ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/multipole/Dielectric_Supplemental.tex
(Generate patch)

Comparing trunk/multipole/Dielectric_Supplemental.tex (file contents):
Revision 4409 by mlamichh, Wed Mar 30 21:09:54 2016 UTC vs.
Revision 4419 by gezelter, Tue Apr 12 20:03:08 2016 UTC

# Line 36 | Line 36 | jcp]{revtex4-1}
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[Real space electrostatics for multipoles. III. Dielectric Properties]
52 < {Supplemental Material for: Real space electrostatics for multipoles. III. Dielectric Properties}
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
# Line 63 | Line 68 | of Notre Dame, Notre Dame, IN 46556}
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 density 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 < \newpage
84 > \section{Generating Uniform Field Gradients}
85 > One important task in carrying out the simulations mentioned in the
86 > main text was to generate uniform electric field gradients.  To do
87 > this, we relied heavily on both the notation and results from Torres
88 > del Castillo and Mend\'{e}z Garido.\cite{Torres-del-Castillo:2006uo}
89 > In this work, tensors were expressed in Cartesian components, using at
90 > times a dyadic notation. This proves quite useful for computer
91 > simulations 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 < \section{Boltzmann averages for orientational polarization}
101 < The dielectric properties of the system is mainly arise from two
102 < different ways: i) the applied field distort the charge distributions
103 < so it produces an induced multipolar moment in each molecule; and ii)
104 < the applied field tends to line up originally randomly oriented
105 < molecular moment towards the direction of the applied field. In this
106 < study, we basically focus on the orientational contribution in the
107 < dielectric properties. If we consider a system of molecules in the
108 < presence of external field perturbation, the perturbation experienced
80 < by any molecule will not be only due to external field or field
81 < gradient but also due to the field or field gradient produced by the
82 < all other molecules in the system. In the following subsections
83 < \ref{subsec:boltzAverage-Dipole} and \ref{subsec:boltzAverage-Quad},
84 < we will discuss about the molecular polarization only due to external
85 < field perturbation. The contribution of the field or field gradient
86 < due to all other molecules will be taken into account while
87 < calculating correction factor in the paper.
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 < \subsection{Dipoles}
111 < \label{subsec:boltzAverage-Dipole}
112 < Consider a system of molecules, each with permanent dipole moment
113 < $p_o$. In the absense of external field, thermal agitation orients the
114 < dipoles randomly, reducing the system moment to zero.  External fields
115 < will tend to line up the dipoles in the direction of applied field.
116 < Here we have considered net field from all other molecules is
117 < considered to be zero.  Therefore the total Hamiltonian of each
118 < molecule is,\cite{Jackson98}
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 < H = H_o - \bf{p_o}\cdot \mathbf{E},
136 < \end{equation}
137 < where $H_o$ is a function of the internal coordinates of the molecule.
138 < The Boltzmann average of the dipole moment is given by,
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 < \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}}},
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 < where $\bf{E}$ is selected along z-axis. If we consider that the
167 < applied field is small, \textit{i.e.} $\frac{p_oE\; cos\theta}{k_B T} << 1$,
168 < \begin{equation}
169 < \braket{p_{mol}}  \approx \frac{1}{3}\frac{{p_o}^2}{k_B T}E,
110 < \end{equation}
111 < where $ \alpha_p = \frac{1}{3}\frac{{p_o}^2}{k_B T}$ is a molecular
112 < polarizability. The orientational polarization depends inversely on
113 < the temperature and applied field must overcome the thermal agitation.
