ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/multipole/dielectric.tex
Revision: 4353
Committed: Tue Sep 15 15:56:07 2015 UTC (9 years, 10 months ago) by mlamichh
Content type: application/x-tex
File size: 40716 byte(s)
Log Message:
Added new tex file for dielectric.

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
22 \begin{document}
23
24 \title{Real space electrostatics for multipoles. III. Dielectric Properties}
25
26 \author{Madan Lamichhane}
27 \affiliation{Department of Physics, University
28 of Notre Dame, Notre Dame, IN 46556}
29 \author{Thomas Parsons}
30 \affiliation{Department of Chemistry and Biochemistry, University
31 of Notre Dame, Notre Dame, IN 46556}
32 \author{Kathie E. Newman}
33 \affiliation{Department of Physics, University
34 of Notre Dame, Notre Dame, IN 46556}
35 \author{J. Daniel Gezelter}
36 \email{gezelter@nd.edu.}
37 \affiliation{Department of Chemistry and Biochemistry, University
38 of Notre Dame, Notre Dame, IN 46556}
39
40 \date{\today}% It is always \today, today,
41 % but any date may be explicitly specified
42
43 \begin{abstract}
44 Note: This manuscript is a work in progress.
45
46 We report on the dielectric properties of the shifted potential
47 (SP), gradient shifted force (GSF), and Taylor shifted force (TSF)
48 real-space methods for multipole interactions that were developed in
49 the first two papers in this series. We find that some subtlety is
50 required for computing dielectric properties with the real-space
51 methods, particularly when using the common fluctuation formulae.
52 Three distinct methods for computing the dielectric constant are
53 investigated, including the standard fluctuation formulae,
54 potentials of mean force between solvated ions, and direct
55 measurement of linear solvent polarization in response to applied
56 fields and field gradients.
57 \end{abstract}
58
59 \maketitle
60
61 \section{Introduction}
62
63 Over the past several years, there has been increasing interest in
64 pairwise methods for correcting electrostatic interactions in computer
65 simulations of condensed molecular
66 systems.\cite{Wolf99,Zahn02,Kast03,Beckd.A.C._Bi0486381,Ma05,Fennell06}
67 These techniques were developed from the observations and efforts of
68 Wolf {\it et al.} and their work towards an $\mathcal{O}(N)$
69 Coulombic sum.\cite{Wolf99} Wolf's method of cutoff neutralization is
70 able to obtain excellent agreement with Madelung energies in ionic
71 crystals.\cite{Wolf99} One of the most difficult tests of any new
72 electrostatic method is the fidelity with which that method can
73 reproduce the bulk-phase polarizability or equivalently, the
74 dielectric properties of a fluid. Of particular interest is the
75 static dielectric constant, $\epsilon$. Using the Ewald sum under
76 tin-foil boundary conditions, $\epsilon$ can be calculated for systems
77 of non-polarizable substances via
78 \begin{equation}
79 \epsilon = 1 + \frac{\langle M^2\rangle}{3\epsilon_0k_\textrm{B}TV},
80 \label{eq:staticDielectric}
81 \end{equation}
82 where $\epsilon_0$ is the permittivity of free space and $\langle
83 M^2\rangle$ is the fluctuation of the system dipole
84 moment.\cite{Allen:1989fp} The numerator in the fractional term in equation
85 (\ref{eq:staticDielectric}) is the fluctuation of the simulation-box
86 dipole moment, identical to the quantity calculated in the
87 finite-system Kirkwood $g$ factor ($G_k$):
88 \begin{equation}
89 G_k = \frac{\langle M^2\rangle}{N\mu^2},
90 \label{eq:KirkwoodFactor}
91 \end{equation}
92 where $\mu$ is the dipole moment of a single molecule of the
93 homogeneous system.\cite{Neumann:1983mz,Neumann:1983yq,Neumann84,Neumann85} The
94 fluctuation term in both equation (\ref{eq:staticDielectric}) and
95 (\ref{eq:KirkwoodFactor}) is calculated as follows,
96 \begin{equation}
97 \begin{split}
98 \langle M^2\rangle &= \langle\bm{M}\cdot\bm{M}\rangle
99 - \langle\bm{M}\rangle\cdot\langle\bm{M}\rangle \\
100 &= \langle M_x^2+M_y^2+M_z^2\rangle
101 - (\langle M_x\rangle^2 + \langle M_y\rangle^2
102 + \langle M_z\rangle^2).
103 \end{split}
104 \label{eq:fluctBoxDipole}
105 \end{equation}
106 This fluctuation term can be accumulated during a simulation; however,
107 it converges rather slowly, thus requiring multi-nanosecond simulation
108 times.\cite{Horn04} In the case of tin-foil boundary conditions, the
109 dielectric/surface term of the Ewald summation is equal to zero.
110
111 \section{Dielectric constants for real-space electrostatics}
112
113 The new real-space methods require some care when computing dielectric
114 properties because the interaction between molecular dipoles depends
115 on the level of approximation being utilized. Consider the cases of
116 Stockmayer (dipolar) soft spheres that are represented either by two
117 closely-spaced point charges or by a single point dipole (see
118 Fig. \ref{fig:stockmayer}).
119 \begin{figure}
120 \includegraphics[width=3in]{DielectricFigure}
121 \caption{With the real-space electrostatic methods, the effective
122 dipole tensor, $\mathbf{T}$, governing interactions between
123 molecular dipoles is not the same for charge-charge interactions as
124 for point dipoles.}
125 \label{fig:stockmayer}
126 \end{figure}
127 In the case where point charges are interacting via an electrostatic
128 kernel, $v(r)$, the effective {\it molecular} dipole tensor,
129 $\mathbf{T}$ is obtained from two successive applications of the
130 gradient operator to the electrostatic kernel,
131 \begin{equation}
132 \mathbf{T}_{\alpha \beta}(r) = \nabla_\alpha \nabla_\beta \left(v(r)\right) = \delta_{\alpha \beta}
133 \left(\frac{1}{r} v^\prime(r) \right) + \frac{r_{\alpha}
134 r_{\beta}}{r^2} \left( v^{\prime \prime}(r) - \frac{1}{r}
135 v^{\prime}(r) \right)
136 \end{equation}
137 where $v(r)$ may be either the bare kernel ($1/r$) or one of the
138 modified (Wolf or DSF) kernels. This tensor describes the effective
139 interaction between molecular dipoles ($\mathbf{D}$) in Gaussian
140 units as $-\mathbf{D} \cdot \mathbf{T} \cdot \mathbf{D}$.
