ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/multipole/Dielectric_Supplemental.tex
Revision: 4418
Committed: Sun Apr 10 15:13:19 2016 UTC (8 years, 2 months ago) by gezelter
Content type: application/x-tex
File size: 27482 byte(s)
Log Message:
Latest version of supplemental

File Contents

# Content
1 % ****** Start of file aipsamp.tex ******
2 %
3 % This file is part of the AIP files in the AIP distribution for REVTeX 4.
4 % Version 4.1 of REVTeX, October 2009
5 %
6 % Copyright (c) 2009 American Institute of Physics.
7 %
8 % See the AIP README file for restrictions and more information.
9 %
10 % TeX'ing this file requires that you have AMS-LaTeX 2.0 installed
11 % as well as the rest of the prerequisites for REVTeX 4.1
12 %
13 % It also requires running BibTeX. The commands are as follows:
14 %
15 % 1) latex aipsamp
16 % 2) bibtex aipsamp
17 % 3) latex aipsamp
18 % 4) latex aipsamp
19 %
20 % Use this file as a source of example code for your aip document.
21 % Use the file aiptemplate.tex as a template for your document.
22 \documentclass[%
23 aip,jcp,
24 amsmath,amssymb,
25 preprint,%
26 % reprint,%
27 %author-year,%
28 %author-numerical,%
29 jcp]{revtex4-1}
30
31 \usepackage{graphicx}% Include figure files
32 \usepackage{dcolumn}% Align table columns on decimal point
33 %\usepackage{bm}% bold math
34 \usepackage{times}
35 \usepackage[version=3]{mhchem} % this is a great package for formatting chemical reactions
36 \usepackage{url}
37 \usepackage{rotating}
38 \usepackage{braket}
39 \usepackage{array}
40 \newcolumntype{L}[1]{>{\raggedright\let\newline\\\arraybackslash\hspace{0pt}}m{#1}}
41 \newcolumntype{C}[1]{>{\centering\let\newline\\\arraybackslash\hspace{0pt}}m{#1}}
42 \newcolumntype{R}[1]{>{\raggedleft\let\newline\\\arraybackslash\hspace{0pt}}m{#1}}
43
44
45
46 %\usepackage[mathlines]{lineno}% Enable numbering of text and display math
47 %\linenumbers\relax % Commence numbering lines
48
49 \begin{document}
50
51 \title{Supplemental Material for: Real space electrostatics for
52 multipoles. III. Dielectric Properties}
53
54 \author{Madan Lamichhane}
55 \affiliation{Department of Physics, University
56 of Notre Dame, Notre Dame, IN 46556}
57 \author{Thomas Parsons}
58 \affiliation{Department of Chemistry and Biochemistry, University
59 of Notre Dame, Notre Dame, IN 46556}
60 \author{Kathie E. Newman}
61 \affiliation{Department of Physics, University
62 of Notre Dame, Notre Dame, IN 46556}
63 \author{J. Daniel Gezelter}
64 \email{gezelter@nd.edu.}
65 \affiliation{Department of Chemistry and Biochemistry, University
66 of Notre Dame, Notre Dame, IN 46556}
67
68 \date{\today}% It is always \today, today,
69 % but any date may be explicitly specified
70
71 \begin{abstract}
72 This document includes useful relationships for computing the
73 interactions between fields and field gradients and point multipolar
74 representations of molecular electrostatics. We also provide
75 explanatory derivations of a number of relationships used in the
76 main text. This includes the Boltzmann averages of quadrupole
77 orientations, and the interaction of a quadrupole with the
78 self-generated field gradient. This last relationship is assumed to
79 be zero in the main text but is explicitly shown to be zero here.
80 \end{abstract}
81
82 \maketitle
83
84 \section{Generating Uniform Field Gradients}
85 One important task in performing out the simulations mentioned in the
86 main text was to generate uniform electric field gradients. We rely
87 heavily on both the notation and results from Torres del Castillo and
88 Mend\'{e}z Garido.\cite{Torres-del-Castillo:2006uo} In this work,
89 tensors were expressed in Cartesian components, using at times a
90 dyadic notation. This proves quite useful for computer simulations
91 that make use of toroidal boundary conditions.