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 < \subsection{Quadrupoles}
172 < \label{subsec:boltzAverage-Quad}
117 < Consider a system of molecules with permanent quadrupole moment
118 < $q_{\alpha\beta}$. The average quadrupole moment at temperature T in
119 < the presence of uniform applied field gradient is given
120 < by,\cite{AduGyamfi78, AduGyamfi81}
171 > Applying the gradient operator to Eq. (\ref{eq:quad_phi}) the electric
172 > field due to this potential,
173   \begin{equation}
174 < \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}}},
175 < \label{boltzQuad}
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 < where $\int d\Omega = \int_0^{2\pi} \int_0^\pi \int_0^{2\pi}
183 < sin\theta\; d\theta\ d\phi\ d\psi$ is the integration over Euler
184 < angles, $ H = H_o -q_{\mu\nu}\;\partial_\nu E_\mu $ is the energy of
185 < a quadrupole in the gradient of the  
186 < 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,
187 < \begin{equation}
188 < \begin{split}
189 < &q_{\alpha\beta} = \eta_{\alpha\alpha'}\;\eta_{\beta\beta'}\;{q}^* _{\alpha'\beta'} \\
133 < &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.
134 < \end{split}
135 < \label{energyQuad}
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 < Here the starred tensors are the components in the body fixed
192 < frame. Substituting equation (\ref{energyQuad}) in the equation (\ref{boltzQuad})
193 < and taking linear terms in the expansion we get,
194 < \begin{equation}
195 < \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)},
196 < \end{equation}
197 < where $\eta_{\alpha\alpha'}$ is the inverse of the rotation matrix that transforms
198 < the body fixed co-ordinates to the space co-ordinates,
199 < \[\eta_{\alpha\alpha'}
200 < = \left(\begin{array}{ccc}
201 < cos\phi\; cos\psi - cos\theta\; sin\phi\; sin\psi & -cos\theta\; cos\psi\; sin\phi - cos\phi\; sin\psi & sin\theta\; sin\phi \\
202 < cos\psi\; sin\phi + cos\theta\; cos\phi \; sin\psi & cos\theta\; cos\phi\; cos\psi - sin\phi\; sin\psi & -cos\phi\; sin\theta \\
203 < sin\theta\; sin\psi & -cos\psi\; sin\theta & cos\theta
204 < \end{array} \right).\]
205 < Integration of 1st and 2nd terms in the denominator gives $8 \pi^2$
206 < and $8 \pi^2 /3\;{\nabla}.\mathbf{E}\; Tr(q^*) $ respectively. The
207 < second term vanishes for charge free space, ${\nabla}.\mathbf{E} \; = \; 0$. Similarly integration of the
208 < 1st term in the numerator produces
209 < $8 \pi^2 /3\; Tr(q^*)\delta_{\alpha\beta}$ and the 2nd term produces
210 < $8 \pi^2 /15k_B T (3{q}^*_{\alpha'\beta'}{q}^*_{\beta'\alpha'} -
211 < {q}^*_{\alpha'\alpha'}{q}^*_{\beta'\beta'})\partial_\alpha E_\beta$,
212 < if ${\nabla}.\mathbf{E} \; = \; 0$,
213 < $ \partial_\alpha E_\beta = \partial_\beta E_\alpha$ and
214 < ${q}^*_{\alpha'\beta'}= {q}^*_{\beta'\alpha'}$. Therefore the
215 < Boltzmann average of a quadrupole moment can be written as,
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 < \begin{equation}
232 < \braket{q_{\alpha\beta}}\; = \; \frac{1}{3} Tr(q^*)\;\delta_{\alpha\beta} + \frac{{\bar{q_o}}^2}{15k_BT}\;\partial_\alpha E_\beta,
233 < \end{equation}
234 < where $ \alpha_q = \frac{{\bar{q_o}}^2}{15k_BT} $ is a molecular quadrupole polarizablity  and  ${\bar{q_o}}^2=
235 < 3{q}^*_{\alpha'\beta'}{q}^*_{\beta'\alpha'}-{q}^*_{\alpha'\alpha'}{q}^*_{\beta'\beta'}$ is a square of the net quadrupole moment of a molecule.