141
142 When utilizing the new real-space methods for point dipoles, the
143 tensor is explicitly constructed,
144 \begin{equation}
145 \mathbf{T}_{\alpha \beta}(r) = \delta_{\alpha \beta} v_{21}(r) +
146 \frac{r_{\alpha} r_{\beta}}{r^2} v_{22}(r)
147 \end{equation}
148 where the functions $v_{21}(r)$ and $v_{22}(r)$ depend on the level of
149 the approximation. Although the Taylor-shifted (TSF) and
150 gradient-shifted (GSF) models produce to the same $v(r)$ function for
151 point charges, they have distinct forms for the dipole-dipole
152 interactions.
153
154 Neumann and Steinhauser showed~\cite{Neumann:1983mz,Neumann:1983yq}
155 that the relative dielectric permittivity $\epsilon$ is given by the
156 general fluctuation formula,
157 \begin{equation}
158 \frac{\epsilon - 1}{\epsilon +2} = \frac{4 \pi}{3} \frac{\left<M^2\right>}{3 V k_B
159 T} \left( 1 - \frac{3}{4 \pi} \frac{\epsilon -1}{\epsilon + 2}
160 \tilde{T}(0) \right)
161 \end{equation}
162 where $\left<M^2\right>$ is the mean fluctuation of the square of the
163 net dipole moment of the simulation cell, $V$ is the volume of the
164 cell, and $\tilde{T}(0) = \tilde{T̃}_{xx}(0) = \tilde{T̃}_{yy}(0) =
165 \tilde{T̃}_{zz}(0)$ is the $\mathbf{k} = 0$ limit of Fourier transform
166 of the diagonal term of the effective molecular dipolar tensor,
167 \begin{equation}
168 \tilde{T}(0) = \int_V \mathbf{T}(\mathbf{r}) d\mathbf{r}
169 \end{equation}
170 where the integral is carried out over the relevant geometry for the
171 interaction. For the real-space methods, the integration can be done
172 easily up to the imposed cutoff radius, while for the Ewald sum, the
173 results also depend on details of the $k$-space
174 calculation.\cite{Neumann:1983yq}
175
176 For molecules composed of point charges interacting via the bare
177 $(1/r)$ kernel, the simple cutoff (SC) and minimum image (MI)
178 approaches give $\tilde{T}(0) = 0$ which means that the
179 Clausius-Mosotti equation,
180 \begin{equation}
181 \frac{4\pi}{3}\frac{\left< M^2 \right>}{3 V k_B T} = \frac{\epsilon
182 -1}{\epsilon+2}
183 \end{equation}
184 is the relevant fluctuation formula for the relative permittivity.
185
186 Zahn {\it et al.}\cite{Zahn02} showed that for the damped shifted
187 force charge-charge kernel, $\tilde{T}(0) = 4 \pi / 3$. This was later
188 generalized by Izvekov {\it et al.} for all point-charge kernels which
189 have forces that go to zero at a cutoff radius and which maintain a
190 pole of first order at $r=0$.\cite{Izvekov:2008wo} When $\tilde{T}(0)
191 = 4 \pi/ 3$, the expression for the dielectric constant reduces to the
192 widely-used {\it conducting boundary} formula,
193 \begin{equation}
194 \frac{4\pi}{3}\frac{\left< M^2 \right>}{3 V k_B T} = \frac{\epsilon -
195 1}{3}
196 \end{equation}
197 It is convenient to define a quantity $A = \frac{3}{4 \pi}
198 \tilde{T}(0)$, and to combine all of the fluctuation formulae
199 together:
200 \begin{equation}
201 \frac{4 \pi}{3} \frac{\left< M^2 \right>}{3 V k_B T} =
202 \left\{\begin{array}{ll}
203 \frac{\epsilon-1}{\epsilon+2} \left[1- \frac{\epsilon-1}{\epsilon+2}
204 A\right]^{-1} & \mathrm{General~Case} \\
205 \frac{\epsilon-1}{\epsilon+2} & A \rightarrow 0 \mathrm{~limit} \\
206 \frac{\epsilon-1}{3} & A \rightarrow 1 \mathrm{~limit}
207 \end{array}\right.
208 \end{equation}
209
210 The Clausius-Mossotti $(A\rightarrow 0)$ approach is subject to noise
211 and error magnification,\cite{Allen:1989fp} and the conducting
212 boundary approach $(A \rightarrow 1)$ has become widely used as a
213 result. If the electrostatic method being used has $A < 1$, the
214 relative dielectric permittivity $\epsilon$ can be estimated once the
215 conducting boundary fluctuation result has been found,
216 \begin{equation}
217 \epsilon = \frac{(A+2) (\epsilon_\text{CB}-1)+3} {(A-1) (\epsilon_\text{CB}-1)+3}
218 \end{equation}
219 Note that this expression becomes quite numerically sensitive when the
220 value of $A$ deviates significantly from 1.
221
222 We have derived a set of expressions for the value of $A$ for the new
223 real space methods that are shown in Table \ref{tab:A}.