92
93 An alternative formalism uses the theory of angular momentum and
94 spherical harmonics and is common in standard physics texts such as
95 Jackson,\cite{Jackson98} Morse and Feshbach,\cite{Morse:1946zr} and
96 Stone.\cite{Stone:1997ly} Because this approach has its own
97 advantages, relationships are provided below comparing that
98 terminology to the Cartesian tensor notation.
99
100 The gradient of the electric field,
101 \begin{equation*}
102 \mathsf{G}(\mathbf{r}) = -\nabla \nabla \Phi(\mathbf{r}),
103 \end{equation*}
104 where $\Phi(\mathbf{r})$ is the electrostatic potential. In a
105 charge-free region of space, $\nabla \cdot \mathbf{E}=0$, and
106 $\mathsf{G}$ is a symmetric traceless tensor. From symmetry
107 arguments, we know that this tensor can be written in terms of just
108 five independent components.
109
110 Following Torres del Castillo and Mend\'{e}z Garido's notation, the
111 gradient of the electric field may also be written in terms of two
112 vectors $\mathbf{a}$ and $\mathbf{b}$,
113 \begin{equation*}
114 G_{ij}=\frac{1}{2} (a_i b_j + a_j b_i) - \frac{1}{3}(\mathbf a \cdot \mathbf b) \delta_{ij} .
115 \end{equation*}
116 If the vectors $\mathbf{a}$ and $\mathbf{b}$ are unit vectors, the
117 electrostatic potential that generates a uniform gradient may be
118 written:
119 \begin{align}
120 \Phi(x, y, z) =\; -\frac{g_o}{2} & \left(\left(a_1b_1 -
121 \frac{cos\psi}{3}\right)\;x^2+\left(a_2b_2
122 - \frac{cos\psi}{3}\right)\;y^2 +
123 \left(a_3b_3 -
124 \frac{cos\psi}{3}\right)\;z^2 \right. \nonumber \\
125 & + (a_1b_2 + a_2b_1)\; xy + (a_1b_3 + a_3b_1)\; xz + (a_2b_3 + a_3b_2)\; yz \bigg) .
126 \label{eq:appliedPotential}
127 \end{align}
128 Note $\mathbf{a}\cdot\mathbf{a} = \mathbf{b} \cdot \mathbf{b} = 1$,
129 $\mathbf{a} \cdot \mathbf{b}=\cos \psi$, and $g_0$ is the overall
130 strength of the potential.
131
132 Taking the gradient of Eq. (\ref{eq:appliedPotential}), we find the
133 field due to this potential,
134 \begin{equation}
135 \mathbf{E} = -\nabla \Phi
136 =\frac{g_o}{2} \left(\begin{array}{ccc}
137 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 \\
138 (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 \\
139 (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
140 \end{array} \right),
141 \label{eq:CE}
142 \end{equation}
143 while the gradient of the electric field in this form,
144 \begin{equation}
145 \mathsf{G} = \nabla\mathbf{E}
146 = \frac{g_o}{2}\left(\begin{array}{ccc}
147 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) \\
148 (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) \\
149 (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})
150 \end{array} \right),
151 \label{eq:GC}
152 \end{equation}
153 is uniform over the entire space. Therefore, to describe a uniform
154 gradient in this notation, two unit vectors ($\mathbf{a}$ and
155 $\mathbf{b}$) as well as a potential strength, $g_0$, must be
156 specified. As expected, this requires five independent parameters.
157
158 The common alternative to the Cartesian notation expresses the
159 electrostatic potential using the notation of Morse and
160 Feshbach,\cite{Morse:1946zr}
161 \begin{equation} \label{eq:quad_phi}
162 \Phi(x,y,z) = -\left[ a_{20} \frac{2 z^2 -x^2 - y^2}{2}
163 + 3 a_{21}^e \,xz + 3 a_{21}^o \,yz
164 + 6a_{22}^e \,xy + 3 a_{22}^o (x^2 - y^2) \right].
165 \end{equation}
166 Here we use the standard $(l,m)$ form for the $a_{lm}$ coefficients,
167 with superscript $e$ and $o$ denoting even and odd, respectively.
