ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/multipole/dielectric_new.tex
Revision: 4396
Committed: Wed Dec 16 19:54:40 2015 UTC (8 years, 6 months ago) by gezelter
Content type: application/x-tex
File size: 61990 byte(s)
Log Message:
edits

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 We report on the dielectric properties of the shifted potential
40 (SP), gradient shifted force (GSF), and Taylor shifted force (TSF)
41 real-space methods for multipole interactions that were developed in
42 the first two papers in this series. We find that some subtlety is
43 required for computing dielectric properties with the real-space
44 methods, particularly when using the common fluctuation formulae.
45 Three distinct methods for computing the dielectric constant are
46 investigated, including the standard fluctuation formulae,
47 potentials of mean force between solvated ions, and direct
48 measurement of linear solvent polarization in response to applied
49 fields and field gradients.
50 \end{abstract}
51
52 \maketitle
53
54 \section{Introduction}
55
56 Over the past several years, there has been increasing interest in
57 pairwise or ``real space'' methods for computing electrostatic
58 interactions in condensed phase
59 simulations.\cite{Wolf99,Zahn02,Kast03,Beckd.A.C._Bi0486381,Ma05,Fennell06}
60 These techniques were initially developed by Wolf {\it et al.} in
61 their work towards an $\mathcal{O}(N)$ Coulombic sum.\cite{Wolf99}
62 Wolf's method of using cutoff neutralization and electrostatic damping
63 is able to obtain excellent agreement with Madelung energies in ionic
64 crystals.\cite{Wolf99}
65
66 Zahn \textit{et al.}\cite{Zahn02} and Fennell and Gezelter extended
67 this method using shifted force approximations at the cutoff distance
68 in order to conserve total energy in molecular dynamics
69 simulations.\cite{Fennell06} In the previous two papers in this
70 series, we developed three generalized real space methods: shifted
71 potential (SP), gradient shifted force (GSF), and Taylor shifted force
72 (TSF).\cite{PaperI, PaperII} These methods evaluate electrostatic
73 interactions for charges and higher order multipoles using a
74 finite-radius cutoff sphere. The neutralization and damping of local
75 moments within the cutoff sphere is a multipolar generalization of
76 Wolf's sum. In the GSF and TSF methods, additional terms are added to
77 the potential energy so that forces and torques also vanish smoothly
78 at the cutoff radius. This ensures that the total energy is conserved
79 in molecular dynamics simulations.
80
81 One of the most difficult tests of any new electrostatic method is the
82 fidelity with which that method can reproduce the bulk-phase
83 polarizability or equivalently, the dielectric properties of a
84 fluid. Before the advent of computer simulations, Kirkwood and Onsager
85 developed fluctuation formulae for the dielectric properties of
86 dipolar fluids.\cite{Kirkwood39,Onsagar36}
87
88 Of particular interest is the static dielectric constant, $\epsilon$.
89 Using the Ewald sum under tin-foil boundary conditions, $\epsilon$ can
90 be calculated for systems of non-polarizable substances via
91 \begin{equation}
92 \epsilon = 1 + \frac{\langle M^2\rangle}{3\epsilon_0k_\textrm{B}TV},
93 \label{eq:staticDielectric}
94 \end{equation}
95 where $\epsilon_0$ is the permittivity of free space and
96 $\langle M^2\rangle$ is the fluctuation of the system dipole
97 moment.\cite{Allen:1989fp} The numerator in the fractional term in
98 equation (\ref{eq:staticDielectric}) is identical to the quantity
99 calculated in the finite-system Kirkwood $g$ factor ($G_k$):
100 \begin{equation}
101 G_k = \frac{\langle M^2\rangle}{N\mu^2},
102 \label{eq:KirkwoodFactor}
103 \end{equation}
104 where $\mu$ is the dipole moment of a single molecule of the
105 homogeneous system.\cite{Neumann:1983mz,Neumann:1983yq,Neumann84,Neumann85} The
106 fluctuation term in both equation (\ref{eq:staticDielectric}) and
107 (\ref{eq:KirkwoodFactor}) is calculated as follows,
108 \begin{equation}
109 \begin{split}
110 \langle M^2\rangle &= \langle\bm{M}\cdot\bm{M}\rangle
111 - \langle\bm{M}\rangle\cdot\langle\bm{M}\rangle \\
112 &= \langle M_x^2+M_y^2+M_z^2\rangle
113 - (\langle M_x\rangle^2 + \langle M_x\rangle^2
114 + \langle M_x\rangle^2).
115 \end{split}
116 \label{eq:fluctBoxDipole}
117 \end{equation}
118 This fluctuation term can be accumulated during a simulation; however,
119 it converges rather slowly, thus requiring multi-nanosecond simulation
120 times.\cite{Horn04} In the case of tin-foil boundary conditions, the
121 dielectric/surface term of the Ewald summation is equal to zero.
122
123 Similar formulae were developed by Logan \textit{et al.} for the bulk
124 polarizability of quadrupolar
125 fluids.\cite{LoganI81,LoganII82,LoganIII82}
126
127 In modern simulations, bulk materials are usually treated using
128 periodic replicas of small regions, and this level of approximation
129 requires corrections to the fluctuation formulae that were derived for
130 the bulk fluids. In 1983 Neumann proposed a general formula for
131 evaluating dielectric properties of dipolar fluids using real-space
132 cutoff methods.\cite{Neumann83} Steinhauser and Neumann used this
133 formula to evaluate the corrected dielectric constant for the
134 Stockmayer fluid using two different methods: Ewald-Kornfield (EK) and
135 reaction field (RF) methods.\cite{Neumann-Steinhauser83}
136
137 Zahn \textit{et al.}\cite{Zahn02} utilized this approach and evaluated
138 the correction factor for using damped shifted charge-charge
139 kernel. This was later generalized by Izvekov \textit{et
140 al.},\cite{Izvekov:2008wo} who that the expression for the
141 dielectric constant reduces to widely-used \textit{conducting
142 boundary} formula for real-space cutoff methods that have first
143 derivatives that vanish at the cutoff sphere.
