ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/multipole/dielectric_new.tex
Revision: 4389
Committed: Tue Nov 3 16:30:49 2015 UTC (8 years, 7 months ago) by mlamichh
Content type: application/x-tex
File size: 61064 byte(s)
Log Message:
added Methodology

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 correcting electrostatic interactions in computer
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 One of the most difficult tests of any new electrostatic method is the fidelity with which that method can reproduce the bulk-phase polarizability or equivalently, the dielectric properties of a fluid. Since dielectric properties are macroscopic properties, all interactions between molecules in an entire system are significantly important. But it is computationally infeasible to consider every interactions between molecules in the macroscopically large system. Therefore small molecular system with periodic boundary condition and finite cutoff region of interactions is usually considered in computer simulations. While calculating dielectric properties, the formula should be modified in such a way so that it can accommodate behaviour of electrostatic neutrality and smoothness of energy, force and torque at the cutoff radius. Previously many studies have been conducted to calculate dipolar and quadrupolar dielectric properties using computer simulations. \cite{Kirkwood39, Onsagar36,LoganI81, LoganII82, LoganIII82} But these methods do not specifically take account of the cutoff behavior common in real-space electrosatic methods. In 1983 Neumann proposed a general formula for evaluating dielectric properties for dipolar fluid using real-space cutoff methods. \cite{Neumann83} In the same year Steinhauser and Neumann used this formula to evaluate the correct dielectric constant for the Stockmayer fluid using two different methods: Ewald-Kornfield (EK) and reaction field (RF) methods. \cite{Neumann-Steinhauser83} This formula contains a correction factor which is equal to $\frac{3}{4 \pi} $ times volume integral of the dipole-dipole interactions for a given electrostatic cutoff method (See equation \ref{dipole-diopleTensor}).\cite{Neumann83} Similarly Zahn \textit{et al.}\cite{Zahn02} also evaluated correction factor for dipole-dipole interaction using damped shifted charge-charge kernel (see equation \ref{dipole-chargeTensor}). This later generalized by Izvekove \textit{et al.}, which is equal to $\frac{3}{4 \pi} $. \cite{Izvekov:2008wo} When the correction factor is equal to $\frac{3}{4 \pi} $, the expression for the dielectric constant reduces to widely-used \textit{conducting boundary} formula (see equation (\ref{correctionFormula})). Many studies have also been conducted to understand solvation theory using dielectric properties of quadrupolar fluid.\cite{JeonI03, JeonII03, Chitanvis96}. But these studies do not use correction factor straight forwardly to evaluate correct dielectric properties for quadrupolar fluid.
77
78 In this paper we are proposing general consecutive formulas for calculating the dielectric properties for quadrupolar fluid. Furthermore we have also evaluated the correction factor for SP, GSF, and TSF method for both dipolar and quadrupolar fluid considering charge-charge, dipole-dipole or quadrupole-quadrupole interactions. The relation between quadrupolar susceptibility and dielectric constant is not straight forward for quadrupolar fluid as in the dipolar case. The dielectric constant depends on the geometry of the external field perturbation.\cite{Ernst92} We have also calculated the geometrical factor for two ions immersed quadrupolar system to evaluate dielectric constant from the quadrupolar susceptibility. We have used three different methods: i) external field perturbation, ii) fluctuation formula, and iii) the potential of mean force, to study dielectric properties of the dipolar and quadrupolar system. In the external field perturbation, the net polarization of the system is observed as a linear response of the applied field perturbation, where proportionality constant is determined by the electrostatic interaction between the electrostatic multipoles at a given temperature. The fluctuation formula observes the time average fluctuation of the multipolar moment as a function of temperature. The average fluctuation value of the system is determined by the multipole-multipole interactions between molecules at a given temperature. Since the expression of the electrostatic interaction energy, force, and torque in the real space electrostatic methods are different from their original definition, both fluctuation and external field perturbation formula should also be modified accordingly. The potential of mean force method calculates dielectric constant from the potential energy between ions before and after dielectric material is introduced. All of these different methods for calculating dielectric properties will be discussed in detail in the following sections: \ref{subsec:perturbation}, \ref{subsec:fluctuation}, and \ref{sec:PMF}.
79
80 \section{Boltzmann average for orientational polarization}
81 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}.
82
83 \subsection{Dipole}
84 \label{subsec:boltzAverage-Dipole}
85 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}
86
87 \begin{equation}
88 H = H_o - \bf{p_o} .\bf{E},
89 \end{equation}
90 where $H_o$ is a function of the internal coordinates of the molecule. Now Boltzmann average of the dipole moment is given by,
91 \begin{equation}
92 \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}}},
93 \end{equation}
94 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,
95
96 \begin{equation}
97 \braket{p_{mol}} \approx \frac{1}{3}\frac{{p_o}^2}{k_B T}E,
98 \end{equation}
99 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.