236 <
237 < \section{External application of a uniform field gradient}
238 < \label{Ap:fieldOrGradient}
239 <
240 < To satisfy the condition $ \nabla \cdot \mathbf{E} = 0 $, within the box of molecules we have taken electrostatic potential in the following form
241 < \begin{equation}
242 < \begin{split}
243 < \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. \\
244 < & \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),
245 < \end{split}
246 < \label{eq:appliedPotential}
247 < \end{equation}
248 < 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,
249 < \[\mathbf{E}
250 < = \frac{g_o}{2} \left(\begin{array}{ccc}
251 < 2(a_1\; b_1 - \frac{cos\psi}{3})\;x \;+  (a_1\; b_2 \;+ a_2\; b_1)\;y + (a_1\; b_3 \;+ a_3\; b_1)\;z \\
252 < (a_2\; b_1 \;+ a_1\; b_2)\;x + 2(a_2\; b_2 \;- \frac{cos\psi}{3})\;y +  (a_2\; b_3 \;+ a_3\; b_2)\;z \\
253 < (a_3\; b_1 \;+ a_3\; b_2)\;x +  (a_3\; b_2 \;+ a_2\; b_3)y + 2(a_3\; b_3 \;- \frac{cos\psi}{3})\;z
254 < \end{array} \right).\]
255 < The gradient of the applied field derived from the potential can be written in the following form,
256 < \[\nabla\mathbf{E}
257 < = \frac{g_o}{2}\left(\begin{array}{ccc}
258 < 2(a_1\; b_1 - \frac{cos\psi}{3}) &  (a_1\; b_2 \;+ a_2\; b_1) & (a_1\; b_3 \;+ a_3\; b_1) \\
259 < (a_2\; b_1 \;+ a_1\; b_2) & 2(a_2\; b_2 \;- \frac{cos\psi}{3}) & (a_2\; b_3 \;+ a_3\; b_2) \\
260 < (a_3\; b_1 \;+ a_3\; b_2) & (a_3\; b_2 \;+ a_2\; b_3) & 2(a_3\; b_3 \;- \frac{cos\psi}{3})
261 < \end{array} \right).\]
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 < We want to derive formulas for the force and torque exerted by an external electric field $\mathbf{E}(\mathbf{r})$ on object $a$. Object a has an embedded collection of charges and in simulations will normally represent a molecule or ion. We describe the charge distributions using primitive monopoles defined in paper I by
280 <
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}\\
# Line 204 | Line 289 | where $\mathbf{r}_k$ is the local coordinate system fo
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 (for convenience, the real origin is at the "center" of object $a$). Components of vectors and tensors are given using Green indices, using the Einstein repeated summation notation.Note that the definition of the primitive quadrupole here differs from the standard traceless form, and contains an additional Taylor-series based factor of $1/2$.  In Paper I \cite{PaperI}, we derived the forces and torques each object exerts on the other objects in the system.
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.  
306 < For a particular charge $q_k$, the electric field at that site's
215 < position is given by:
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 once one shrinks object $a$ to point size, the ${E}_\gamma$ terms  are all evaluated at the center of the object (now a point). Thus later the ${E}_\gamma$ terms can be written using the same global origin for all objects $a, b, c, ...$ in the system. The force exerted on object $a$ by the electric field is given by,
313 <
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,\cite{Raab:2004ve}
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}
# Line 228 | Line 322 | Thus in terms of the global origin $\mathbf{r}$, ${F}_
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.
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}
# Line 260 | Line 354 | The results has been summarized in Table I.
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 < The results has been summarized in Table I.
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 $\mathrm{r}$ in an
362 <  electric field, $\mathbf{E}(\mathbf{r})$.
363 < \label{tab:UFT}}
364 < \begin{tabular}{r|ccc}
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 + If we consider a collection of molecules in the presence of external
374 + field, the perturbation experienced by any one molecule will include
375 + contributions to the field or field gradient produced by the all other
376 + molecules in the system. In subsections
377 + \ref{subsec:boltzAverage-Dipole} and \ref{subsec:boltzAverage-Quad},
378 + we discuss the molecular polarization due solely to external field
379 + perturbations.  This illustrates the origins of the polarizability
380 + equations (Eqs. 6, 20, and 21) in the main text.