144
145 In quadrupolar fluids, the relationship between quadrupolar
146 susceptibility and the dielectric constant is not as straightforward
147 as in the dipolar case. The dielectric constant depends on the
148 geometry of the external field perturbation.\cite{Ernst92} Significant
149 efforts have been made to increase our understanding the dielectric
150 properties of these fluids,\cite{JeonI03,JeonII03,Chitanvis96}
151 although a correction formula for different real-space methods has not
152 yet been developed.
153
154 In this paper we derive general formulae for calculating the
155 dielectric properties of quadrupolar fluids. We also evaluate the
156 correction factor for SP, GSF, and TSF methods for both dipolar and
157 quadrupolar fluids interacting via point charges, point dipoles or
158 through quadrupole-quadrupole interactions.
159
160 We have also calculated the screening behavior for two ions immersed
161 in a quadrupolar fluid to estimate the dielectric screening from the
162 quadrupolar susceptibility. We have used three different methods to
163 compare our results with computer simulations:
164 \begin{enumerate}
165 \item responses of the fluid to external perturbations,
166 \item fluctuations of box multipole moments, and
167 \item potentials of mean force between solvated ions,
168 \end{enumerate}
169
170 In the external field perturbation, the net polarization of the system
171 is observed as a linear response of the applied field perturbation,
172 where proportionality constant is determined by the electrostatic
173 interaction between the electrostatic multipoles at a given
174 temperature. The fluctuation formula observes the time average
175 fluctuation of the multipolar moment as a function of temperature. The
176 average fluctuation value of the system is determined by the
177 multipole-multipole interactions between molecules at a given
178 temperature. Since the expression of the electrostatic interaction
179 energy, force, and torque in the real space electrostatic methods are
180 different from their original definition, both fluctuation and
181 external field perturbation formula should also be modified
182 accordingly. The potential of mean force method calculates dielectric
183 constant from the potential energy between ions before and after
184 dielectric material is introduced. All of these different methods for
185 calculating dielectric properties will be discussed in detail in the
186 following sections: \ref{subsec:perturbation},
187 \ref{subsec:fluctuation}, and \ref{sec:PMF}.
188
189 \section{Boltzmann average for orientational polarization}
190 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}.
191
192 \subsection{Dipole}
193 \label{subsec:boltzAverage-Dipole}
194 Consider a system of molecules, each with permanent dipole moment
195 $p_o$. In the absense of external field, thermal agitation orients the
196 dipoles randomly, reducing the system moment to zero. External fields
197 will tend to line up the dipoles in the direction of applied field.
198 Here we have considered net field from all other molecules is
199 considered to be zero. Therefore the total Hamiltonian of each
200 molecule is,\cite{Jackson98}
201 \begin{equation}
202 H = H_o - \bf{p_o} .\bf{E},
203 \end{equation}
204 where $H_o$ is a function of the internal coordinates of the molecule.
205 The Boltzmann average of the dipole moment is given by,
206 \begin{equation}
207 \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}}},
208 \end{equation}
209 where $\bf{E}$ is selected along z-axis. If we consider that the
210 applied field is small, \textit{i.e.} $\frac{p_oE\; cos\theta}{k_B T} << 1$,
211 \begin{equation}
212 \braket{p_{mol}} \approx \frac{1}{3}\frac{{p_o}^2}{k_B T}E,
213 \end{equation}
214 where $ \alpha_p = \frac{1}{3}\frac{{p_o}^2}{k_B T}$ is a molecular
215 polarizability. The orientational polarization depends inversely on
216 the temperature and applied field must overcome the thermal agitation.
217
218 \subsection{Quadrupole}
219 \label{subsec:boltzAverage-Quad}
220 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}
221 \begin{equation}
222 \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}}},
223 \label{boltzQuad}
224 \end{equation}
225 where $\int d\Omega = \int_0^{2\pi} \int_0^\pi \int_0^{2\pi}
226 sin\theta\; d\theta\ d\phi\ d\psi$ is the integration over Euler
227 angles, $ H = H_o -q_{\mu\nu}\;\partial_\nu E_\mu $ is the energy of
228 a quadrupole in the gradient of the
229 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,
230 \begin{equation}
231 \begin{split}
232 &q_{\alpha\beta} = \eta_{\alpha\alpha'}\;\eta_{\beta\beta'}\;{q}^* _{\alpha'\beta'} \\
233 &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.