100
101
102 \subsection{Quadrupole}
103 \label{subsec:boltzAverage-Quad}
104 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}
105 \begin{equation}
106 \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}}},
107 \label{boltzQuad}
108 \end{equation}
109 where $\int d\Omega = \int_0^{2\pi} \int_0^\pi \int_0^{2\pi}
110 sin\theta\; d\theta\ d\phi\ d\psi$ is the integration over Euler
111 angles, $ H = H_o -q_{\mu\nu}\;\partial_\nu E_\mu $ is the energy of
112 a quadrupole in the gradient of the
113 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,
114 \begin{equation}
115 \begin{split}
116 &q_{\alpha\beta} = \eta_{\alpha\alpha'}\;\eta_{\beta\beta'}\;{q}^* _{\alpha'\beta'} \\
117 &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.
118 \end{split}
119 \label{energyQuad}
120 \end{equation}
121 Here the starred tensors are the components in the body fixed
122 frame. Substituting equation (\ref{energyQuad}) in the equation (\ref{boltzQuad})
123 and taking linear terms in the expansion we get,
124 \begin{equation}
125 \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)},
126 \end{equation}
127 where $\eta_{\alpha\alpha'}$ is the inverse of the rotation matrix that transforms
128 the body fixed co-ordinates to the space co-ordinates,
129 \[\eta_{\alpha\alpha'}
130 = \left(\begin{array}{ccc}
131 cos\phi\; cos\psi - cos\theta\; sin\phi\; sin\psi & -cos\theta\; cos\psi\; sin\phi - cos\phi\; sin\psi & sin\theta\; sin\phi \\
132 cos\psi\; sin\phi + cos\theta\; cos\phi \; sin\psi & cos\theta\; cos\phi\; cos\psi - sin\phi\; sin\psi & -cos\phi\; sin\theta \\
133 sin\theta\; sin\psi & -cos\psi\; sin\theta & cos\theta
134 \end{array} \right).\]
135 Integration of 1st and 2nd terms in the denominator gives $8 \pi^2$
136 and $8 \pi^2 /3\;\vec{\nabla}.\vec{E}\; Tr(q^*) $ respectively. The
137 second term vanishes for charge free space
138 (i.e. $\vec{\nabla}.\vec{E} \; = \; 0)$. Similarly integration of the
139 1st term in the numerator produces
140 $8 \pi^2 /3\; Tr(q^*)\delta_{\alpha\beta}$ and the 2nd term produces
141 $8 \pi^2 /15k_B T (3{q}^*_{\alpha'\beta'}{q}^*_{\beta'\alpha'} -
142 {q}^*_{\alpha'\alpha'}{q}^*_{\beta'\beta'})\partial_\alpha E_\beta$,
143 if $\vec{\nabla}.\vec{E} \; = \; 0$,
144 $ \partial_\alpha E_\beta = \partial_\beta E_\alpha$ and
145 ${q}^*_{\alpha'\beta'}= {q}^*_{\beta'\alpha'}$. Therefore the
146 Boltzmann average of a quadrupole moment can be written as,
147
148 \begin{equation}
149 \braket{q_{\alpha\beta}}\; = \; \frac{1}{3} Tr(q^*)\;\delta_{\alpha\beta} + \frac{{\bar{q_o}}^2}{15k_BT}\;\partial_\alpha E_\beta,
150 \end{equation}
151 where $ \alpha_q = \frac{{\bar{q_o}}^2}{15k_BT} $ is a molecular quadrupolarizablity and ${\bar{q_o}}^2=
152 3{q}^*_{\alpha'\beta'}{q}^*_{\beta'\alpha'}-{q}^*_{\alpha'\alpha'}{q}^*_{\beta'\beta'}$ is a square of the net quadrupole moment of a molecule.
153
154 \section{Macroscopic Polarizability}
155 \label{sec:MacPolarizablity}
156
157 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.
158
159 \subsection{External field perturbation}
160 \label{subsec:perturbation}
161 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,
162 \begin{equation}
163 \textbf{P} = \epsilon_o \alpha_{D}\; \textbf{E}^o.
164 \label{pertDipole}
165 \end{equation}
166 The constant $\alpha_D$ is a macroscopic polarizability, which is a property of the dipolar fluid in a given density and temperature.
167
168 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,
169 \begin{equation}
170 \begin{split}
171 & {Q}_{\alpha\beta} = \frac{1}{3}\; Tr({Q})\; \delta_{\alpha\beta} + \epsilon_o\; \alpha_Q \; \partial_{\alpha} E^o_{\beta}
172 \\ & or \\
173 & \frac{1}{3}\;\Theta_{\alpha\beta} = \epsilon_o\; \alpha_Q \; \partial_{\alpha} E^o_{\beta}
174 \end{split}
175 \label{pertQuad}
176 \end{equation}
177 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.