381 +
382 + \subsection{Dipoles}
383 + \label{subsec:boltzAverage-Dipole}
384 + Consider a system of molecules, each with permanent dipole moment
385 + $p_o$. In the absense of an external field, thermal agitation orients
386 + the dipoles randomly, and the system moment, $\mathbf{P}$, is zero.
387 + External fields will line up the dipoles in the direction of applied
388 + field.  Here we consider the net field from all other molecules to be
389 + zero.  Therefore the total Hamiltonian acting on each molecule
390 + is,\cite{Jackson98}
391 + \begin{equation}
392 + H = H_o - \mathbf{p}_o \cdot \mathbf{E},
393 + \end{equation}
394 + where $H_o$ is a function of the internal coordinates of the molecule.
395 + The Boltzmann average of the dipole moment in the direction of the
396 + field is given by,
397 + \begin{equation}
398 + \langle p_{mol} \rangle = \frac{\displaystyle\int p_o \cos\theta
399 +  e^{~p_o E \cos\theta /k_B T}\; d\Omega}{\displaystyle\int  e^{~p_o E \cos\theta/k_B
400 +    T}\; d\Omega},
401 + \end{equation}
402 + where the $z$-axis is taken in the direction of the applied field,
403 + $\bf{E}$ and
404 + $\int d\Omega = \int_0^\pi \sin\theta\; d\theta \int_0^{2\pi} d\phi
405 + \int_0^{2\pi} d\psi$
406 + is an integration over Euler angles describing the orientation of the
407 + molecule.
408 +
409 + If the external fields are small, \textit{i.e.}
410 + $p_oE \cos\theta / k_B T << 1$,
411 + \begin{equation}
412 + \langle p_{mol} \rangle \approx \frac{{p_o}^2}{3 k_B T}E,
413 + \end{equation}
414 + where $ \alpha_p = \frac{{p_o}^2}{3 k_B T}$ is the molecular
415 + polarizability. The orientational polarization depends inversely on
416 + the temperature as the applied field must overcome thermal agitation
417 + to orient the dipoles.
418 +
419 + \subsection{Quadrupoles}
420 + \label{subsec:boltzAverage-Quad}
421 + If instead, our system consists of molecules with permanent
422 + \textit{quadrupole} tensor $q_{\alpha\beta}$. The average quadrupole
423 + at temperature $T$ in the presence of uniform applied field gradient
424 + is given by,\cite{AduGyamfi78, AduGyamfi81}
425 + \begin{equation}
426 + \langle q_{\alpha\beta} \rangle \;=\; \frac{\displaystyle\int
427 +  q_{\alpha\beta}\; e^{-H/k_B T}\; d\Omega}{\displaystyle\int
428 +  e^{-H/k_B T}\; d\Omega} \;=\; \frac{\displaystyle\int
429 +  q_{\alpha\beta}\; e^{~q_{\mu\nu}\;\partial_\nu E_\mu /k_B T}\;
430 +  d\Omega}{\displaystyle\int  e^{~q_{\mu\nu}\;\partial_\nu E_\mu /k_B
431 +    T}\; d\Omega },
432 + \label{boltzQuad}
433 + \end{equation}
434 + where $H = H_o - q_{\mu\nu}\;\partial_\nu E_\mu $ is the energy of a
435 + quadrupole in the gradient of the applied field and $H_o$ is a
436 + function of internal coordinates of the molecule. The energy and
437 + quadrupole moment can be transformed into the body frame using a
438 + rotation matrix $\mathsf{\eta}^{-1}$,
439 + \begin{align}
440 + q_{\alpha\beta} &= \eta_{\alpha\alpha'}\;\eta_{\beta\beta'}\;{q}^* _{\alpha'\beta'} \\
441 + H &= H_o - q:{\nabla}\mathbf{E} \\
442 +  &= H_o - q_{\mu\nu}\;\partial_\nu E_\mu  \\
443 +  &= H_o
444 +    -\eta_{\mu\mu'}\;\eta_{\nu\nu'}\;{q}^*_{\mu'\nu'}\;\partial_\nu
445 +    E_\mu. \label{energyQuad}
446 + \end{align}
447 + Here the starred tensors are the components in the body fixed
448 + frame. Substituting equation (\ref{energyQuad}) in the equation
449 + (\ref{boltzQuad}) and taking linear terms in the expansion we obtain,
450 + \begin{equation}
451 + \braket{q_{\alpha\beta}} = \frac{\displaystyle \int q_{\alpha\beta} \left(1 +
452 +    \frac{\eta_{\mu\mu'}\;\eta_{\nu\nu'}\;{q}^*_{\mu'\nu'}\;\partial_\nu
453 +      E_\mu }{k_B T}\right)\;  d\Omega}{\displaystyle \int \left(1 + \frac{\eta_{\mu\mu'}\;\eta_{\nu\nu'}\;{q}^*_{\mu'\nu'}\;\partial_\nu E_\mu }{k_B T}\right)\; d\Omega}.