168 This form makes the functional analogy to ``d'' atomic states
169 apparent.
170
171 Applying the gradient operator to Eq. (\ref{eq:quad_phi}) the electric
172 field due to this potential,
173 \begin{equation}
174 \mathbf{E} = -\nabla \Phi
175 = \left(\begin{array}{ccc}
176 \left( 6a_{22}^o -a_{20} \right)\; x &+\; 6a_{22}^e\; y &+\; 3a_{21}^e\; z \\
177 6a_{22}^e\; x & -\; (a_{20} + 6a_{22}^o)\; y & +\; 3a_{21}^o\; z \\
178 3a_{21}^e\; x & +\; 3a_{21}^o\; y & +\; 2a_{20}\; z
179 \end{array} \right),
180 \label{eq:MFE}
181 \end{equation}
182 while the gradient of the electric field in this form is:
183 \begin{equation} \label{eq:grad_e2}
184 \mathsf{G} =
185 \begin{pmatrix}
186 6 a_{22}^o - a_{20} & 6a_{22}^e & 3a_{21}^e\\
187 6a_{22}^e & -(a_{20}+6a_{22}^o) & 3a_{21}^o \\
188 3a_{21}^e & 3a_{21}^o & 2a_{20} \\
189 \end{pmatrix} \\
190 \end{equation}
191 which is also uniform over the entire space. This form for the
192 gradient can be factored as
193 \begin{gather}
194 \begin{aligned}
195 \mathsf{G} = a_{20}
196 \begin{pmatrix}
197 -1 & 0 & 0\\
198 0 & -1 & 0\\
199 0 & 0 & 2\\
200 \end{pmatrix}
201 +3a_{21}^e
202 \begin{pmatrix}
203 0 & 0 & 1\\
204 0 & 0 & 0\\
205 1 & 0 & 0\\
206 \end{pmatrix}
207 +3a_{21}^o
208 \begin{pmatrix}
209 0 & 0 & 0\\
210 0 & 0 & 1\\
211 0 & 1 & 0\\
212 \end{pmatrix}
213 +6a_{22}^e
214 \begin{pmatrix}
215 0 & 1 & 0\\
216 1 & 0 & 0\\
217 0 & 0 & 0\\
218 \end{pmatrix}
219 +6a_{22}^o
220 \begin{pmatrix}
221 1 & 0 & 0\\
222 0 & -1 & 0\\
223 0 & 0 & 0\\
224 \end{pmatrix}
225 \end{aligned}
226 \label{eq:intro_tensors}
227 \end{gather}
228 The five matrices in the expression above represent five different
229 symmetric traceless tensors of rank 2.
230
231 It is useful to find the Cartesian vectors $\mathbf a$ and $\mathbf b$
232 that generate the five types of tensors shown in
233 Eq. (\ref{eq:intro_tensors}). If the two vectors are co-linear, e.g.,
234 $\psi=0$, $\mathbf{a}=(0,0,1)$ and $\mathbf{b}=(0,0,1)$, then
235 \begin{equation*}
236 \mathsf{G} = \frac{g_0}{3}
237 \begin{pmatrix}
238 -1 & 0 & 0 \\
239 0 & -1 & 0 \\
240 0 & 0 & 2 \\
241 \end{pmatrix} ,
242 \end{equation*}
243 which is the $a_{20}$ symmetry.
244 To generate the $a_{22}^o$ symmetry, we take:
245 $\mathbf{a}= (\frac{1}{\sqrt{2}}, \frac{1}{\sqrt{2}},0)$ and
246 $\mathbf{b}=(\frac{1}{\sqrt{2}}, -\frac{1}{\sqrt{2}},0)$
247 and find:
248 \begin{equation*}
249 \mathsf{G}=\frac{g_0}{2}
250 \begin{pmatrix}
251 1 & 0 & 0 \\
252 0 & -1 & 0 \\
253 0 & 0 & 0 \\
254 \end{pmatrix} .