234 \end{split}
235 \label{energyQuad}
236 \end{equation}
237 Here the starred tensors are the components in the body fixed
238 frame. Substituting equation (\ref{energyQuad}) in the equation (\ref{boltzQuad})
239 and taking linear terms in the expansion we get,
240 \begin{equation}
241 \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)},
242 \end{equation}
243 where $\eta_{\alpha\alpha'}$ is the inverse of the rotation matrix that transforms
244 the body fixed co-ordinates to the space co-ordinates,
245 \[\eta_{\alpha\alpha'}
246 = \left(\begin{array}{ccc}
247 cos\phi\; cos\psi - cos\theta\; sin\phi\; sin\psi & -cos\theta\; cos\psi\; sin\phi - cos\phi\; sin\psi & sin\theta\; sin\phi \\
248 cos\psi\; sin\phi + cos\theta\; cos\phi \; sin\psi & cos\theta\; cos\phi\; cos\psi - sin\phi\; sin\psi & -cos\phi\; sin\theta \\
249 sin\theta\; sin\psi & -cos\psi\; sin\theta & cos\theta
250 \end{array} \right).\]
251 Integration of 1st and 2nd terms in the denominator gives $8 \pi^2$
252 and $8 \pi^2 /3\;\vec{\nabla}.\vec{E}\; Tr(q^*) $ respectively. The
253 second term vanishes for charge free space
254 (i.e. $\vec{\nabla}.\vec{E} \; = \; 0)$. Similarly integration of the
255 1st term in the numerator produces
256 $8 \pi^2 /3\; Tr(q^*)\delta_{\alpha\beta}$ and the 2nd term produces
257 $8 \pi^2 /15k_B T (3{q}^*_{\alpha'\beta'}{q}^*_{\beta'\alpha'} -
258 {q}^*_{\alpha'\alpha'}{q}^*_{\beta'\beta'})\partial_\alpha E_\beta$,
259 if $\vec{\nabla}.\vec{E} \; = \; 0$,
260 $ \partial_\alpha E_\beta = \partial_\beta E_\alpha$ and
261 ${q}^*_{\alpha'\beta'}= {q}^*_{\beta'\alpha'}$. Therefore the
262 Boltzmann average of a quadrupole moment can be written as,
263
264 \begin{equation}
265 \braket{q_{\alpha\beta}}\; = \; \frac{1}{3} Tr(q^*)\;\delta_{\alpha\beta} + \frac{{\bar{q_o}}^2}{15k_BT}\;\partial_\alpha E_\beta,
266 \end{equation}
267 where $ \alpha_q = \frac{{\bar{q_o}}^2}{15k_BT} $ is a molecular quadrupolarizablity and ${\bar{q_o}}^2=
268 3{q}^*_{\alpha'\beta'}{q}^*_{\beta'\alpha'}-{q}^*_{\alpha'\alpha'}{q}^*_{\beta'\beta'}$ is a square of the net quadrupole moment of a molecule.
269
270 \section{Macroscopic Polarizability}
271 \label{sec:MacPolarizablity}
272
273 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.
274
275 \subsection{External field perturbation}
276 \label{subsec:perturbation}
277 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,
278 \begin{equation}
279 \textbf{P} = \epsilon_o \alpha_{D}\; \textbf{E}^o.
280 \label{pertDipole}
281 \end{equation}
282 The constant $\alpha_D$ is a macroscopic polarizability, which is a property of the dipolar fluid in a given density and temperature.
283
284 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,
285 \begin{equation}
286 \begin{split}
287 & {Q}_{\alpha\beta} = \frac{1}{3}\; Tr({Q})\; \delta_{\alpha\beta} + \epsilon_o\; \alpha_Q \; \partial_{\alpha} E^o_{\beta}
288 \\ & or \\
289 & \frac{1}{3}\;\Theta_{\alpha\beta} = \epsilon_o\; \alpha_Q \; \partial_{\alpha} E^o_{\beta}
290 \end{split}
291 \label{pertQuad}
292 \end{equation}
293 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.
294
295
296 \subsection{Fluctuation formula}
297 \label{subsec:fluctuation}
298 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}
299 \begin{equation}
300 \braket{\bf{P}} = \epsilon_o \frac{\braket{\bf{M}^2}-{\braket{\bf{M}}}^2}{3 \epsilon_o V k_B T}\bf{E}^o
301 \label{flucDipole}
302 \end{equation}
303 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,
304 \begin{equation}
305 \alpha_D = \frac{\braket{\bf{M}^2}-{\braket{\bf{M}}}^2}{3 \epsilon_o V k_B T}
306 \end{equation}
307 This is a macroscopic property of dipolar material which is true even if applied field $ \textbf{E}^o \rightarrow 0 $.
308
309 Analogous formula can also be written for a system with quadrupolar molecules,
310 \begin{equation}
311 \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}
312 \label{flucQuad}
313 \end{equation}
314 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,
315 \begin{equation}
316 \alpha_Q = \frac{\braket{\textbf{Q}^2}-{\braket{\textbf{Q}}}^2}{15 \epsilon_o V k_B T}
317 \label{propConstQuad}
318 \end{equation}
319
320
321 \section{Potential of mean force}
322 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,
323 \begin{equation}
324 w(r) = \int_{\inf}^{r}\braket{\frac{\partial f}{\partial r'}}dr',
325 \end{equation}
326 where $\braket{\partial f/\partial r'}$ is the mean constraint force.
327 Since the ions have a protecting Lennard-Jones (LJ) potential,
328 \begin{equation}
329 w(r) = w_{LJ}(r) + \frac{q_iq_j}{4\pi \epsilon_o \epsilon(r)}U_{method}(r).
330 \label{eq:pmf}
331 \end{equation}
332 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.
333
334 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.
335 \label{sec:PMF}
336
337 \section{Correction factor}
338 \label{sec:corrFactor}
339 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.
340 \subsection{Dipolar system}
341 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}
342 \begin{equation}
343 \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}')}.