178
179
180 \subsection{Fluctuation formula}
181 \label{subsec:fluctuation}
182 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}
183 \begin{equation}
184 \braket{\bf{P}} = \epsilon_o \frac{\braket{\bf{M}^2}-{\braket{\bf{M}}}^2}{3 \epsilon_o V k_B T}\bf{E}^o
185 \label{flucDipole}
186 \end{equation}
187 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,
188 \begin{equation}
189 \alpha_D = \frac{\braket{\bf{M}^2}-{\braket{\bf{M}}}^2}{3 \epsilon_o V k_B T}
190 \end{equation}
191 This is a macroscopic property of dipolar material which is true even if applied field $ \textbf{E}^o \rightarrow 0 $.
192
193 Analogous formula can also be written for a system with quadrupolar molecules,
194 \begin{equation}
195 \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}
196 \label{flucQuad}
197 \end{equation}
198 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,
199 \begin{equation}
200 \alpha_Q = \frac{\braket{\textbf{Q}^2}-{\braket{\textbf{Q}}}^2}{15 \epsilon_o V k_B T}
201 \label{propConstQuad}
202 \end{equation}
203
204
205 \section{Potential of mean force}
206 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,
207 \begin{equation}
208 w(r) = \int_{\inf}^{r}\braket{\frac{\partial f}{\partial r'}}dr',
209 \end{equation}
210 where $\braket{\partial f/\partial r'}$ is the mean constraint force.
211 Since the ions have a protecting Lennard-Jones (LJ) potential,
212 \begin{equation}
213 w(r) = w_{LJ}(r) + \frac{q_iq_j}{4\pi \epsilon_o \epsilon(r)}U_{method}(r).
214 \label{eq:pmf}
215 \end{equation}
216 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.
217
218 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.
219 \label{sec:PMF}
220
221 \section{Correction factor}
222 \label{sec:corrFactor}
223 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.
224 \subsection{Dipolar system}
225 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}
226 \begin{equation}
227 \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}')}.
228 \end{equation}
229
230 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}).
231 \begin{figure}
232 \includegraphics[width=3in]{DielectricFigure}
233 \caption{With the real-space electrostatic methods, the effective
234 dipole tensor, $\mathbf{T}$, governing interactions between
235 molecular dipoles is not the same for charge-charge interactions as
236 for point dipoles.}
237 \label{fig:stockmayer}
238 \end{figure}
239 In the case where point charges are interacting via an electrostatic
240 kernel, $v(r)$, the effective {\it molecular} dipole tensor,
241 $\mathbf{T}$ is obtained from two successive applications of the
242 gradient operator to the electrostatic kernel,
243 \begin{equation}
244 \mathbf{T}_{\alpha \beta}(r) = \nabla_\alpha \nabla_\beta \left(v(r)\right) = \delta_{\alpha \beta}
245 \left(\frac{1}{r} v^\prime(r) \right) + \frac{r_{\alpha}
246 r_{\beta}}{r^2} \left( v^{\prime \prime}(r) - \frac{1}{r}
247 v^{\prime}(r) \right)
248 \label{dipole-chargeTensor}
249 \end{equation}
250 where $v(r)$ may be either the bare kernel ($1/r$) or one of the
251 modified (Wolf or DSF) kernels. This tensor describes the effective
252 interaction between molecular dipoles ($\mathbf{D}$) in Gaussian
253 units as $-\mathbf{D} \cdot \mathbf{T} \cdot \mathbf{D}$.
254
255 When utilizing the new real-space methods for point dipoles, the
256 tensor is explicitly constructed,
257 \begin{equation}
258 \mathbf{T}_{\alpha \beta}(r) = \delta_{\alpha \beta} v_{21}(r) +
259 \frac{r_{\alpha} r_{\beta}}{r^2} v_{22}(r)
260 \label{dipole-diopleTensor}
261 \end{equation}
262 where the functions $v_{21}(r)$ and $v_{22}(r)$ depend on the level of
263 the approximation. Although the Taylor-shifted (TSF) and
264 gradient-shifted (GSF) models produce to the same $v(r)$ function for
265 point charges, they have distinct forms for the dipole-dipole
266 interactions.
267
268 Using constitutive relation, the polarization density $\textbf{P}(\textbf{r})$ is given by,
269 \begin{equation}
270 \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).
271 \label{constDipole}
272 \end{equation}
273 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}
274 \begin{equation}
275 \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}
276 \label{singular}
277 \end{equation}
278 Substituting equation (\ref{singular}) in the equation (\ref{constDipole}) we get,
279 \begin{equation}
280 \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).
281 \end{equation}
282 For both polarization and electric field homogeneous, this can be easily solved using Fourier transformation,
283 \begin{equation}
284 \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}).
285 \end{equation}
286 For homogeneous applied field Fourier component is non-zero only if $\kappa = 0$. Therefore,
287 \begin{equation}
288 \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}).