255 \end{equation*}
256 To generate the $a_{22}^e$ symmetry, we take:
257 $\mathbf{a}= (1, 0, 0)$ and $\mathbf{b} = (0,1,0)$ and find:
258 \begin{equation*}
259 \mathsf{G}=\frac{g_0}{2}
260 \begin{pmatrix}
261 0 & 1 & 0 \\
262 1 & 0 & 0 \\
263 0 & 0 & 0 \\
264 \end{pmatrix} .
265 \end{equation*}
266 The pattern is straightforward to continue for the other symmetries.
267
268 We find the notation of Ref. \onlinecite{Torres-del-Castillo:2006uo}
269 helpful when creating specific types of constant gradient electric
270 fields in simulations. For this reason,
271 Eqs. (\ref{eq:appliedPotential}), (\ref{eq:GC}), and (\ref{eq:CE}) are
272 implemented in our code. In the simulations using constant applied
273 gradients that are mentioned in the main text, we utilized a field
274 with the $a_{22}^e$ symmetry using vectors, $\mathbf{a}= (1, 0, 0)$
275 and $\mathbf{b} = (0,1,0)$.
276
277 \section{Point-multipolar interactions with a spatially-varying electric field}
278
279 This section develops formulas for the force and torque exerted by an
280 external electric field, $\mathbf{E}(\mathbf{r})$, on object
281 $a$. Object $a$ has an embedded collection of charges and in
282 simulations will represent a molecule, ion, or a coarse-grained
283 substructure. We describe the charge distributions using primitive
284 multipoles defined in Ref. \onlinecite{PaperI} by
285 \begin{align}
286 C_a =&\sum_{k \, \text{in }a} q_k , \label{eq:charge} \\
287 D_{a\alpha} =&\sum_{k \, \text{in }a} q_k r_{k\alpha}, \label{eq:dipole}\\
288 Q_{a\alpha\beta} =& \frac{1}{2} \sum_{k \, \text{in } a} q_k
289 r_{k\alpha} r_{k\beta},
290 \label{eq:quadrupole}
291 \end{align}
292 where $\mathbf{r}_k$ is the local coordinate system for the object
293 (usually the center of mass of object $a$). Components of vectors and
294 tensors are given using the Einstein repeated summation notation. Note
295 that the definition of the primitive quadrupole here differs from the
296 standard traceless form, and contains an additional Taylor-series
297 based factor of $1/2$. In Ref. \onlinecite{PaperI}, we derived the
298 forces and torques each object exerts on the other objects in the
299 system.
300
301 Here we must also consider an external electric field that varies in
302 space: $\mathbf E(\mathbf r)$. Each of the local charges $q_k$ in
303 object $a$ will then experience a slightly different field. This
304 electric field can be expanded in a Taylor series around the local
305 origin of each object. For a particular charge $q_k$, the electric
306 field at that site's position is given by:
307 \begin{equation}
308 \mathbf{E}(\mathbf{r}_k) = E_\gamma|_{\mathbf{r}_k = 0} + \nabla_\delta E_\gamma |_{\mathbf{r}_k = 0} r_{k \delta}
309 + \frac {1}{2} \nabla_\delta \nabla_\varepsilon E_\gamma|_{\mathbf{r}_k = 0} r_{k \delta}
310 r_{k \varepsilon} + ...
311 \end{equation}
312 Note that if one shrinks object $a$ to a single point, the
313 ${E}_\gamma$ terms are all evaluated at the center of the object (now
314 a point). Thus later the ${E}_\gamma$ terms can be written using the
315 same (molecular) origin for all point charges in the object. The force
316 exerted on object $a$ by the electric field is given by,
317 \begin{align}
318 F^a_\gamma = \sum_{k \textrm{~in~} a} E_\gamma(\mathbf{r}_k) &= \sum_{k \textrm{~in~} a} q_k \lbrace E_\gamma + \nabla_\delta E_\gamma r_{k \delta}
319 + \frac {1}{2} \nabla_\delta \nabla_\varepsilon E_\gamma r_{k \delta}
320 r_{k \varepsilon} + ... \rbrace \\
321 &= C_a E_\gamma + D_{a \delta} \nabla_\delta E_\gamma
322 + Q_{a \delta \varepsilon} \nabla_\delta \nabla_\varepsilon E_\gamma +
323 ...