344 \end{equation}
345
346 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}).
347 \begin{figure}
348 \includegraphics[width=3in]{DielectricFigure}
349 \caption{With the real-space electrostatic methods, the effective
350 dipole tensor, $\mathbf{T}$, governing interactions between
351 molecular dipoles is not the same for charge-charge interactions as
352 for point dipoles.}
353 \label{fig:stockmayer}
354 \end{figure}
355 In the case where point charges are interacting via an electrostatic
356 kernel, $v(r)$, the effective {\it molecular} dipole tensor,
357 $\mathbf{T}$ is obtained from two successive applications of the
358 gradient operator to the electrostatic kernel,
359 \begin{equation}
360 \mathbf{T}_{\alpha \beta}(r) = \nabla_\alpha \nabla_\beta \left(v(r)\right) = \delta_{\alpha \beta}
361 \left(\frac{1}{r} v^\prime(r) \right) + \frac{r_{\alpha}
362 r_{\beta}}{r^2} \left( v^{\prime \prime}(r) - \frac{1}{r}
363 v^{\prime}(r) \right)
364 \label{dipole-chargeTensor}
365 \end{equation}
366 where $v(r)$ may be either the bare kernel ($1/r$) or one of the
367 modified (Wolf or DSF) kernels. This tensor describes the effective
368 interaction between molecular dipoles ($\mathbf{D}$) in Gaussian
369 units as $-\mathbf{D} \cdot \mathbf{T} \cdot \mathbf{D}$.
370
371 When utilizing the new real-space methods for point dipoles, the
372 tensor is explicitly constructed,
373 \begin{equation}
374 \mathbf{T}_{\alpha \beta}(r) = \delta_{\alpha \beta} v_{21}(r) +
375 \frac{r_{\alpha} r_{\beta}}{r^2} v_{22}(r)
376 \label{dipole-diopleTensor}
377 \end{equation}
378 where the functions $v_{21}(r)$ and $v_{22}(r)$ depend on the level of
379 the approximation. Although the Taylor-shifted (TSF) and
380 gradient-shifted (GSF) models produce to the same $v(r)$ function for
381 point charges, they have distinct forms for the dipole-dipole
382 interactions.
383
384 Using constitutive relation, the polarization density $\textbf{P}(\textbf{r})$ is given by,
385 \begin{equation}
386 \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).
387 \label{constDipole}
388 \end{equation}
389 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}
390 \begin{equation}
391 \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}
392 \label{singular}
393 \end{equation}
394 Substituting equation (\ref{singular}) in the equation (\ref{constDipole}) we get,
395 \begin{equation}
396 \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).
397 \end{equation}
398 For both polarization and electric field homogeneous, this can be easily solved using Fourier transformation,
399 \begin{equation}
400 \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}).
401 \end{equation}
402 For homogeneous applied field Fourier component is non-zero only if $\kappa = 0$. Therefore,
403 \begin{equation}
404 \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}).
405 \label{fourierDipole}
406 \end{equation}
407 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,
408 \begin{equation}
409 \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}
410 \end{equation}
411 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,
412 \begin{equation}
413 \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}
414 \label{correctionFormula}
415 \end{equation}
416 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}.
417 \begin{table}
418 \caption{Expressions for the dipolar correction factor ($A$) for the real-space electrostatic methods in terms of the damping parameter
419 ($\alpha$) and the cutoff radius ($r_c$). The Ewald-Kornfeld result
420 derived in Refs. \onlinecite{Adams:1980rt,Adams:1981fr,Neumann83} is shown for comparison using the Ewald
421 convergence parameter ($\kappa$) and the real-space cutoff value ($r_c$). }
422 \label{tab:A}
423 {%
424 \begin{tabular}{l|c|c|c|}
425
426 Method & $A_\mathrm{charges}$ & $A_\mathrm{dipoles}$ \\
427 \hline
428 Spherical Cutoff (SC) & $\mathrm{erf}(r_c \alpha) - \frac{2 \alpha r_c}{\sqrt{\pi}} e^{-\alpha^2 r_c^2}$ & $\mathrm{erf}(r_c \alpha) - \frac{2 \alpha r_c}{\sqrt{\pi}} e^{-\alpha^2 r_c^2}$ \\
429 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} $\\
430 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} $ \\
431 Taylor-shifted (TSF) & 1 & 1 \\
432 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
433 \end{tabular}%
434 }
435 \end{table}
436 \subsection{Quadrupolar system}
437 In the presence of the field gradient $\partial_\alpha {E}_\beta $, a
438 non-vanishing quadrupolar polarization (quadrupole moment per unit
439 volume) $\bar{Q}_{\alpha\beta}$ will be induced in the entire volume
440 of a sample. The total field at any point $\vec{r}$ in the sample is
441 given by,
442 \begin{equation}
443 \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'
444 \label{gradMaxwell}
445 \end{equation}
446 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 $v(r)$.
447
448 \begin{equation}
449 \begin{split}
450 T_{\alpha\beta\gamma\delta}(r) &=\nabla_\alpha \nabla_\beta \nabla_\gamma \nabla_\delta\;v(r)
451 \\ &= \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)
452 \\ &+ \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)
453 \\ &+ 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),
454 \end{split}
455 \label{quadCharge}
456 \end{equation}
457 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,
458
459 \begin{equation}
460 \begin{split}
461 T_{\alpha\beta\gamma\delta}(r) &=\nabla_\alpha \nabla_\beta \;v_{\gamma\delta}(r)
462 \\ &= \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}
463 \\ &+ \delta_{\gamma\delta} r_\alpha r_\beta \left(\frac{v''_{21}(r)}{r^2}-\frac{v'_{21}(r)}{r^3} \right)
464 \\ &+\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)
465 \\ &+ 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),
466 \end{split}
467 \label{quadDip}
468 \end{equation}
469 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,
470 \begin{equation}
471 \begin{split}
472 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),
473 \end{split}
474 \label{quadRadial}
475 \end{equation}
476 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.