289 \label{fourierDipole}
290 \end{equation}
291 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,
292 \begin{equation}
293 \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}
294 \end{equation}
295 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,
296 \begin{equation}
297 \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}
298 \label{correctionFormula}
299 \end{equation}
300 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}.
301 \begin{table}
302 \caption{Expressions for the dipolar correction factor ($A$) for the real-space electrostatic methods in terms of the damping parameter
303 ($\alpha$) and the cutoff radius ($r_c$). The Ewald-Kornfeld result
304 derived in Refs. \onlinecite{Adams:1980rt,Adams:1981fr,Neumann83} is shown for comparison using the Ewald
305 convergence parameter ($\kappa$) and the real-space cutoff value ($r_c$). }
306 \label{tab:A}
307 {%
308 \begin{tabular}{l|c|c|c|}
309
310 Method & $A_\mathrm{charges}$ & $A_\mathrm{dipoles}$ \\
311 \hline
312 Spherical Cutoff (SC) & $\mathrm{erf}(r_c \alpha) - \frac{2 \alpha r_c}{\sqrt{\pi}} e^{-\alpha^2 r_c^2}$ & $\mathrm{erf}(r_c \alpha) - \frac{2 \alpha r_c}{\sqrt{\pi}} e^{-\alpha^2 r_c^2}$ \\
313 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} $\\
314 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} $ \\
315 Taylor-shifted (TSF) & 1 & 1 \\
316 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
317 \end{tabular}%
318 }
319 \end{table}
320 \subsection{Quadrupolar system}
321 In the presence of the field gradient $\partial_\alpha {E}_\beta $, a
322 non-vanishing quadrupolar polarization (quadrupole moment per unit
323 volume) $\bar{Q}_{\alpha\beta}$ will be induced in the entire volume
324 of a sample. The total field at any point $\vec{r}$ in the sample is
325 given by,
326 \begin{equation}
327 \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'
328 \label{gradMaxwell}
329 \end{equation}
330 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 $v(r)$.
331
332 \begin{equation}
333 \begin{split}
334 T_{\alpha\beta\gamma\delta}(r) &=\nabla_\alpha \nabla_\beta \nabla_\gamma \nabla_\delta\;v(r)
335 \\ &= \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)
336 \\ &+ \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)
337 \\ &+ 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),
338 \end{split}
339 \label{quadCharge}
340 \end{equation}
341 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,
342
343 \begin{equation}
344 \begin{split}
345 T_{\alpha\beta\gamma\delta}(r) &=\nabla_\alpha \nabla_\beta \;v_{\gamma\delta}(r)
346 \\ &= \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}
347 \\ &+ \delta_{\gamma\delta} r_\alpha r_\beta \left(\frac{v''_{21}(r)}{r^2}-\frac{v'_{21}(r)}{r^3} \right)
348 \\ &+\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)
349 \\ &+ 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),
350 \end{split}
351 \label{quadDip}
352 \end{equation}
353 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,
354 \begin{equation}
355 \begin{split}
356 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),
357 \end{split}
358 \label{quadRadial}
359 \end{equation}
360 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.
361 \begin{figure}
362 \includegraphics[width=3in]{QuadrupoleFigure}
363 \caption{With the real-space electrostatic methods, the effective
364 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.}
365 \label{fig:quadrupolarFluid}
366 \end{figure}
367 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
368 field gradient due to quadrupolar fluid vanishes at the singularity (see Appendix \ref{singularQuad}), equation (\ref{gradMaxwell}) can be written as,
369 \begin{equation}
370 \partial_\alpha E_\beta(\textbf{r}) = \partial_\alpha {E^o}_\beta(\textbf{r}) +
371 \frac{1}{4\pi \epsilon_o}\int\limits_{|\textbf{r}-\textbf{r}'|> 0 }
372 T_{\alpha\beta\gamma\delta}(|\textbf{r}-\textbf{r}'|)\;{Q}_{\gamma\delta}(\textbf{r}')\;
373 d^3r'.
374 \end{equation}
375 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}
376 \begin{equation}
377 \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',
378 \end{equation}
379 where $\Theta_{\alpha\beta} = 3Q_{\alpha\beta} - \delta_{\alpha\beta}Tr(Q)$
380 is the traceless quadrupole moment. The total quadrupolar polarization is written as,
381 \begin{equation}
382 {Q}_{\alpha\beta}(\textbf{r}) = \frac{1}{3}\delta_{\alpha\beta}\;Tr({Q})+\epsilon_o {\chi}^*_Q\;\partial_\alpha E_\beta(\textbf{r}),
383 \label{constQaud}
384 \end{equation}
385 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,
386 \begin{equation}
387 \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)
388 \end{equation}
389 For toroidal boundary conditions, both $\partial_\alpha E_\beta$ and
390 ${\Theta}_{\alpha\beta}$ are uniform over the entire space. After
391 performing a Fourier transform (see the Appendix in the Neumann's Paper \cite{Neumann83}) we get,
392 \begin{equation}
393 \frac{1}{3}{{\Theta}}_{\alpha\beta}({\kappa})=
394 \epsilon_o {\chi}^*_Q \;\left[{\partial_\alpha
395 {E^o}_\beta}({\kappa})+ \frac{1}{12\pi
396 \epsilon_o}\;{T}_{\alpha\beta\gamma\delta}({\kappa})\;
397 {{\Theta}}_{\gamma\delta}({\kappa})\right]
398 \end{equation}
399 Since the quadrupolar polarization is in the direction of the applied
400 field, we can write
401 ${{\Theta}}_{\gamma\delta}({\kappa}) =
402 {{\Theta}}_{\alpha\beta}({\kappa})$
403 and
404 ${T}_{\alpha\beta\gamma\delta}({\kappa}) =
405 {T}_{\alpha\beta\alpha\beta}({\kappa})$. Therefore we can consider each component of the interaction tensor as scalar and perform calculation.