324 \end{align}
325 Thus in terms of the global origin $\mathbf{r}$, ${F}_\gamma(\mathbf{r}) = C {E}_\gamma(\mathbf{r})$ etc.
326
327 Similarly, the torque exerted by the field on $a$ can be expressed as
328 \begin{align}
329 \tau^a_\alpha &= \sum_{k \textrm{~in~} a} (\mathbf r_k \times q_k \mathbf E)_\alpha \\
330 & = \sum_{k \textrm{~in~} a} \epsilon_{\alpha \beta \gamma} q_k
331 r_{k\beta} E_\gamma(\mathbf r_k) \\
332 & = \epsilon_{\alpha \beta \gamma} D_\beta E_\gamma
333 + 2 \epsilon_{\alpha \beta \gamma} Q_{\beta \delta} \nabla_\delta
334 E_\gamma + ...
335 \end{align}
336 We note that the Levi-Civita symbol can be eliminated by utilizing the matrix cross product as defined in Ref. \onlinecite{Smith98}:
337 \begin{equation}
338 \left[\mathsf{A} \times \mathsf{B}\right]_\alpha = \sum_\beta
339 \left[\mathsf{A}_{\alpha+1,\beta} \mathsf{B}_{\alpha+2,\beta}
340 -\mathsf{A}_{\alpha+2,\beta} \mathsf{B}_{\alpha+1,\beta}
341 \right]
342 \label{eq:matrixCross}
343 \end{equation}
344 where $\alpha+1$ and $\alpha+2$ are regarded as cyclic permuations of
345 the matrix indices. Finally, the interaction energy $U^a$ of object $a$ with the external field is given by,
346 \begin{equation}
347 U^a = \sum_{k~in~a} q_k \phi_k (\mathrm{r}_k)
348 \end{equation}
349 Performing another Taylor series expansion about the local body origin,
350 \begin{equation}
351 \phi({\mathbf{r}_k}) = \phi|_{\mathbf{r}_k = 0 } + r_{k \alpha} \nabla_\alpha \phi_\alpha|_{\mathbf{r}_k = 0 } + \frac{1}{2} r_{k\alpha}r_{k\beta}\nabla_\alpha \nabla_\beta \phi|_{\mathbf{r}_k = 0} + ...
352 \end{equation}
353 Writing this in terms of the global origin $\mathrm{r}$, we find
354 \begin{equation}
355 U(\mathbf{r}) = \mathrm{C} \phi(\mathbf{r}) - \mathrm{D}_\alpha \mathrm{E}_\alpha - \mathrm{Q}_{\alpha\beta}\nabla_\alpha \mathrm{E}_\beta + ...
356 \end{equation}
357 These results have been summarized in Table \ref{tab:UFT}.
358
359 \begin{table}
360 \caption{Potential energy $(U)$, force $(\mathbf{F})$, and torque
361 $(\mathbf{\tau})$ expressions for a multipolar site at $\mathbf{r}$ in an
362 electric field, $\mathbf{E}(\mathbf{r})$ using the definitions of the multipoles in Eqs. (\ref{eq:charge}), (\ref{eq:dipole}) and (\ref{eq:quadrupole}).