224
225 \begin{table}
226 \caption{Expressions for the dielectric correction factor ($A$) for the
227 real-space electrostatic methods in terms of the damping parameter
228 ($\alpha$) and the cutoff radius ($r_c$). The Ewald-Kornfeld result
229 derived in Refs. \onlinecite{Adams:1980rt,Adams:1981fr,Neumann:1983yq} is shown for comparison using the Ewald
230 convergence parameter ($\kappa$) and the real-space cutoff value ($r_c$). }
231 \label{tab:A}
232 \resizebox{\columnwidth}{!}{%
233 \begin{tabular}{l|c|c|c|c|c|}
234 & \multicolumn{3}{c|}{damped} & \multicolumn{2}{c|}{undamped} \\ \cline{2-4} \cline{5-6}
235 Method & $A_\mathrm{charges}$ & $A_\mathrm{dipoles}$ & $A_\mathrm{quad} (length^{-2})$ & $A_\mathrm{charges}$ & $A_\mathrm{dipoles}$ \\
236 \hline
237 Spherical Cutoff (SC) & $\mathrm{erf}(r_c \alpha) - \frac{2 \alpha r_c}{\sqrt{\pi}} e^{-\alpha^2 r_c^2}$ & 0 &$-\frac{8{\alpha}^5{r_c}^3 }{5\sqrt{\pi}}e^{-{\alpha}^2{r_c}^2 }$ &0 &0\\
238 Shifted Potental (SP) & $ \mathrm{erf}(r_c \alpha) - \frac{2 \alpha r_c}{\sqrt{\pi}} e^{-\alpha^2 r_c^2}$ & $\mathrm{erf}(\alpha r_c)-\frac{2 \alpha r_c }{\sqrt{\pi }} \left(1 + \frac{2 \alpha^2 r_c^2}{3} \right) e^{-\alpha ^2 r_c^2} $&$-\frac{16{\alpha}^7{r_c}^5 }{15\sqrt{\pi}}e^{-{\alpha}^2{r_c}^2 }$ &0 &0\\
239 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} $ &$-\frac{4{\alpha}^7{r_c}^5 }{15\sqrt{\pi}}(-1+2\alpha ^2 r_c^2)e^{-\alpha^2 r_c^2}$ &1 &0\\
240 Taylor-shifted (TSF) & 1 & 1 &$\frac{6\;\mathrm{erf}(r_c \alpha)}{{r_c}^2} + \frac{4{\alpha}{r_c}}{15\sqrt{\pi}{r_c}^2}\left(45 + 30\alpha ^2 {r_c}^2 + 12\alpha^4 {r_c}^4 + 3\alpha^6 {r_c}^6 + 2 \alpha^8 {r_c}^8\right)e^{-\alpha^2 r_c^2}$ &1 & 1\\
241 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
242 \end{tabular}%
243 }
244 \end{table}
245
246 \section{Boltzman average of the dipole polarization density}
247 Consider a dipolar fluid with polarization density $\mathbf{P}(r)$ at
248 equlibrium at temperature $T$. The average of polarization density is
249 given by,
250 \begin{equation}
251 \left<\mathbf{P}(r)\right> = \frac{\int\mathbf{P}(r) e^{-H/k_B T} d
252 \Gamma^N}{\int e^{-H/k_B T} d \Gamma^N}
253 \end{equation}
254 where $d \Gamma^N = dr_1...dr_N dp_1...dp_N$ is the volume element in
255 phase space, and $H$ is the Hamiltonian of the $N$-particle system in
256 the absence of applied external field. The Hamiltonian of the system
257 is given by,
258 \begin{equation}
259 H = \frac{1}{2}\sum_{i=1}^N m_i {\dot{\mathbf{r}}_i}^2 + \sum_{i<j} U(\mathbf{r}_i - \mathbf{r}_j)
260 \end{equation}
261 where $m_i$ is mass of site $i$, and $U(\mathbf{r}_i - \mathbf{r}_j)$
262 is the interaction potential between sites $i$ and $j$.
263
264 Let $E'$ be the applied external electric field. The Hamiltonian for
265 the polarized system then becomes
266 \begin{equation}
267 H' = H - {M}_\alpha {E'}_\alpha
268 \end{equation}
269 where $\mathbf{M}$ is the total dipole moment in zero field. The
270 polarization density changes in the presence of the external field
271 $\mathbf{E'}$. The modified polarization is given by
272 \begin{equation}
273 \left< P'_\alpha(r) \right>_{E'} = \frac{\int P_\alpha(r) e^{-H'/k_B
274 T} d \Gamma^N}{\int e^{-H'/k_B T} d \Gamma^N} \equiv \frac{F}{G}
275 \end{equation}
276 For small values of the applied electric field we can expand the
277 polarization density as,
278 \begin{equation}
279 \left< P'_\alpha(r) \right>_{E'} = \left< P'_\alpha(r) \right> + \frac{\partial
280 \left< P'_\alpha(r) \right>}{\partial E'_\beta }\bigg|_{E'=0} E'_\beta
281 \end{equation}
282 and
283 \begin{equation}
284 \frac{\partial{\left<P'_\alpha(r)\right>}}{\partial{E'_\beta}}\bigg|_{E'=0} = \frac{F'G-FG'}{G^2}
285 \end{equation}
286 where
287 \begin{equation}
288 \begin{split}
289 F' & = \frac{1}{k_BT} \int {P}_\alpha M_\beta\; e^{-H/k_B T} d\Gamma^N\\
290 \textit{and} \\
291 G' & = \frac{1}{k_BT} \int M_\beta\; e^{-H/k_B T} d\Gamma^N
292 \end{split}
293 \end{equation}
294 Therefore
295 \begin{equation}
296 \braket{{P'_\alpha(r)}}_{E'}=\braket{{P'_\alpha(r)}}+\frac{1}{k_B T}(\braket{P_{\alpha}M_{\beta}}-\braket{P_{\alpha}}\braket{M_{\beta}}){E'}_\beta,
297 \end{equation}
298 where $\braket{A} = \frac{\int A\; e^{-H/k_B T}d\Gamma^N}{\int e^{-H/k_B T}d\Gamma^N}$ is the ensemble average of a physical quantity $A$ at temperature $T$. Now the polarization density due to applied external field is given by,
299 \begin{equation}
300 {\Delta P}_\alpha =\frac{1}{k_B T}\left(\braket{P_{\alpha}M_{\beta}}-\braket{P_{\alpha}}\braket{M_{\beta}}\right){E'}_\beta.
301 \end{equation}
302 Therefore the total polarization for the system can be obtained by integrating polarization density over the volume,
303 \begin{equation}
304 {\Delta M}_\alpha =\frac{1}{k_B T} \left(\braket{M_{\alpha}M_{\beta}}-\braket{M_{\alpha}}\braket{M_{\beta}}\right){E'}_\beta.
305 \end{equation}
306 If we consider both applied field and polarization in the z-direction,
307 \begin{equation}
308 {\Delta M}_z =\frac{1}{k_B T}\left(\braket{{M_{z}}^2}-\braket{M_{z}}^2\right){E'}_z.
309 \end{equation}
310 In general applied field $ \braket{{M_{z}}^2} = \frac{1}{3}\braket{{\mathbf{M}}^2}$ and $ \braket{{M_{z}}} = \frac{1}{3}\braket{{\mathbf{M}}}$ therefore,
311 \begin{equation}
312 {\Delta P}_z = \frac{{\Delta M}_z}{V} = \epsilon_o \frac{\braket{{\mathbf{M}}^2}-\braket{{\mathbf{M}}}^2}{3 \epsilon_o \; k_B T V}{E'}_z.