406 \begin{equation}
407 \begin{split}
408 \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] \\
409 &= \epsilon_o {\chi}^*_Q\;\left(1-\frac{1}{4\pi} {\chi}^*_Q\;
410 {T}_{\alpha\beta\alpha\beta}({\kappa})\right)^{-1}
411 {\partial_\alpha E^o_\beta}({\kappa})
412 \end{split}
413 \label{fourierQuad}
414 \end{equation}
415 If the field gradient is homogeneous over the
416 entire volume, ${\partial_ \alpha E_\beta}({\kappa}) = 0 $ except at
417 $ {\kappa} = 0$, hence it is sufficient to know
418 ${T}_{\alpha\beta\alpha\beta}({\kappa})$ at $ {\kappa} =
419 0$. Therefore equation (\ref{fourierQuad}) can be written as,
420 \begin{equation}
421 \begin{split}
422 \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})
423 \end{split}
424 \label{fourierQuad2}
425 \end{equation}
426 where $ {T}_{\alpha\beta\alpha\beta}({0})$ can be evaluated as,
427 \begin{equation}
428 {T}_{\alpha\beta\alpha\beta}({0}) = \int {T}_{\alpha\beta\alpha\beta}\;(\textbf{r})d^3r
429 \label{realTensorQaud}
430 \end{equation}
431
432 In terms of traced quadrupole moment equation (\ref{fourierQuad2}) can be written as,
433 \begin{equation}
434 {{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
435 \label{tracedConstQuad}
436 \end{equation}
437 Comparing (\ref{tracedConstQuad}) and (\ref{flucQuad}) we get,
438 \begin{equation}
439 \begin{split}
440 &\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}, \\
441 &{\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}
442 \end{split}
443 \end{equation}
444 Finally the quadrupolar susceptibility cab be written in terms of quadrupolar correction factor ($A_{quad}$) as below,
445 \begin{equation}
446 {\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}
447 \label{eq:quadrupolarSusceptiblity}
448 \end{equation}
449 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}
450
451 \begin{equation}
452 \epsilon = 1 + \chi^*_Q\; G = 1 + G \; \alpha_Q\left(1 + \alpha_Q\; A_{quad}\right)^{-1}
453 \label{eq:dielectricFromQuadrupoles}
454 \end{equation}
455 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,
456 \begin{equation}
457 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}
458 \label{eq:geometricalFactor}
459 \end{equation}
460 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.
461 \begin{table}
462 \caption{Expressions for the quadrupolar correction factor ($A$) for the real-space electrostatic methods in terms of the damping parameter
463 ($\alpha$) and the cutoff radius ($r_c$). The dimension of the correction factor is $ length^{-2}$ in case of quadrupolar fluid.}
464 \label{tab:B}
465 \centering
466 \resizebox{\columnwidth}{!}{%
467
468 \begin{tabular}{l|c|c|c|c|}
469
470 Method & $A_\mathrm{charges}$ & $A_\mathrm{dipoles}$ &$A_\mathrm{quadrupoles}$ \\\hline
471 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}$ \\
472 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}$ \\
473 Gradient-shifted (GSF) & $- \frac{8 \alpha^5 {r_c}^3}{3\sqrt{\pi}} e^{-\alpha^2 r_c^2}$ & 0 & $-\frac{4{\alpha}^7{r_c}^5 }{9\sqrt{\pi}}e^{-\alpha^2 r_c^2}(-1+2\alpha ^2 r_c^2)$\\
474 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
475 \end{tabular}%
476 }
477 \end{table}
478 \section{Methodology}
479 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}.