363 \label{tab:UFT}}
364 \begin{tabular}{r|C{3cm}C{3cm}C{3cm}}
365 & Charge & Dipole & Quadrupole \\ \hline
366 $U(\mathbf{r})$ & $C \phi(\mathbf{r})$ & $-\mathbf{D} \cdot \mathbf{E}(\mathbf{r})$ & $- \mathsf{Q}:\nabla \mathbf{E}(\mathbf{r})$ \\
367 $\mathbf{F}(\mathbf{r})$ & $C \mathbf{E}(\mathbf{r})$ & $\mathbf{D} \cdot \nabla \mathbf{E}(\mathbf{r})$ & $\mathsf{Q} : \nabla\nabla\mathbf{E}(\mathbf{r})$ \\
368 $\mathbf{\tau}(\mathbf{r})$ & & $\mathbf{D} \times \mathbf{E}(\mathbf{r})$ & $2 \mathsf{Q} \times \nabla \mathbf{E}(\mathbf{r})$
369 \end{tabular}
370 \end{table}
371
372 \section{Boltzmann averages for orientational polarization}
373 The dielectric properties of the system mainly arise from two
374 different ways: i) the applied field distort the charge distributions
375 so it produces an induced multipolar moment in each molecule; and ii)
376 the applied field tends to line up originally randomly oriented
377 molecular moment towards the direction of the applied field. In this
378 study, we basically focus on the orientational contribution in the
379 dielectric properties. If we consider a system of molecules in the
380 presence of external field perturbation, the perturbation experienced
381 by any molecule will not be only due to external field or field
382 gradient but also due to the field or field gradient produced by the
383 all other molecules in the system. In the following subsections
384 \ref{subsec:boltzAverage-Dipole} and \ref{subsec:boltzAverage-Quad},
385 we will discuss about the molecular polarization only due to external
386 field perturbation. The contribution of the field or field gradient
387 due to all other molecules will be taken into account while
388 calculating correction factor in the paper.
389
390 \subsection{Dipoles}
391 \label{subsec:boltzAverage-Dipole}
392 Consider a system of molecules, each with permanent dipole moment
393 $p_o$. In the absense of external field, thermal agitation orients the
394 dipoles randomly, reducing the system moment to zero. External fields
395 will tend to line up the dipoles in the direction of applied field.
396 Here we have considered net field from all other molecules is
397 considered to be zero. Therefore the total Hamiltonian of each
398 molecule is,\cite{Jackson98}
399 \begin{equation}
400 H = H_o - \mathbf{p_o}\cdot \mathbf{E},
401 \end{equation}
402 where $H_o$ is a function of the internal coordinates of the molecule.
403 The Boltzmann average of the dipole moment is given by,
404 \begin{equation}
405 \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}}},
406 \end{equation}
407 where $\bf{E}$ is selected along z-axis. If we consider that the
408 applied field is small, \textit{i.e.} $\frac{p_oE\; cos\theta}{k_B T} << 1$,
409 \begin{equation}
410 \braket{p_{mol}} \approx \frac{1}{3}\frac{{p_o}^2}{k_B T}E,
411 \end{equation}
412 where $ \alpha_p = \frac{1}{3}\frac{{p_o}^2}{k_B T}$ is a molecular
413 polarizability. The orientational polarization depends inversely on
414 the temperature and applied field must overcome the thermal agitation.
415
416 \subsection{Quadrupoles}
417 \label{subsec:boltzAverage-Quad}
418 Consider a system of molecules with permanent quadrupole moment
419 $q_{\alpha\beta}$. The average quadrupole moment at temperature T in
420 the presence of uniform applied field gradient is given
421 by,\cite{AduGyamfi78, AduGyamfi81}
422 \begin{equation}
423 \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}}},
424 \label{boltzQuad}
425 \end{equation}
426 where $\int d\Omega = \int_0^{2\pi} \int_0^\pi \int_0^{2\pi}
427 sin\theta\; d\theta\ d\phi\ d\psi$ is the integration over Euler
428 angles, $ H = H_o -q_{\mu\nu}\;\partial_\nu E_\mu $ is the energy of
429 a quadrupole in the gradient of the
430 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,
431 \begin{equation}
432 \begin{split}
433 &q_{\alpha\beta} = \eta_{\alpha\alpha'}\;\eta_{\beta\beta'}\;{q}^* _{\alpha'\beta'} \\
434 &H = H_o - q:{\nabla}\mathbf{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.
435 \end{split}
436 \label{energyQuad}
437 \end{equation}
438 Here the starred tensors are the components in the body fixed
439 frame. Substituting equation (\ref{energyQuad}) in the equation (\ref{boltzQuad})
440 and taking linear terms in the expansion we get,
441 \begin{equation}
442 \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)},
443 \end{equation}
444 where $\eta_{\alpha\alpha'}$ is the inverse of the rotation matrix that transforms
445 the body fixed co-ordinates to the space co-ordinates.