313 \label{eq:fluc}
314 \end{equation}
315 Comparing equation (\ref{eq:fluc}) with $ \Delta P_z = \epsilon_o \chi \ {E'}_z$ the fluctuation formula for susceptibility ($\chi$) can be written as,
316 \begin{equation}
317 \chi = \frac{\braket{{\mathbf{M}}^2}-\braket{{\mathbf{M}}}^2}{3 \epsilon_o \; k_B T V}
318 \end{equation}
319 \section{Boltzmann average of a quadrupole moment}
320 The average quadrupole moment at temperature T in the presence of uniform applied
321 field gradient is given by,
322 \begin{equation}
323 \braket{q_{\alpha\beta}} \;=\; \frac{\int d\Omega\; e^{\frac{-U(r)}{k_B T}}q_{\alpha\beta}}{\int d\Omega\; e^{\frac{-U(r)}{k_B T}}} \;=\; \frac{\int d\Omega\; e^{\frac{q_{\mu\nu}\;\partial_\nu E_\mu}{k_B T}}q_{\alpha\beta}}{\int d\Omega\; e^{\frac{q_{\mu\nu}\;\partial_\nu E_\mu}{k_B T}}},
324 \label{eq:1}
325 \end{equation}
326 where $\int \Omega = \int_0^{2\pi} \int_0^\pi \int_0^{2\pi}
327 sin\theta\; d\theta\ d\phi\ d\psi$ is the integration over Euler
328 angles and $ U(r) = -q_{\mu\nu}\;\partial_\nu E_\mu $ is the energy of
329 a quadrupole in the gradient of the
330 applied field. The energy and quadrupole moment can be transformed
331 into body frame using following relation,
332 \begin{equation}
333 \begin{split}
334 q_{\alpha\beta} &= \eta_{\alpha\alpha'}{q}^* _{\alpha'\beta'}\eta_{\beta'\beta} \\
335 U(r) &= - q:\vec{\nabla}\vec{E} = - q_{\mu\nu}\;\partial_\nu E_\mu = -\eta_{\mu\mu'}{q}^*_{\mu'\nu'}\eta_{\nu'\nu}\;\partial_\nu E_\mu.
336 \end{split}
337 \label{eq:2}
338 \end{equation}
339 Here the starred tensors are the components in the body fixed
340 frame. Substituting equation (\ref{eq:2}) in the equation (\ref{eq:1})
341 and taking linear terms in the expansion we get,
342 \begin{equation}
343 \braket{q_{\alpha\beta}} = \frac{ \int d\Omega \left(1 + \frac{\eta_{\mu\mu'}{q}^*_{\mu'\nu'}\eta_{\nu'\nu}\;\partial_\nu E_\mu }{k_B T}\right)q_{\alpha\beta}}{ \int d\Omega \left(1 + \frac{\eta_{\mu\mu'}{q}^*_{\mu'\nu'}\eta_{\nu'\nu}\;\partial_\nu E_\mu }{k_B T}\right)},
344 \label{eq:3}
345 \end{equation}
346 where $\eta_{\alpha\alpha'}$ is the rotation matrix that transforms
347 the body fixed to the space fixed frame,
348 \[\eta_{\alpha\alpha'}
349 = \left(\begin{array}{ccc}
350 cos\phi\; cos\psi - cos\theta\; sin\phi\; sin\psi & cos\theta\; cos\psi\; sin\phi + cos\phi\; sin\psi & sin\theta\; sin\phi \\
351 -cos\psi\; sin\phi - cos\theta\; cos\phi \; sin\psi & cos\theta\; cos\phi\; cos\psi - sin\phi\; sin\psi & cos\phi\; cos\theta \\
352 sin\theta\; sin\psi & -cos\psi\; sin\theta & cos\theta
353 \end{array} \right).\]
354 Integration of 1st and 2nd terms in the denominator gives $8 \pi^2$
355 and $8 \pi^2 /3\;\vec{\nabla}.\vec{E}\; Tr(q^*) $ respectively. The
356 second term vanishes for charge free space
357 (i.e. $\vec{\nabla}.\vec{E} \; = \; 0)$. Similarly integration of the
358 1st term in the numerator produces
359 $8 \pi^2 /3\; Tr(q^*)\delta_{\alpha\beta}$ and the 2nd term produces
360 $8 \pi^2 /15k_B T (3{q}^*_{\alpha'\beta'}{q}^*_{\beta'\alpha'} -
361 {q}^*_{\alpha'\alpha'}{q}^*_{\beta'\beta'})\partial_\alpha E_\beta$,
362 if $\vec{\nabla}.\vec{E} \; = \; 0$,
363 $ \partial_\alpha E_\beta = \partial_\beta E_\alpha$ and
364 ${q}^*_{\alpha'\beta'}= {q}^*_{\beta'\alpha'}$. Therefore the
365 Boltzmann average of a quadrupole moment can be written as,
366
367 \begin{equation}
368 \braket{q_{\alpha\beta}}\; = \; \frac{1}{3} Tr(q^*)\;\delta_{\alpha\beta} + \frac{{\bar{q_o}}^2}{15k_BT}\;\partial_\alpha E_\beta,
369 \label{eq:4}
370 \end{equation}
371 here ${\bar{q_o}}^2=
372 3{q}^*_{\alpha'\beta'}{q}^*_{\beta'\alpha'}-{q}^*_{\alpha'\alpha'}{q}^*_{\beta'\beta'}$
373 which is a net quadrupole moment of a molecule.
374
375 To find the average quadrupole moment for the entire system, we can
376 multiply above equation with number of molecules $N$.
377 \begin{equation}
378 \braket{Q_{\alpha\beta}} \;=\; N\braket{q_{\alpha\beta}}\; = \; \frac{N}{3} Tr(q^*)\;\delta_{\alpha\beta} + \frac{{N\;\bar{q_o}}^2}{15k_BT}\;\partial_\alpha E_\beta
379 \label{eq:5}
380 \end{equation}
381 The analogous relation in a system of the $N$ quadrupoles is,
382 \begin{equation}
383 \braket{Q_{\alpha\beta}} \; = \; \frac{1}{3}\;Tr(Q)\;\delta_{\alpha\beta} + \frac{{\braket{Q^2}}_{en} - {\braket{Q}}_{en}^2}{15k_BT}\;\partial_\alpha E_\beta
384 \label{eq:6}
385 \end{equation}
386 where $\braket{{Q^2}}_{en}$ and $\braket{Q}_{en}^2$ are the ensemble
387 average of the square of the magnitude of the box quadrupole moment
388 and the square of the ensemble average of the box quadrupole moment
389 respectively. The ensemble averages in the above equation can be
390 obtained using relations,
391 ${\braket{Q^2}}_{en} =
392 {\braket{3Q_{\alpha'\beta'}Q_{\beta'\alpha'}-Q_{\alpha'\alpha'}Q_{\beta'\beta'}}}_{en}$
393 and
394 ${\braket{Q}}_{en}^2 = 3{\braket{Q_{\alpha'\beta'}}}_{en}\;
395 \braket{Q_{\beta'\alpha'}}_{en} -{\braket{Q_{\alpha'\alpha'}}}_{en}\;
396 {\braket{Q_{\beta'\beta'}}}_{en}$
397 respectively. The non-zero value of the $\braket{Q}_{en}^2$ is due to
398 the applied gradient, therefore its contribution is subtracted to find
399 the actual fluctuation in the box quadrupole moment. From the
400 equations (\ref{eq:5}) and (\ref{eq:6}) we get,
401 \begin{equation}
402 \begin{split}
403 \braket{{Q^2}}_{en}-{\braket{Q}_{en}}^2 &= N\; {\bar{q_o}}^2\\
404 & and \\
405 \braket{Q_{\alpha\beta}}\; &= \; N\braket{q_{\alpha\beta}}.