480
481 \begin{table}
482 \caption{\label{Tab:C}}
483 \begin{tabularx}{\textwidth}{r|cc|YYccc|Yccc} \hline
484 & \multicolumn{2}{c|}{LJ parameters} &
485 \multicolumn{5}{c|}{Electrostatic moments} & & & & \\
486 Test system & $\sigma$& $\epsilon$ & $C$ & $D$ &
487 $Q_{xx}$ & $Q_{yy}$ & $Q_{zz}$ & mass & $I_{xx}$ & $I_{yy}$ &
488 $I_{zz}$ \\ \cline{6-8}\cline{10-12}
489 & (\AA) & (kcal/mol) & (e) & (debye) & \multicolumn{3}{c|}{(debye \AA)} & (amu) & \multicolumn{3}{c}{(amu
490 \AA\textsuperscript{2})} \\ \hline
491 Stockmayer fluid & 3.41 & 0.2381 & - & 1.4026 &-&-&-& 39.948 & 11.613 & 11.613 & 0.0 \\
492 Quadrupolar fluid & 2.985 & 0.265 & - & - & 0.0 & 0.0 &-2.139 & 18.0153 & 43.0565 & 43.0565 & 0.0 \\
493 \ce{q+} & 1.0 & 0.1 & +1 & - & - & - & - & 22.98 & - & - & - \\
494 \ce{q-} & 1.0 & 0.1 & -1 & - & - & - & - & 22.98 & - & - & - \\ \hline
495 \end{tabularx}
496 \end{table}
497
498 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.
499
500 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}).
501
502 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.
503
504 \subsection{Implementation}
505 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.
506
507 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}.
508 \subsection{Model systems}
509 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.
510
511 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.
512
513 \section{Results}
514
515 \section{Conclusion}
516
517 \newpage
518
519 \appendix
520 \section{Point-multipolar interactions with a spatially-varying electric field}
521
522 We can treat objects $a$, $b$, and $c$ containing embedded collections
523 of charges. When we define the primitive moments, we sum over that
524 collections of charges using a local coordinate system within each
525 object. The point charge, dipole, and quadrupole for object $a$ are
526 given by $C_a$, $\mathbf{D}_a$, and $\mathsf{Q}_a$, respectively.
527 These are the primitive multipoles which can be expressed as a
528 distribution of charges,
529 \begin{align}
530 C_a =&\sum_{k \, \text{in }a} q_k , \label{eq:charge} \\
531 D_{a\alpha} =&\sum_{k \, \text{in }a} q_k r_{k\alpha}, \label{eq:dipole}\\
532 Q_{a\alpha\beta} =& \frac{1}{2} \sum_{k \, \text{in } a} q_k
533 r_{k\alpha} r_{k\beta} . \label{eq:quadrupole}
534 \end{align}
535 Note that the definition of the primitive quadrupole here differs from
536 the standard traceless form, and contains an additional Taylor-series
537 based factor of $1/2$. In Paper 1, we derived the forces and torques
538 each object exerts on the others.
539
540 Here we must also consider an external electric field that varies in
541 space: $\mathbf E(\mathbf r)$. Each of the local charges $q_k$ in
542 object $a$ will then experience a slightly different field. This
543 electric field can be expanded in a Taylor series around the local
544 origin of each object. A different Taylor series expansion is carried
545 out for each object.
546
547 For a particular charge $q_k$, the electric field at that site's
548 position is given by:
549 \begin{equation}
550 E_\gamma + \nabla_\delta E_\gamma r_{k \delta}
551 + \frac {1}{2} \nabla_\delta \nabla_\varepsilon E_\gamma r_{k \delta}
552 r_{k \varepsilon} + ...
553 \end{equation}
554 Note that the electric field is always evaluated at the origin of the
555 objects, and treating each object using point multipoles simplifies
556 this greatly.
557
558 To find the force exerted on object $a$ by the electric field, one
559 takes the electric field expression, and multiplies it by $q_k$, and
560 then sum over all charges in $a$:
561
562 \begin{align}
563 F_\gamma &= \sum_{k \textrm{~in~} a} q_k \lbrace E_\gamma + \nabla_\delta E_\gamma r_{k \delta}
564 + \frac {1}{2} \nabla_\delta \nabla_\varepsilon E_\gamma r_{k \delta}
565 r_{k \varepsilon} + ... \rbrace \\
566 &= C_a E_\gamma + D_{a \delta} \nabla_\delta E_\gamma
567 + Q_{a \delta \varepsilon} \nabla_\delta \nabla_\varepsilon E_\gamma +
568 ...