446 % \[\eta_{\alpha\alpha'}
447 % = \left(\begin{array}{ccc}
448 % cos\phi\; cos\psi - cos\theta\; sin\phi\; sin\psi & -cos\theta\; cos\psi\; sin\phi - cos\phi\; sin\psi & sin\theta\; sin\phi \\
449 % cos\psi\; sin\phi + cos\theta\; cos\phi \; sin\psi & cos\theta\; cos\phi\; cos\psi - sin\phi\; sin\psi & -cos\phi\; sin\theta \\
450 % sin\theta\; sin\psi & -cos\psi\; sin\theta & cos\theta
451 % \end{array} \right).\]
452
453 Integration of 1st and 2nd terms in the denominator gives $8 \pi^2$
454 and $8 \pi^2 /3\;{\nabla}.\mathbf{E}\; Tr(q^*) $ respectively. The
455 second term vanishes for charge free space, ${\nabla}.\mathbf{E} \; = \; 0$. Similarly integration of the
456 1st term in the numerator produces
457 $8 \pi^2 /3\; Tr(q^*)\delta_{\alpha\beta}$ and the 2nd term produces
458 $8 \pi^2 /15k_B T (3{q}^*_{\alpha'\beta'}{q}^*_{\beta'\alpha'} -
459 {q}^*_{\alpha'\alpha'}{q}^*_{\beta'\beta'})\partial_\alpha E_\beta$,
460 if ${\nabla}.\mathbf{E} \; = \; 0$,
461 $ \partial_\alpha E_\beta = \partial_\beta E_\alpha$ and
462 ${q}^*_{\alpha'\beta'}= {q}^*_{\beta'\alpha'}$. Therefore the
463 Boltzmann average of a quadrupole moment can be written as,
464
465 \begin{equation}
466 \braket{q_{\alpha\beta}}\; = \; \frac{1}{3} Tr(q^*)\;\delta_{\alpha\beta} + \frac{{\bar{q_o}}^2}{15k_BT}\;\partial_\alpha E_\beta,
467 \end{equation}
468 where $ \alpha_q = \frac{{\bar{q_o}}^2}{15k_BT} $ is a molecular quadrupole polarizablity and ${\bar{q_o}}^2=
469 3{q}^*_{\alpha'\beta'}{q}^*_{\beta'\alpha'}-{q}^*_{\alpha'\alpha'}{q}^*_{\beta'\beta'}$ is a square of the net quadrupole moment of a molecule.
470
471 \section{Gradient of the field due to quadrupolar polarization}
472 \label{singularQuad}
473 In this section, we will discuss the gradient of the field produced by
474 quadrupolar polarization. For this purpose, we consider a distribution
475 of charge ${\rho}(\mathbf r)$ which gives rise to an electric field
476 $\mathbf{E}(\mathbf r)$ and gradient of the field ${\nabla} \mathbf{E}(\mathbf r)$
477 throughout space. The total gradient of the electric field over volume
478 due to the all charges within the sphere of radius $R$ is given by
479 (cf. Jackson equation 4.14):
480 \begin{equation}
481 \int_{r<R} {\nabla}\mathbf{E}\;d^3r = -\int_{r=R} R^2 \mathbf{E}\;\hat{n}\; d\Omega
482 \label{eq:8}
483 \end{equation}
484 where $d\Omega$ is the solid angle and $\hat{n}$ is the normal vector
485 of the surface of the sphere,
486 $\hat{n} = sin[\theta]cos[\phi]\hat{x} + sin[\theta]sin[\phi]\hat{y} +
487 cos[\theta]\hat{z}$
488 in spherical coordinates. For the charge density ${\rho}(\mathbf r')$, the
489 total gradient of the electric field can be written as, ~\cite{Jackson98}
490 \begin{equation}
491 \int_{r<R} {\nabla}\mathbf {E}\; d^3r=-\int_{r=R} R^2\; {\nabla}\Phi\; \hat{n}\; d\Omega =-\frac{1}{4\pi\;\epsilon_o}\int_{r=R} R^2\; {\nabla}\;\left(\int \frac{\rho(\mathbf r')}{|\mathbf{r}-\mathbf{r}'|}\;d^3r'\right) \hat{n}\; d\Omega
492 \label{eq:9}
493 \end{equation}
494 The radial function in the equation (\ref{eq:9}) can be expressed in
495 terms of spherical harmonics as,\cite{Jackson98}
496 \begin{equation}
497 \frac{1}{|\mathbf{r} - \mathbf{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)
498 \label{eq:10}
499 \end{equation}
500 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,
501 \begin{equation}
502 \begin{split}
503 \int_{r<R} {\nabla}\mathbf{E}\;d^3r &=-\frac{R^2}{\epsilon_o}\int_{r=R} \; {\nabla}\;\left(\int \rho(\mathbf 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 \\
504 &= -\frac{R^2}{\epsilon_o}\sum_{l=0}^{\infty}\sum_{m=-l}^{m=l}\frac{1}{2l+1}\;\int \rho(\mathbf 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
505 '
506 \end{split}
507 \label{eq:11}
508 \end{equation}
509 The gradient of the product of radial function and spherical harmonics
510 is given by:\cite{Arfkan}
511 \begin{equation}
512 \begin{split}
513 {\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
514 {\partial}{\partial r}+\frac{l}{r} \right]f(r)\; Y_{l, l-1, m}(\theta, \phi).