406 \end{split}
407 \label{eq:7}
408 \end{equation}
409
410
411
412 \section{Relation between quadrupolar polarization and applied field gradient}
413 In the presence of the field gradient $\partial_\alpha {E}_\beta $, a
414 non-vanishing quadrupolar polarization (quadrupole moment per unit
415 volume) $\bar{Q}_{\alpha\beta}$ will be induced in the entire volume
416 of a sample. The total field at any point $\vec{r}$ in the sample is
417 given by,
418 \begin{equation}
419 \partial_\alpha E_\beta(\vec{r}) = \partial_\alpha {E^o}_\beta(\vec{r}) + \frac{1}{4\pi \epsilon_o}\int T_{\alpha\beta\gamma\delta}(\vec{r}-\vec{r'})\;\bar{Q}_{\gamma\delta}(\vec{r'})\; d^3r'
420 \label{eq:20}
421 \end{equation}
422 where $\partial_\alpha {E^o}_\beta$ is the applied field gradient and $ T_{\alpha\beta\gamma\delta}$ is the quadrupole-quadrupole interaction tensor which is obtained from applications of four successive gradient operator to the electrostatic kernal $v(r)$.
423 \begin{equation}
424 \begin{split}
425 T_{\alpha\beta\gamma\delta}(r) &=\nabla_\alpha \nabla_\beta \nabla_\gamma \nabla_\delta\;v(r) \\ &= \left(\delta_{\alpha\beta}\delta_{\gamma\delta} + \delta_{\alpha\gamma}\delta_{\beta\delta}+ \delta_{\alpha\delta}\delta_{\beta\gamma}\right)v_{41}(r) + \left(\delta_{\alpha\beta} r_\gamma r_\delta + 5 \; permutations \right) \frac{v_{42}(r)}{r^2} \\ &+ r_\alpha r_\beta r_\gamma r_\delta\; \frac{v_{43}(r)}{r^4},
426 \end{split}
427 \label{quadRadial}
428 \end{equation}
429 where $v_{41}(r),\; v_{42}(r), \; and v_{43}(r)$ are defined in Paper I of the series (Refer Paper I) which have different functional forms for different cutoff methods. The integral in equation
430 (\ref{eq:20}) can be divided into two parts, $|\vec{r}-\vec{r'}|< R $
431 and $|\vec{r}-\vec{r'}|> R $ where $R \rightarrow 0 $. Since the total
432 field gradient vanishes at the singularity (see Appendix \ref{singular}),
433 equation (\ref{eq:20}) can be written as,
434 \begin{equation}
435 \partial_\alpha E_\beta(\vec{r}) = \partial_\alpha {E^o}_\beta(\vec{r}) +
436 \frac{1}{4\pi \epsilon_o}\lim_{R \to
437 0}\int\limits_{|\vec{r}-\vec{r'}|> R }
438 T_{\alpha\beta\gamma\delta}(|\vec{r}-\vec{r'}|)\;\bar{Q}_{\gamma\delta}(\vec{r'})\;
439 d^3r'.
440 \label{eq:23}
441 \end{equation}
442 If $r = r'$ is excluded from the integration, the gradient of the electric can be expressed in terms of traceless quadrupole moment as below (Refer eq 2.7 of Logan I) ,
443 \begin{equation}
444 \partial_\alpha E_\beta(\vec{r}) = \partial_\alpha {E^o}_\beta(\vec{r}) + \frac{1}{12\pi \epsilon_o}\int\limits_{|\vec{r}-\vec{r'}|> R } T_{\alpha\beta\gamma\delta}(|\vec{r}-\vec{r'}|)\;\bar{\Theta}_{\gamma\delta}(\vec{r'})\; d^3r',
445 \label{eq:22}
446 \end{equation}
447 where $\Theta_{\alpha\beta} = 3Q_{\alpha\beta} - \delta_{\alpha\beta}Tr(Q)$
448 is the traceless quadrupole moment. The total quadrupolar polarization is written as,
449 \begin{equation}
450 \bar{Q}_{\alpha\beta}(\vec{r}) = \frac{1}{3}\delta_{\alpha\beta}\;Tr(\bar{Q})+\epsilon_o {\chi}_Q\;\partial_\alpha E_\beta(\vec{r}),
451 \label{eq:21}
452 \end{equation}
453 where ${\chi}_Q = \frac{\braket{{Q^2}} - \braket{Q}^2}{15\epsilon_o Vk_BT}$ is the susceptibility of the quadrupolar fluid. In terms of traceless quadrupole moment, equation (\ref{eq:21}) can be written as,
454 \begin{equation}
455 \frac{1}{3}\bar{\Theta}_{\alpha\beta}(\vec{r}) = \epsilon_o {\chi}_Q \; \partial_\alpha E_\beta (\vec{r})= \epsilon_o {\chi}_Q \left(\partial_\alpha {E^o}_\beta(\vec{r}) + \frac{1}{12\pi \epsilon_o}\int\limits_{|\vec{r}-\vec{r'}|> R } T_{\alpha\beta\gamma\delta}(|\vec{r}-\vec{r'}|)\;\bar{\Theta}_{\gamma\delta}(\vec{r'})\; d^3r'\right)
456 \label{eq:22}
457 \end{equation}
458 For toroidal boundary conditions, both $\partial_\alpha E_\beta$ and
459 $\bar{\Theta}_{\alpha\beta}$ are uniform over the entire space. After
460 performing a Fourier transform (see the Appendix in the Neumann Paper),
461 we get,
462 \begin{equation}
463 \frac{1}{3}\widetilde{\bar{\Theta}}_{\alpha\beta}(\vec{k})=
464 \epsilon_o {\chi}_Q \;\left[{\partial_\alpha
465 {E^o}_\beta}(\vec{k})+ \frac{1}{12\pi
466 \epsilon_o}\;\widetilde{T}_{\alpha\beta\gamma\delta}(\vec{k})\;
467 {\bar{\Theta}}_{\gamma\delta}(\vec{k})\right]
468 \label{eq:24}
469 \end{equation}
470 Since the quadrupolar polarization is in the direction of the applied
471 field
472 $\widetilde{\bar{\Theta}}_{\gamma\delta}(\vec{k}) =
473 \widetilde{\bar{\Theta}}_{\alpha\beta}(\vec{k})$
474 and
475 $\widetilde{T}_{\alpha\beta\gamma\delta}(\vec{k}) =
476 \widetilde{T}_{\alpha\beta\alpha\beta}(\vec{k})$.