569 \end{align}
570
571 Similarly, the torque exerted by the field on $a$ can be expressed as
572 \begin{align}
573 \tau_\alpha &= \sum_{k \textrm{~in~} a} (\mathbf r_k \times q_k \mathbf E)_\alpha \\
574 & = \sum_{k \textrm{~in~} a} \epsilon_{\alpha \beta \gamma} q_k
575 r_{k\beta} E_\gamma(\mathbf r_k) \\
576 & = \epsilon_{\alpha \beta \gamma} D_\beta E_\gamma
577 + 2 \epsilon_{\alpha \beta \gamma} Q_{\beta \delta} \nabla_\delta
578 E_\gamma + ...
579 \end{align}
580
581 The last term is essentially identical with form derived by Torres del
582 Castillo and M\'{e}ndez Garrido,\cite{Torres-del-Castillo:2006uo} although their derivation
583 utilized a traceless form of the quadrupole that is different than the
584 primitive definition in use here. We note that the Levi-Civita symbol
585 can be eliminated by utilizing the matrix cross product in an
586 identical form as in Ref. \onlinecite{Smith98}:
587 \begin{equation}
588 \left[\mathsf{A} \times \mathsf{B}\right]_\alpha = \sum_\beta
589 \left[\mathsf{A}_{\alpha+1,\beta} \mathsf{B}_{\alpha+2,\beta}
590 -\mathsf{A}_{\alpha+2,\beta} \mathsf{B}_{\alpha+1,\beta}
591 \right]
592 \label{eq:matrixCross}
593 \end{equation}
594 where $\alpha+1$ and $\alpha+2$ are regarded as cyclic permuations of
595 the matrix indices. In table \ref{tab:UFT} we give compact
596 expressions for how the multipole sites interact with an external
597 field that has exhibits spatial variations.
598
599 \begin{table}
600 \caption{Potential energy $(U)$, force $(\mathbf{F})$, and torque
601 $(\mathbf{\tau})$ expressions for a multipolar site embedded in an
602 electric field with spatial variations, $\mathbf{E}(\mathbf{r})$.
603 \label{tab:UFT}}
604 \begin{tabular}{r|ccc}
605 & Charge & Dipole & Quadrupole \\ \hline
606 $U$ & $C \phi(\mathbf{r})$ & $-\mathbf{D} \cdot \mathbf{E}(\mathbf{r})$ & $- \mathsf{Q}:\nabla \mathbf{E}(\mathbf{r})$ \\
607 $\mathbf{F}$ & $C \mathbf{E}(\mathbf{r})$ & $+\mathbf{D} \cdot \nabla \mathbf{E}(\mathbf{r})$ & $+\mathsf{Q} : \nabla\nabla\mathbf{E}(\mathbf{r})$ \\
608 $\mathbf{\tau}$ & & $\mathbf{D} \times \mathbf{E}(\mathbf{r})$ & $+2 \mathsf{Q} \times \nabla \mathbf{E}(\mathbf{r})$
609 \end{tabular}
610 \end{table}
611 \section{Gradient of the field due to quadrupolar polarization}
612 \label{singularQuad}
613 In this section, we will discuss the gradient of the field produced by
614 quadrupolar polarization. For this purpose, we consider a distribution
615 of charge ${\rho}(r)$ which gives rise to an electric field
616 $\vec{E}(r)$ and gradient of the field $\vec{\nabla} \vec{E}(r)$
617 throughout space. The total gradient of the electric field over volume
618 due to the all charges within the sphere of radius $R$ is given by
619 (cf. Jackson equation 4.14):
620 \begin{equation}
621 \int_{r<R} \vec{\nabla}\vec{E}\;d^3r = -\int_{r=R} R^2 \vec{E}\;\hat{n}\; d\Omega
622 \label{eq:8}
623 \end{equation}
624 where $d\Omega$ is the solid angle and $\hat{n}$ is the normal vector
625 of the surface of the sphere which is equal to
626 $sin[\theta]cos[\phi]\hat{x} + sin[\theta]sin[\phi]\hat{y} +
627 cos[\theta]\hat{z}$
628 in spherical coordinates. For the charge density ${\rho}(r')$, the
629 total gradient of the electric field can be written as (cf. Jackson
630 equation 4.16),
631 \begin{equation}
632 \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
633 \label{eq:9}
634 \end{equation}
635 The radial function in the equation (\ref{eq:9}) can be expressed in
636 terms of spherical harmonics as (cf. Jackson equation 3.70),
637 \begin{equation}
638 \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)
639 \label{eq:10}
640 \end{equation}
641 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,
642 \begin{equation}
643 \begin{split}
644 \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 \\
645 &= -\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
646 '
647 \end{split}
648 \label{eq:11}
649 \end{equation}
650 The gradient of the product of radial function and spherical harmonics
651 is given by (cf. Arfken, p.811 eq. 16.94):
652 \begin{equation}
653 \begin{split}
654 \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
655 {\partial}{\partial r}+\frac{l}{r} \right]f(r)\; Y_{l, l-1, m}(\theta, \phi).
