m_1$ are -1,
805 1, and 0 then $m_2$ takes the values 1, -1, and 0, respectively
806 provided that $m = m_1 + m_2$. Equation \ref{eq:11} can therefore be
807 modified,
808 \begin{equation}
809 \begin{split}
810 \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(
811 1, 1, 0|0,0,0)\;{\hat{e}_{0}}{\hat{e}_{0}} ]\; d^3r'.
812 \end{split}
813 \label{eq:18}
814 \end{equation}
815 After substituting ${Y^*}_{00} = \frac{1}{\sqrt{4\pi}} $ and using the
816 values of the Clebsch-Gorden coefficients: $ C(1, 1, 0|-1,1,0) =
817 \frac{1}{\sqrt{3}}, \; C(1, 1, 0|-1,1,0)= \frac{1}{\sqrt{3}}$ and $
818 C(1, 1, 0|0,0,0) = -\frac{1}{\sqrt{3}}$ in equation \ref{eq:18} we
819 obtain,
820 \begin{equation}
821 \begin{split}
822 \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)\\
823 &= - \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).
824 \end{split}
825 \label{eq:19}
826 \end{equation}
827 Equation (\ref{eq:19}) gives the total gradient of the field over a
828 sphere due to the distribution of the charges. For quadrupolar fluids
829 the total charge within a sphere is zero, therefore
830 $ \int_{r<R} \vec{\nabla}\vec{E}\;d^3r = 0 $. Hence the quadrupolar
831 polarization produces zero net gradient of the field inside the
832 sphere.
833
834 \section{Applied field or field gradient}
835 \label{Ap:fieldOrGradient}
836
837 To satisfy the condition $ \nabla . E = 0 $, within the box of molecules we have taken electrostatic potential in the following form
838 \begin{equation}
839 \begin{split}
840 \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. \\
841 & \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),
842 \end{split}
843 \label{eq:appliedPotential}
844 \end{equation}
845 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,
846 \[\bf{E}
847 =\frac{g_o}{2} \left(\begin{array}{ccc}
848 2(a_1\; b_1 - \frac{cos\psi}{3})\;x \;+ (a_1\; b_2 \;+ a_2\; b_1)\;y + (a_1\; b_3 \;+ a_3\; b_1)\;z \\
849 (a_2\; b_1 \;+ a_1\; b_2)\;x + 2(a_2\; b_2 \;- \frac{cos\psi}{3})\;y + (a_2\; b_3 \;+ a_3\; b_3)\;z \\
850 (a_3\; b_1 \;+ a_3\; b_2)\;x + (a_3\; b_2 \;+ a_2\; b_3)y + 2(a_3\; b_3 \;- \frac{cos\psi}{3})\;z
851 \end{array} \right).\]
852 The gradient of the applied field derived from the potential can be written in the following form,
853 \[\nabla\bf{E}
854 = \frac{g_o}{2}\left(\begin{array}{ccc}
855 2(a_1\; b_1 - \frac{cos\psi}{3}) & (a_1\; b_2 \;+ a_2\; b_1) & (a_1\; b_3 \;+ a_3\; b_1)\;z \\
856 (a_2\; b_1 \;+ a_1\; b_2) & 2(a_2\; b_2 \;- \frac{cos\psi}{3}) & (a_2\; b_3 \;+ a_3\; b_3)\;z \\
857 (a_3\; b_1 \;+ a_3\; b_2) & (a_3\; b_2 \;+ a_2\; b_3) & 2(a_3\; b_3 \;- \frac{cos\psi}{3})\;z
858 \end{array} \right).\]
859 \newpage
860
861 \bibliography{multipole}
862
863 \end{document}