477 (Note: This is true for orientational polarization but might not be
478 true for polarization due to changes in the atomic distribution. For
479 the orientational case, we need only to work with individual
480 components of the matrix.)
481 \begin{equation}
482 \begin{split}
483 \frac{1}{3}\widetilde{\bar{\Theta}}_{\alpha\beta}(\vec{k}) &= \epsilon_o \bar{\chi}_Q \left[{\partial_\alpha E^o_\beta}(\vec{k})+ \frac{1}{12\pi \epsilon_o}\widetilde{T}_{\alpha\beta\alpha\beta}(\vec{k})\;\widetilde{\bar{\Theta}}_{\alpha\beta}(\vec{k})\right] \\
484 &= \epsilon_o {\chi}_Q\;\left(1-\frac{1}{4\pi} {\chi}_Q
485 \widetilde{T}_{\alpha\beta\alpha\beta}(\vec{k})\right)^{-1}
486 {\partial_\alpha E^o_\beta}(\vec{k})
487 \end{split}
488 \label{eq:25}
489 \end{equation}
490 Matrix contraction of quadrupole-quadrupole tensor and quadrupole
491 moment gives the identity matrix. Therefore each component can be
492 treated as scalar. If the field gradient is homogeneous over the
493 entire volume, ${\partial_ \alpha E_\beta}(\vec{k}) = 0 $ except at
494 $ \vec{k} = 0$, hence it is sufficient to know
495 $ \widetilde{T}_{\alpha\beta\alpha\beta}(\vec{k})$ at $ \vec{k} =
496 0$. Therefore equation (\ref{eq:25}) can be written as,
497 \begin{equation}
498 \begin{split}
499 \frac{1}{3}\widetilde{\bar{\Theta}}_{\alpha\beta}({0}) &= \epsilon_o {\chi}_Q\left(1-\frac{1}{4\pi} {\chi}_Q\;\widetilde{T}_{\alpha\beta\alpha\beta}({0})\right)^{-1} \partial_\alpha E^o_\beta({0})
500 \end{split}
501 \label{eq:26}
502 \end{equation}
503 where $ \widetilde{T}_{\alpha\beta\alpha\beta}(\vec{0})$ can be evaluated as,
504 \begin{equation}
505 \widetilde{T}_{\alpha\beta\alpha\beta}({0}) = \int \widetilde{T}_{\alpha\beta\alpha\beta}\;(\vec{r})d^3r
506 \label{eq:27}
507 \end{equation}
508 The susceptibility is modified as,
509 \begin{equation}
510 \Xi_Q = \bar{\chi}_Q\left(1-\frac{1}{4\pi} \bar{\chi}_Q\;\widetilde{T}_{\alpha\beta\alpha\beta}({0})\right)^{-1}.
511 \label{eq:28}
512 \end{equation}
513 In terms of traced quadrupole moment equation (\ref{eq:25}) can be written as,
514 \begin{equation}
515 {\bar{Q}}_{\alpha\beta} = \frac{1}{3}\delta_{\alpha\beta}\;Tr(\bar{Q}) + \epsilon_o\; \Xi_Q\; \partial_\alpha E^o_\beta
516 \label{eq:29}
517 \end{equation}
518 From equations (\ref{eq:6}), (\ref{eq:28}), and (\ref{eq:29}) we get,
519 \begin{equation}
520 \begin{split}
521 &\frac{\braket{{Q^2}}_{en} - \braket{Q}_{en}^2}{15 \epsilon_o Vk_BT}\; =\; {\chi}_Q\left(1-\frac{1}{4\pi} {\chi}_Q\;\widetilde{T}_{\alpha\beta\alpha\beta}({0})\right)^{-1}, \\
522 &{\chi}_Q \;=\; \frac{\braket{{Q^2}}_{en} - \braket{Q}_{en}^2}{15 \epsilon_o Vk_BT}\left(1 + \frac{1}{4\pi} \frac{\braket{{Q^2}}_{en} - \braket{Q}_{en}^2}{15 \epsilon_o Vk_BT}\;\widetilde{T}_{\alpha\beta\alpha\beta}({0})\right)^{-1}
523 \end{split}
524 \end{equation}
525 Finally in terms of quadrupolar correction factor ($A_{quad}$) the quadrupolar susceptibility is written as,
526 \begin{equation}
527 {\chi}_Q \;=\; \frac{\braket{{Q^2}}_{en} - \braket{Q}_{en}^2}{15 \epsilon_o Vk_BT}\left(1 + \frac{\braket{{Q^2}}_{en} - \braket{Q}_{en}^2}{15 \epsilon_o Vk_BT}\; A_{quad}\right)^{-1}
528 \end{equation}
529 The $A_{quad}$ has dimension of the $length^{-2}$ is different for different cutoff methods which is listed in Table I.