515 \end{split}
516 \label{eq:12}
517 \end{equation}
518 Using equation (\ref{eq:12}) we get,
519 \begin{equation}
520 {\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}},
521 \label{eq:13}
522 \end{equation}
523 where $ Y_{l,l+1,m}(\theta, \phi)$ is a vector spherical harmonics \cite{Arfkan}. Using Clebsch-Gorden coefficients $C(l+1, 1, l|m_1,m_2,m) $, equation \ref{eq:14} can be written in spherical harmonics,
524 \begin{equation}
525 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}.
526 \label{eq:14}
527 \end{equation}
528 Here $\hat{e}_{m_2}$ is a spherical tensor of rank 1 which can be expressed
529 in terms of Cartesian coordinates,
530 \begin{equation}
531 {\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}}.
532 \label{eq:15}
533 \end{equation}
534 The normal vector $\hat{n} $ is then expressed in terms of spherical tensor of rank 1 as shown in below,
535 \begin{equation}
536 \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).
537 \label{eq:16}
538 \end{equation}
539 The surface integral of the product of $\hat{n}$ and
540 ${Y_{l+1}}^{m_1}(\theta, \phi)$ gives,
541 \begin{equation}
542 \begin{split}
543 \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 \\
544 &= \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 \\
545 &= \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),
546 \end{split}
547 \label{eq:17}
548 \end{equation}
549 where ${Y_{l}}^{-m} = (-1)^m\;{{Y_{l}}^{m}}^* $ and
550 $ \int {{Y_{l}}^{m}}^*\;{Y_{l'}}^{m'}\;d\Omega =
551 \delta_{ll'}\delta_{mm'} $.
552 Non-vanishing values of equation \ref{eq:17} require $l = 0$,
553 therefore the value of $ m = 0 $. Since the values of $ m_1$ are -1,
554 1, and 0 then $m_2$ takes the values 1, -1, and 0, respectively
555 provided that $m = m_1 + m_2$. Equation \ref{eq:11} can therefore be
556 modified,
557 \begin{equation}
558 \begin{split}
559 \int_{r<R} {\nabla}\mathbf{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(
560 1, 1, 0|0,0,0)\;{\hat{e}_{0}}{\hat{e}_{0}} ]\; d^3r' \\
561 &= -\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)\\
562 &= - \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).
563 \end{split}
564 \label{eq:19}
565 \end{equation}
566 In the last step, the charge density was integrated over the sphere, yielding a total charge $\mathrm{C_total}$.Equation (\ref{eq:19}) gives the total gradient of the field over a sphere due to the distribution of the charges.
567 For quadrupolar fluids the total charge within a sphere is zero, therefore
568 $ \int_{r<R} {\nabla}\mathbf{E}\;d^3r = 0 $. Hence the quadrupolar
569 polarization produces zero net gradient of the field inside the
570 sphere.
571
572 \section{Geometric Factors for Two Embedded Point Charges}
573
574 \bibliography{dielectric_new}
575 \end{document}
576 %
577 % ****** End of file multipole.tex ******