477 \begin{figure}
478 \includegraphics[width=3in]{QuadrupoleFigure}
479 \caption{With the real-space electrostatic methods, the effective
480 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.}
481 \label{fig:quadrupolarFluid}
482 \end{figure}
483 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
484 field gradient due to quadrupolar fluid vanishes at the singularity (see Appendix \ref{singularQuad}), equation (\ref{gradMaxwell}) can be written as,
485 \begin{equation}
486 \partial_\alpha E_\beta(\textbf{r}) = \partial_\alpha {E^o}_\beta(\textbf{r}) +
487 \frac{1}{4\pi \epsilon_o}\int\limits_{|\textbf{r}-\textbf{r}'|> 0 }
488 T_{\alpha\beta\gamma\delta}(|\textbf{r}-\textbf{r}'|)\;{Q}_{\gamma\delta}(\textbf{r}')\;
489 d^3r'.
490 \end{equation}
491 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}
492 \begin{equation}
493 \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',
494 \end{equation}
495 where $\Theta_{\alpha\beta} = 3Q_{\alpha\beta} - \delta_{\alpha\beta}Tr(Q)$
496 is the traceless quadrupole moment. The total quadrupolar polarization is written as,
497 \begin{equation}
498 {Q}_{\alpha\beta}(\textbf{r}) = \frac{1}{3}\delta_{\alpha\beta}\;Tr({Q})+\epsilon_o {\chi}^*_Q\;\partial_\alpha E_\beta(\textbf{r}),
499 \label{constQaud}
500 \end{equation}
501 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,
502 \begin{equation}
503 \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)
504 \end{equation}
505 For toroidal boundary conditions, both $\partial_\alpha E_\beta$ and
506 ${\Theta}_{\alpha\beta}$ are uniform over the entire space. After
507 performing a Fourier transform (see the Appendix in the Neumann's Paper \cite{Neumann83}) we get,
508 \begin{equation}
509 \frac{1}{3}{{\Theta}}_{\alpha\beta}({\kappa})=
510 \epsilon_o {\chi}^*_Q \;\left[{\partial_\alpha
511 {E^o}_\beta}({\kappa})+ \frac{1}{12\pi
512 \epsilon_o}\;{T}_{\alpha\beta\gamma\delta}({\kappa})\;
513 {{\Theta}}_{\gamma\delta}({\kappa})\right]
514 \end{equation}
515 Since the quadrupolar polarization is in the direction of the applied
516 field, we can write
517 ${{\Theta}}_{\gamma\delta}({\kappa}) =
518 {{\Theta}}_{\alpha\beta}({\kappa})$
519 and
520 ${T}_{\alpha\beta\gamma\delta}({\kappa}) =
521 {T}_{\alpha\beta\alpha\beta}({\kappa})$. Therefore we can consider each component of the interaction tensor as scalar and perform calculation.
522 \begin{equation}
523 \begin{split}
524 \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] \\
525 &= \epsilon_o {\chi}^*_Q\;\left(1-\frac{1}{4\pi} {\chi}^*_Q\;
526 {T}_{\alpha\beta\alpha\beta}({\kappa})\right)^{-1}
527 {\partial_\alpha E^o_\beta}({\kappa})
528 \end{split}
529 \label{fourierQuad}
530 \end{equation}
531 If the field gradient is homogeneous over the
532 entire volume, ${\partial_ \alpha E_\beta}({\kappa}) = 0 $ except at
533 $ {\kappa} = 0$, hence it is sufficient to know
534 ${T}_{\alpha\beta\alpha\beta}({\kappa})$ at $ {\kappa} =
535 0$. Therefore equation (\ref{fourierQuad}) can be written as,
536 \begin{equation}
537 \begin{split}
538 \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})
539 \end{split}
540 \label{fourierQuad2}
541 \end{equation}
542 where $ {T}_{\alpha\beta\alpha\beta}({0})$ can be evaluated as,
543 \begin{equation}
544 {T}_{\alpha\beta\alpha\beta}({0}) = \int {T}_{\alpha\beta\alpha\beta}\;(\textbf{r})d^3r
545 \label{realTensorQaud}
546 \end{equation}
547
548 In terms of traced quadrupole moment equation (\ref{fourierQuad2}) can be written as,
549 \begin{equation}
550 {{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
551 \label{tracedConstQuad}
552 \end{equation}
553 Comparing (\ref{tracedConstQuad}) and (\ref{flucQuad}) we get,
554 \begin{equation}
555 \begin{split}
556 &\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}, \\
557 &{\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}
558 \end{split}
559 \end{equation}
560 Finally the quadrupolar susceptibility cab be written in terms of quadrupolar correction factor ($A_{quad}$) as below,
561 \begin{equation}
562 {\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}
563 \label{eq:quadrupolarSusceptiblity}
564 \end{equation}
565 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}
566
567 \begin{equation}
568 \epsilon = 1 + \chi^*_Q\; G = 1 + G \; \alpha_Q\left(1 + \alpha_Q\; A_{quad}\right)^{-1}
569 \label{eq:dielectricFromQuadrupoles}
570 \end{equation}
571 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,
572 \begin{equation}
573 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}
574 \label{eq:geometricalFactor}
575 \end{equation}
576 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.