530 \section{The effective dielectric constant of a quadrupolar fluid}
531 The dielectric constant of a material is a measure of the screening of
532 the electrostatic energy when a material is placed in the
533 electrostatic field. If $U_{field}(r)$ and $U_{int}(r)$ are the total
534 electrostatic energies of a system due to the applied field and
535 interaction of the quadrupole with the gradient of the field
536 respectively. The dielectric constant can be written as,
537 \begin{equation}
538 \epsilon = \frac{U_{field} + U_{int}}{U_{field}}
539 \label{eq:31}
540 \end{equation}
541 The energy due to field can be written as,
542 \begin{equation}
543 U_{field}(r) = \int_V u(r)dv = \int_V \frac{1}{2}\epsilon_o {|E^o|}^2 dv,
544 \label{eq:32}
545 \end{equation}
546 where $u(r)$ is the energy density for applied electric field and $V$ is the total volume
547 of the system. The interaction energy can be assumed to be created by assembling quadrupoles from infinitely far away against the existing field gradient(Analogy Jackson's equations 4.83 - 4.94). Therefore
548 \begin{equation}
549 U_{int}(r) = -\frac{1}{2}\int_V \bar{Q}_{\alpha\beta}\;\partial_\beta E^o_{\alpha}\; dv,
550 \label{eq:33}
551 \end{equation}
552 where $\bar{Q}_{\alpha\beta}$ is the quadrupolar polarization. The quadrupoloarization can be expressed as linear function of the gradient of the applied field as below,
553 \begin{equation}
554 \bar{Q}_{\alpha\beta} = \epsilon_{o}\; {\chi_Q}\; \partial_\alpha E_\beta \approx \epsilon_{o}\; {\chi_Q}\; \partial_\alpha E^o_\beta ,
555 \label{eq:34}
556 \end{equation}
557 where ${\chi_Q} $ is a quadrupolar susceptibility. The local field gradient can be ignored in the region of low polarizability. In terms of the box quadrupole
558 moment it can be expressed as,
559 \begin{equation}
560 \chi_Q \;=\; \frac{\braket{{Q^2}}_{en} - \braket{Q}_{en}^2}{15 \epsilon_o Vk_BT}\left(1 + \frac{\braket{{Q^2}}_{en} - \braket{Q}_{en}^2}{15 \epsilon_o Vk_BT}\;A_{quad}\right
561 )^{-1}.
562 \label{eq:36}
563 \end{equation}
564 Using above relations (equation (\ref{eq:32}\;-\;\ref{eq:34}) in the
565 equation (\ref{eq:31}) produce,
566 \begin{equation}
567 \epsilon = 1 + {\chi}_Q \frac{\int_V |\partial_\alpha E^o_\beta|^2 dv}{\int_V{|E^o|}^2 dv}.
568 \end{equation}
569
570 \newpage
571
572 \appendix
573 \section{Point-multipolar interactions with a spatially-varying electric field}
574
575 We can treat objects $a$, $b$, and $c$ containing embedded collections
576 of charges. When we define the primitive moments, we sum over that
577 collections of charges using a local coordinate system within each
578 object. The point charge, dipole, and quadrupole for object $a$ are
579 given by $C_a$, $\mathbf{D}_a$, and $\mathsf{Q}_a$, respectively.
580 These are the primitive multipoles which can be expressed as a
581 distribution of charges,
582 \begin{align}
583 C_a =&\sum_{k \, \text{in }a} q_k , \label{eq:charge} \\
584 D_{a\alpha} =&\sum_{k \, \text{in }a} q_k r_{k\alpha}, \label{eq:dipole}\\
585 Q_{a\alpha\beta} =& \frac{1}{2} \sum_{k \, \text{in } a} q_k
586 r_{k\alpha} r_{k\beta} . \label{eq:quadrupole}
587 \end{align}
588 Note that the definition of the primitive quadrupole here differs from
589 the standard traceless form, and contains an additional Taylor-series
590 based factor of $1/2$. In Paper 1, we derived the forces and torques
591 each object exerts on the others.
592
593 Here we must also consider an external electric field that varies in
594 space: $\mathbf E(\mathbf r)$. Each of the local charges $q_k$ in
595 object $a$ will then experience a slightly different field. This
596 electric field can be expanded in a Taylor series around the local
597 origin of each object. A different Taylor series expansion is carried
598 out for each object.
599
600 For a particular charge $q_k$, the electric field at that site's
601 position is given by:
602 \begin{equation}
603 E_\gamma + \nabla_\delta E_\gamma r_{k \delta}
604 + \frac {1}{2} \nabla_\delta \nabla_\varepsilon E_\gamma r_{k \delta}
605 r_{k \varepsilon} + ...
606 \end{equation}
607 Note that the electric field is always evaluated at the origin of the
608 objects, and treating each object using point multipoles simplifies
609 this greatly.
610
611 To find the force exerted on object $a$ by the electric field, one
612 takes the electric field expression, and multiplies it by $q_k$, and
613 then sum over all charges in $a$:
614
615 \begin{align}
616 F_\gamma &= \sum_{k \textrm{~in~} a} q_k \lbrace E_\gamma + \nabla_\delta E_\gamma r_{k \delta}
617 + \frac {1}{2} \nabla_\delta \nabla_\varepsilon E_\gamma r_{k \delta}
618 r_{k \varepsilon} + ... \rbrace \\
619 &= C_a E_\gamma + D_{a \delta} \nabla_\delta E_\gamma
620 + Q_{a \delta \varepsilon} \nabla_\delta \nabla_\varepsilon E_\gamma +
621 ...
622 \end{align}
623
624 Similarly, the torque exerted by the field on $a$ can be expressed as
625 \begin{align}
626 \tau_\alpha &= \sum_{k \textrm{~in~} a} (\mathbf r_k \times q_k \mathbf E)_\alpha \\
627 & = \sum_{k \textrm{~in~} a} \epsilon_{\alpha \beta \gamma} q_k
628 r_{k\beta} E_\gamma(\mathbf r_k) \\
629 & = \epsilon_{\alpha \beta \gamma} D_\beta E_\gamma
630 + 2 \epsilon_{\alpha \beta \gamma} Q_{\beta \delta} \nabla_\delta
631 E_\gamma + ...
632 \end{align}
633
634 The last term is essentially identical with form derived by Torres del
635 Castillo and M\'{e}ndez Garrido,\cite{Torres-del-Castillo:2006uo} although their derivation
636 utilized a traceless form of the quadrupole that is different than the
637 primitive definition in use here. We note that the Levi-Civita symbol
638 can be eliminated by utilizing the matrix cross product in an
639 identical form as in Ref. \onlinecite{Smith98}:
640 \begin{equation}
641 \left[\mathsf{A} \times \mathsf{B}\right]_\alpha = \sum_\beta
642 \left[\mathsf{A}_{\alpha+1,\beta} \mathsf{B}_{\alpha+2,\beta}
643 -\mathsf{A}_{\alpha+2,\beta} \mathsf{B}_{\alpha+1,\beta}
644 \right]
645 \label{eq:matrixCross}
646 \end{equation}
647 where $\alpha+1$ and $\alpha+2$ are regarded as cyclic permuations of
648 the matrix indices. In table \ref{tab:UFT} we give compact
649 expressions for how the multipole sites interact with an external
650 field that has exhibits spatial variations.