656 \end{split}
657 \label{eq:12}
658 \end{equation}
659 Using equation (\ref{eq:12}) we get,
660 \begin{equation}
661 \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}},
662 \label{eq:13}
663 \end{equation}
664 where $ Y_{l,l+1,m}(\theta, \phi)$ is the vector spherical harmonics
665 which can be expressed in terms of spherical harmonics as shown in
666 below (cf. Arfkan p.811),
667 \begin{equation}
668 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},
669 \label{eq:14}
670 \end{equation}
671 where $C(l+1,1,l|m_1,m_2,m)$ is a Clebsch-Gordan coefficient and
672 $\hat{e}_{m_2}$ is a spherical tensor of rank 1 which can be expressed
673 in terms of Cartesian coordinates,
674 \begin{equation}
675 {\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}}
676 \label{eq:15}
677 \end{equation}
678 The normal vector $\hat{n} $ can be expressed in terms of spherical tensor of rank 1 as shown in below,
679 \begin{equation}
680 \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)
681 \label{eq:16}
682 \end{equation}
683 The surface integral of the product of $\hat{n}$ and
684 ${Y_{l+1}}^{m_1}(\theta, \phi)$ gives,
685 \begin{equation}
686 \begin{split}
687 \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 \\
688 &= \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 \\
689 &= \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),
690 \end{split}
691 \label{eq:17}
692 \end{equation}
693 where ${Y_{l}}^{-m} = (-1)^m\;{{Y_{l}}^{m}}^* $ and
694 $ \int {{Y_{l}}^{m}}^*\;{Y_{l'}}^{m'}\;d\Omega =
695 \delta_{ll'}\delta_{mm'} $.
696 Non-vanishing values of equation \ref{eq:17} require $l = 0$,
697 therefore the value of $ m = 0 $. Since the values of $ m_1$ are -1,
698 1, and 0 then $m_2$ takes the values 1, -1, and 0, respectively
699 provided that $m = m_1 + m_2$. Equation \ref{eq:11} can therefore be
700 modified,
701 \begin{equation}
702 \begin{split}
703 \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(
704 1, 1, 0|0,0,0)\;{\hat{e}_{0}}{\hat{e}_{0}} ]\; d^3r'.
705 \end{split}
706 \label{eq:18}
707 \end{equation}
708 After substituting ${Y^*}_{00} = \frac{1}{\sqrt{4\pi}} $ and using the
709 values of the Clebsch-Gorden coefficients: $ C(1, 1, 0|-1,1,0) =
710 \frac{1}{\sqrt{3}}, \; C(1, 1, 0|-1,1,0)= \frac{1}{\sqrt{3}}$ and $
711 C(1, 1, 0|0,0,0) = -\frac{1}{\sqrt{3}}$ in equation \ref{eq:18} we
712 obtain,
713 \begin{equation}
714 \begin{split}
715 \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)\\
716 &= - \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).
717 \end{split}
718 \label{eq:19}
719 \end{equation}
720 Equation (\ref{eq:19}) gives the total gradient of the field over a
721 sphere due to the distribution of the charges. For quadrupolar fluids
722 the total charge within a sphere is zero, therefore
723 $ \int_{r<R} \vec{\nabla}\vec{E}\;d^3r = 0 $. Hence the quadrupolar
724 polarization produces zero net gradient of the field inside the
725 sphere.
726
727 \section{Applied field or field gradient}
728 \label{Ap:fieldOrGradient}
729
730 To satisfy the condition $ \nabla . E = 0 $, within the box of molecules we have taken electrostatic potential in the following form
731 \begin{equation}
732 \begin{split}
733 \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. \\
734 & \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),
735 \end{split}
736 \label{eq:appliedPotential}
737 \end{equation}
738 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,
739 \[\bf{E}
740 =\frac{g_o}{2} \left(\begin{array}{ccc}
741 2(a_1\; b_1 - \frac{cos\psi}{3})\;x \;+ (a_1\; b_2 \;+ a_2\; b_1)\;y + (a_1\; b_3 \;+ a_3\; b_1)\;z \\
742 (a_2\; b_1 \;+ a_1\; b_2)\;x + 2(a_2\; b_2 \;- \frac{cos\psi}{3})\;y + (a_2\; b_3 \;+ a_3\; b_3)\;z \\
743 (a_3\; b_1 \;+ a_3\; b_2)\;x + (a_3\; b_2 \;+ a_2\; b_3)y + 2(a_3\; b_3 \;- \frac{cos\psi}{3})\;z
744 \end{array} \right).\]
745 The gradient of the applied field derived from the potential can be written in the following form,
746 \[\nabla\bf{E}
747 = \frac{g_o}{2}\left(\begin{array}{ccc}
748 2(a_1\; b_1 - \frac{cos\psi}{3}) & (a_1\; b_2 \;+ a_2\; b_1) & (a_1\; b_3 \;+ a_3\; b_1)\;z \\
749 (a_2\; b_1 \;+ a_1\; b_2) & 2(a_2\; b_2 \;- \frac{cos\psi}{3}) & (a_2\; b_3 \;+ a_3\; b_3)\;z \\
750 (a_3\; b_1 \;+ a_3\; b_2) & (a_3\; b_2 \;+ a_2\; b_3) & 2(a_3\; b_3 \;- \frac{cos\psi}{3})\;z
751 \end{array} \right).\]
752 \newpage
753
754 \bibliography{multipole}
755
756 \end{document}