577 \begin{table}
578 \caption{Expressions for the quadrupolar correction factor ($A$) for the real-space electrostatic methods in terms of the damping parameter
579 ($\alpha$) and the cutoff radius ($r_c$). The dimension of the correction factor is $ length^{-2}$ in case of quadrupolar fluid.}
580 \label{tab:B}
581 \centering
582 \resizebox{\columnwidth}{!}{%
583
584 \begin{tabular}{l|c|c|c|c|}
585
586 Method & $A_\mathrm{charges}$ & $A_\mathrm{dipoles}$ &$A_\mathrm{quadrupoles}$ \\\hline
587 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}$ \\
588 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}$ \\
589 Gradient-shifted (GSF) & $- \frac{8 \alpha^5 {r_c}^3}{3\sqrt{\pi}} e^{-\alpha^2 r_c^2}$ & 0 & $-\frac{4{\alpha}^7{r_c}^5 }{9\sqrt{\pi}}e^{-\alpha^2 r_c^2}(-1+2\alpha ^2 r_c^2)$\\
590 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
591 \end{tabular}%
592 }
593 \end{table}
594 \section{Methodology}
595 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}.
596
597 \begin{table}
598 \caption{\label{Tab:C}}
599 \begin{tabularx}{\textwidth}{r|cc|YYccc|Yccc} \hline
600 & \multicolumn{2}{c|}{LJ parameters} &
601 \multicolumn{5}{c|}{Electrostatic moments} & & & & \\
602 Test system & $\sigma$& $\epsilon$ & $C$ & $D$ &
603 $Q_{xx}$ & $Q_{yy}$ & $Q_{zz}$ & mass & $I_{xx}$ & $I_{yy}$ &
604 $I_{zz}$ \\ \cline{6-8}\cline{10-12}
605 & (\AA) & (kcal/mol) & (e) & (debye) & \multicolumn{3}{c|}{(debye \AA)} & (amu) & \multicolumn{3}{c}{(amu
606 \AA\textsuperscript{2})} \\ \hline
607 Stockmayer fluid & 3.41 & 0.2381 & - & 1.4026 &-&-&-& 39.948 & 11.613 & 11.613 & 0.0 \\
608 Quadrupolar fluid & 2.985 & 0.265 & - & - & 0.0 & 0.0 &-2.139 & 18.0153 & 43.0565 & 43.0565 & 0.0 \\
609 \ce{q+} & 1.0 & 0.1 & +1 & - & - & - & - & 22.98 & - & - & - \\
610 \ce{q-} & 1.0 & 0.1 & -1 & - & - & - & - & 22.98 & - & - & - \\ \hline
611 \end{tabularx}
612 \end{table}
613
614 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.
615
616 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}).
617
618 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.
619
620 \subsection{Implementation}
621 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.
622
623 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}.
624 \subsection{Model systems}
625 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.
626
627 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.
628
629 \section{Results}
630
631 \section{Conclusion}
632
633 \newpage
634
635 \appendix
636 \section{Point-multipolar interactions with a spatially-varying electric field}
637
638 We can treat objects $a$, $b$, and $c$ containing embedded collections
639 of charges. When we define the primitive moments, we sum over that
640 collections of charges using a local coordinate system within each
641 object. The point charge, dipole, and quadrupole for object $a$ are
642 given by $C_a$, $\mathbf{D}_a$, and $\mathsf{Q}_a$, respectively.
643 These are the primitive multipoles which can be expressed as a
644 distribution of charges,
645 \begin{align}
646 C_a =&\sum_{k \, \text{in }a} q_k , \label{eq:charge} \\
647 D_{a\alpha} =&\sum_{k \, \text{in }a} q_k r_{k\alpha}, \label{eq:dipole}\\
648 Q_{a\alpha\beta} =& \frac{1}{2} \sum_{k \, \text{in } a} q_k
649 r_{k\alpha} r_{k\beta} . \label{eq:quadrupole}
650 \end{align}
651 Note that the definition of the primitive quadrupole here differs from
652 the standard traceless form, and contains an additional Taylor-series
653 based factor of $1/2$. In Paper 1, we derived the forces and torques
654 each object exerts on the others.
655
656 Here we must also consider an external electric field that varies in
657 space: $\mathbf E(\mathbf r)$. Each of the local charges $q_k$ in
658 object $a$ will then experience a slightly different field. This
659 electric field can be expanded in a Taylor series around the local
660 origin of each object. A different Taylor series expansion is carried
661 out for each object.
662
663 For a particular charge $q_k$, the electric field at that site's
664 position is given by:
665 \begin{equation}
666 E_\gamma + \nabla_\delta E_\gamma r_{k \delta}
667 + \frac {1}{2} \nabla_\delta \nabla_\varepsilon E_\gamma r_{k \delta}
668 r_{k \varepsilon} + ...
669 \end{equation}
670 Note that the electric field is always evaluated at the origin of the
671 objects, and treating each object using point multipoles simplifies
672 this greatly.