651
652 \begin{table}
653 \caption{Potential energy $(U)$, force $(\mathbf{F})$, and torque
654 $(\mathbf{\tau})$ expressions for a multipolar site embedded in an
655 electric field with spatial variations, $\mathbf{E}(\mathbf{r})$.
656 \label{tab:UFT}}
657 \begin{tabular}{r|ccc}
658 & Charge & Dipole & Quadrupole \\ \hline
659 $U$ & $C \phi(\mathbf{r})$ & $-\mathbf{D} \cdot \mathbf{E}(\mathbf{r})$ & $- \mathsf{Q}:\nabla \mathbf{E}(\mathbf{r})$ \\
660 $\mathbf{F}$ & $C \mathbf{E}(\mathbf{r})$ & $+\mathbf{D} \cdot \nabla \mathbf{E}(\mathbf{r})$ & $+\mathsf{Q} : \nabla\nabla\mathbf{E}(\mathbf{r})$ \\
661 $\mathbf{\tau}$ & & $\mathbf{D} \times \mathbf{E}(\mathbf{r})$ & $+2 \mathsf{Q} \times \nabla \mathbf{E}(\mathbf{r})$
662 \end{tabular}
663 \end{table}
664 \section{Gradient of the field due to quadrupolar polarization}
665 \label{singular}
666 In this section, we will discuss the gradient of the field produced by
667 quadrupolar polarization. For this purpose, we consider a distribution
668 of charge ${\rho}(r)$ which gives rise to an electric field
669 $\vec{E}(r)$ and gradient of the field $\vec{\nabla} \vec{E}(r)$
670 throughout space. The total gradient of the electric field over volume
671 due to the all charges within the sphere of radius $R$ is given by
672 (cf. Jackson equation 4.14):
673 \begin{equation}
674 \int_{r<R} \vec{\nabla}\vec{E}\;d^3r = -\int_{r=R} R^2 \vec{E}\;\hat{n}\; d\Omega
675 \label{eq:8}
676 \end{equation}
677 where $d\Omega$ is the solid angle and $\hat{n}$ is the normal vector
678 of the surface of the sphere which is equal to
679 $sin[\theta]cos[\phi]\hat{x} + sin[\theta]sin[\phi]\hat{y} +
680 cos[\theta]\hat{z}$
681 in spherical coordinates. For the charge density ${\rho}(r')$, the
682 total gradient of the electric field can be written as (cf. Jackson
683 equation 4.16),
684 \begin{equation}
685 \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
686 \label{eq:9}
687 \end{equation}
688 The radial function in the equation (\ref{eq:9}) can be expressed in
689 terms of spherical harmonics as (cf. Jackson equation 3.70),
690 \begin{equation}
691 \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)
692 \label{eq:10}
693 \end{equation}
694 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,
695 \begin{equation}
696 \begin{split}
697 \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 \\
698 &= -\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
699 '
700 \end{split}
701 \label{eq:11}
702 \end{equation}
703 The gradient of the product of radial function and spherical harmonics
704 is given by (cf. Arfken, p.811 eq. 16.94):
705 \begin{equation}
706 \begin{split}
707 \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
708 {\partial}{\partial r}+\frac{l}{r} \right]f(r)\; Y_{l, l-1, m}(\theta, \phi).
709 \end{split}
710 \label{eq:12}
711 \end{equation}
712 Using equation (\ref{eq:12}) we get,
713 \begin{equation}
714 \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}},
715 \label{eq:13}
716 \end{equation}
717 where $ Y_{l,l+1,m}(\theta, \phi)$ is the vector spherical harmonics
718 which can be expressed in terms of spherical harmonics as shown in
719 below (cf. Arfkan p.811),
720 \begin{equation}
721 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},
722 \label{eq:14}
723 \end{equation}
724 where $C(l+1,1,l|m_1,m_2,m)$ is a Clebsch-Gordan coefficient and
725 $\hat{e}_{m_2}$ is a spherical tensor of rank 1 which can be expressed
726 in terms of Cartesian coordinates,
727 \begin{equation}
728 {\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}}
729 \label{eq:15}
730 \end{equation}
731 The normal vector $\hat{n} $ can be expressed in terms of spherical tensor of rank 1 as shown in below,
732 \begin{equation}
733 \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)
734 \label{eq:16}
735 \end{equation}
736 The surface integral of the product of $\hat{n}$ and
737 ${Y_{l+1}}^{m_1}(\theta, \phi)$ gives,
738 \begin{equation}
739 \begin{split}
740 \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 \\
741 &= \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 \\
742 &= \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),
743 \end{split}
744 \label{eq:17}
745 \end{equation}
746 where ${Y_{l}}^{-m} = (-1)^m\;{{Y_{l}}^{m}}^* $ and
747 $ \int {{Y_{l}}^{m}}^*\;{Y_{l'}}^{m'}\;d\Omega =
748 \delta_{ll'}\delta_{mm'} $.
749 Non-vanishing values of equation \ref{eq:17} require $l = 0$,
750 therefore the value of $ m = 0 $. Since the values of $ m_1$ are -1,
751 1, and 0 then $m_2$ takes the values 1, -1, and 0, respectively
752 provided that $m = m_1 + m_2$. Equation \ref{eq:11} can therefore be
753 modified,
754 \begin{equation}
755 \begin{split}
756 \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(
757 1, 1, 0|0,0,0)\;{\hat{e}_{0}}{\hat{e}_{0}} ]\; d^3r'.
758 \end{split}
759 \label{eq:18}
760 \end{equation}
761 After substituting ${Y^*}_{00} = \frac{1}{\sqrt{4\pi}} $ and using the
762 values of the Clebsch-Gorden coefficients: $ C(1, 1, 0|-1,1,0) =
763 \frac{1}{\sqrt{3}}, \; C(1, 1, 0|-1,1,0)= \frac{1}{\sqrt{3}}$ and $
764 C(1, 1, 0|0,0,0) = -\frac{1}{\sqrt{3}}$ in equation \ref{eq:18} we
765 obtain,
766 \begin{equation}
767 \begin{split}
768 \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)\\
769 &= - \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).
770 \end{split}
771 \label{eq:19}
772 \end{equation}
773 Equation (\ref{eq:19}) gives the total gradient of the field over a
774 sphere due to the distribution of the charges. For quadrupolar fluids
775 the total charge within a sphere is zero, therefore
776 $ \int_{r<R} \vec{\nabla}\vec{E}\;d^3r = 0 $. Hence the quadrupolar
777 polarization produces zero net gradient of the field inside the
778 sphere.
779
780 \newpage
781
782 \bibliography{multipole}
783
784 \end{document}