673
674 To find the force exerted on object $a$ by the electric field, one
675 takes the electric field expression, and multiplies it by $q_k$, and
676 then sum over all charges in $a$:
677
678 \begin{align}
679 F_\gamma &= \sum_{k \textrm{~in~} a} q_k \lbrace E_\gamma + \nabla_\delta E_\gamma r_{k \delta}
680 + \frac {1}{2} \nabla_\delta \nabla_\varepsilon E_\gamma r_{k \delta}
681 r_{k \varepsilon} + ... \rbrace \\
682 &= C_a E_\gamma + D_{a \delta} \nabla_\delta E_\gamma
683 + Q_{a \delta \varepsilon} \nabla_\delta \nabla_\varepsilon E_\gamma +
684 ...
685 \end{align}
686
687 Similarly, the torque exerted by the field on $a$ can be expressed as
688 \begin{align}
689 \tau_\alpha &= \sum_{k \textrm{~in~} a} (\mathbf r_k \times q_k \mathbf E)_\alpha \\
690 & = \sum_{k \textrm{~in~} a} \epsilon_{\alpha \beta \gamma} q_k
691 r_{k\beta} E_\gamma(\mathbf r_k) \\
692 & = \epsilon_{\alpha \beta \gamma} D_\beta E_\gamma
693 + 2 \epsilon_{\alpha \beta \gamma} Q_{\beta \delta} \nabla_\delta
694 E_\gamma + ...
695 \end{align}
696
697 The last term is essentially identical with form derived by Torres del
698 Castillo and M\'{e}ndez Garrido,\cite{Torres-del-Castillo:2006uo} although their derivation
699 utilized a traceless form of the quadrupole that is different than the
700 primitive definition in use here. We note that the Levi-Civita symbol
701 can be eliminated by utilizing the matrix cross product in an
702 identical form as in Ref. \onlinecite{Smith98}:
703 \begin{equation}
704 \left[\mathsf{A} \times \mathsf{B}\right]_\alpha = \sum_\beta
705 \left[\mathsf{A}_{\alpha+1,\beta} \mathsf{B}_{\alpha+2,\beta}
706 -\mathsf{A}_{\alpha+2,\beta} \mathsf{B}_{\alpha+1,\beta}
707 \right]
708 \label{eq:matrixCross}
709 \end{equation}
710 where $\alpha+1$ and $\alpha+2$ are regarded as cyclic permuations of
711 the matrix indices. In table \ref{tab:UFT} we give compact
712 expressions for how the multipole sites interact with an external
713 field that has exhibits spatial variations.
714
715 \begin{table}
716 \caption{Potential energy $(U)$, force $(\mathbf{F})$, and torque
717 $(\mathbf{\tau})$ expressions for a multipolar site embedded in an
718 electric field with spatial variations, $\mathbf{E}(\mathbf{r})$.
719 \label{tab:UFT}}
720 \begin{tabular}{r|ccc}
721 & Charge & Dipole & Quadrupole \\ \hline
722 $U$ & $C \phi(\mathbf{r})$ & $-\mathbf{D} \cdot \mathbf{E}(\mathbf{r})$ & $- \mathsf{Q}:\nabla \mathbf{E}(\mathbf{r})$ \\
723 $\mathbf{F}$ & $C \mathbf{E}(\mathbf{r})$ & $+\mathbf{D} \cdot \nabla \mathbf{E}(\mathbf{r})$ & $+\mathsf{Q} : \nabla\nabla\mathbf{E}(\mathbf{r})$ \\
724 $\mathbf{\tau}$ & & $\mathbf{D} \times \mathbf{E}(\mathbf{r})$ & $+2 \mathsf{Q} \times \nabla \mathbf{E}(\mathbf{r})$
725 \end{tabular}
726 \end{table}
727 \section{Gradient of the field due to quadrupolar polarization}
728 \label{singularQuad}
729 In this section, we will discuss the gradient of the field produced by
730 quadrupolar polarization. For this purpose, we consider a distribution
731 of charge ${\rho}(r)$ which gives rise to an electric field
732 $\vec{E}(r)$ and gradient of the field $\vec{\nabla} \vec{E}(r)$
733 throughout space. The total gradient of the electric field over volume
734 due to the all charges within the sphere of radius $R$ is given by
735 (cf. Jackson equation 4.14):
736 \begin{equation}
737 \int_{r<R} \vec{\nabla}\vec{E}\;d^3r = -\int_{r=R} R^2 \vec{E}\;\hat{n}\; d\Omega
738 \label{eq:8}
739 \end{equation}
740 where $d\Omega$ is the solid angle and $\hat{n}$ is the normal vector
741 of the surface of the sphere which is equal to
742 $sin[\theta]cos[\phi]\hat{x} + sin[\theta]sin[\phi]\hat{y} +
743 cos[\theta]\hat{z}$
744 in spherical coordinates. For the charge density ${\rho}(r')$, the
745 total gradient of the electric field can be written as (cf. Jackson
746 equation 4.16),
747 \begin{equation}
748 \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
749 \label{eq:9}
750 \end{equation}
751 The radial function in the equation (\ref{eq:9}) can be expressed in
752 terms of spherical harmonics as (cf. Jackson equation 3.70),
753 \begin{equation}
754 \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)
755 \label{eq:10}
756 \end{equation}
757 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,
758 \begin{equation}
759 \begin{split}
760 \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 \\
761 &= -\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
762 '
763 \end{split}
764 \label{eq:11}
765 \end{equation}
766 The gradient of the product of radial function and spherical harmonics
767 is given by (cf. Arfken, p.811 eq. 16.94):
768 \begin{equation}
769 \begin{split}
770 \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
771 {\partial}{\partial r}+\frac{l}{r} \right]f(r)\; Y_{l, l-1, m}(\theta, \phi).
772 \end{split}
773 \label{eq:12}
774 \end{equation}
775 Using equation (\ref{eq:12}) we get,
776 \begin{equation}
777 \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}},
778 \label{eq:13}
779 \end{equation}
780 where $ Y_{l,l+1,m}(\theta, \phi)$ is the vector spherical harmonics
781 which can be expressed in terms of spherical harmonics as shown in
782 below (cf. Arfkan p.811),
783 \begin{equation}
784 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},
785 \label{eq:14}
786 \end{equation}
787 where $C(l+1,1,l|m_1,m_2,m)$ is a Clebsch-Gordan coefficient and
788 $\hat{e}_{m_2}$ is a spherical tensor of rank 1 which can be expressed
789 in terms of Cartesian coordinates,
790 \begin{equation}
791 {\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}}
792 \label{eq:15}
793 \end{equation}
794 The normal vector $\hat{n} $ can be expressed in terms of spherical tensor of rank 1 as shown in below,
795 \begin{equation}
796 \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)
797 \label{eq:16}
798 \end{equation}
799 The surface integral of the product of $\hat{n}$ and
800 ${Y_{l+1}}^{m_1}(\theta, \phi)$ gives,
801 \begin{equation}
802 \begin{split}
803 \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 \\
804 &= \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 \\
805 &= \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),
806 \end{split}
807 \label{eq:17}
808 \end{equation}
809 where ${Y_{l}}^{-m} = (-1)^m\;{{Y_{l}}^{m}}^* $ and
810 $ \int {{Y_{l}}^{m}}^*\;{Y_{l'}}^{m'}\;d\Omega =
811 \delta_{ll'}\delta_{mm'} $.
812 Non-vanishing values of equation \ref{eq:17} require $l = 0$,
813 therefore the value of $ m = 0 $. Since the values of $ m_1$ are -1,
814 1, and 0 then $m_2$ takes the values 1, -1, and 0, respectively
815 provided that $m = m_1 + m_2$. Equation \ref{eq:11} can therefore be
816 modified,
817 \begin{equation}
818 \begin{split}
819 \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(
820 1, 1, 0|0,0,0)\;{\hat{e}_{0}}{\hat{e}_{0}} ]\; d^3r'.
821 \end{split}
822 \label{eq:18}
823 \end{equation}
824 After substituting ${Y^*}_{00} = \frac{1}{\sqrt{4\pi}} $ and using the
825 values of the Clebsch-Gorden coefficients: $ C(1, 1, 0|-1,1,0) =
826 \frac{1}{\sqrt{3}}, \; C(1, 1, 0|-1,1,0)= \frac{1}{\sqrt{3}}$ and $
827 C(1, 1, 0|0,0,0) = -\frac{1}{\sqrt{3}}$ in equation \ref{eq:18} we
828 obtain,
829 \begin{equation}
830 \begin{split}
831 \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)\\
832 &= - \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).
833 \end{split}
834 \label{eq:19}
835 \end{equation}
836 Equation (\ref{eq:19}) gives the total gradient of the field over a
837 sphere due to the distribution of the charges. For quadrupolar fluids
838 the total charge within a sphere is zero, therefore
839 $ \int_{r<R} \vec{\nabla}\vec{E}\;d^3r = 0 $. Hence the quadrupolar
840 polarization produces zero net gradient of the field inside the
841 sphere.
842
843 \section{Applied field or field gradient}
844 \label{Ap:fieldOrGradient}
845
846 To satisfy the condition $ \nabla . E = 0 $, within the box of molecules we have taken electrostatic potential in the following form
847 \begin{equation}
848 \begin{split}
849 \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. \\
850 & \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),
851 \end{split}
852 \label{eq:appliedPotential}
853 \end{equation}
854 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,
855 \[\bf{E}
856 =\frac{g_o}{2} \left(\begin{array}{ccc}
857 2(a_1\; b_1 - \frac{cos\psi}{3})\;x \;+ (a_1\; b_2 \;+ a_2\; b_1)\;y + (a_1\; b_3 \;+ a_3\; b_1)\;z \\
858 (a_2\; b_1 \;+ a_1\; b_2)\;x + 2(a_2\; b_2 \;- \frac{cos\psi}{3})\;y + (a_2\; b_3 \;+ a_3\; b_3)\;z \\
859 (a_3\; b_1 \;+ a_3\; b_2)\;x + (a_3\; b_2 \;+ a_2\; b_3)y + 2(a_3\; b_3 \;- \frac{cos\psi}{3})\;z
860 \end{array} \right).\]
861 The gradient of the applied field derived from the potential can be written in the following form,
862 \[\nabla\bf{E}
863 = \frac{g_o}{2}\left(\begin{array}{ccc}
864 2(a_1\; b_1 - \frac{cos\psi}{3}) & (a_1\; b_2 \;+ a_2\; b_1) & (a_1\; b_3 \;+ a_3\; b_1)\;z \\
865 (a_2\; b_1 \;+ a_1\; b_2) & 2(a_2\; b_2 \;- \frac{cos\psi}{3}) & (a_2\; b_3 \;+ a_3\; b_3)\;z \\
866 (a_3\; b_1 \;+ a_3\; b_2) & (a_3\; b_2 \;+ a_2\; b_3) & 2(a_3\; b_3 \;- \frac{cos\psi}{3})\;z
867 \end{array} \right).\]
868 \newpage
869
870 \bibliography{multipole}
871
872 \end{document}