454 + \end{equation}
455 + Recall that $\eta_{\alpha\alpha'}$ is the inverse of the rotation
456 + matrix that transforms the body fixed co-ordinates to the space
457 + co-ordinates.
458 + % \[\eta_{\alpha\alpha'}
459 + % = \left(\begin{array}{ccc}
460 + % cos\phi\; cos\psi - cos\theta\; sin\phi\; sin\psi & -cos\theta\; cos\psi\; sin\phi - cos\phi\; sin\psi & sin\theta\; sin\phi \\
461 + % cos\psi\; sin\phi + cos\theta\; cos\phi \; sin\psi & cos\theta\; cos\phi\; cos\psi - sin\phi\; sin\psi & -cos\phi\; sin\theta \\
462 + % sin\theta\; sin\psi & -cos\psi\; sin\theta & cos\theta
463 + % \end{array} \right).\]
464  
465 + Integration of the first and second terms in the denominator gives
466 + $8 \pi^2$ and
467 + $8 \pi^2 ({\nabla} \cdot \mathbf{E}) \mathrm{Tr}(q^*) / 3 $
468 + respectively. The second term vanishes for charge free space (where
469 + ${\nabla} \cdot \mathbf{E}=0$). Similarly, integration of the first
470 + term in the numerator produces
471 + $8 \pi^2 \delta_{\alpha\beta} \mathrm{Tr}(q^*) / 3$ while the second
472 + produces
473 + $8 \pi^2 (3{q}^*_{\alpha'\beta'}{q}^*_{\beta'\alpha'} -
474 + {q}^*_{\alpha'\alpha'}{q}^*_{\beta'\beta'})\partial_\alpha E_\beta /
475 + 15 k_B T $.
476 + Therefore the Boltzmann average of a quadrupole moment can be written
477 + as,
478 + \begin{equation}
479 + \langle q_{\alpha\beta} \rangle =  \frac{1}{3} \mathrm{Tr}(q^*)\;\delta_{\alpha\beta} + \frac{{\bar{q_o}}^2}{15k_BT}\;\partial_\alpha E_\beta,
480 + \end{equation}
481 + where $\alpha_q = \frac{{\bar{q_o}}^2}{15k_BT} $ is a molecular
482 + quadrupole polarizablity and
483 + ${\bar{q_o}}^2=
484 + 3{q}^*_{\alpha'\beta'}{q}^*_{\beta'\alpha'}-{q}^*_{\alpha'\alpha'}{q}^*_{\beta'\beta'}$
485 + is the square of the net quadrupole moment of a molecule.
486 +
487   \section{Gradient of the field due to quadrupolar polarization}
488   \label{singularQuad}
489 < In this section, we will discuss the gradient of the field produced by
490 < quadrupolar polarization. For this purpose, we consider a distribution
491 < of charge ${\rho}(\mathbf r)$ which gives rise to an electric field
492 < $\mathbf{E}(\mathbf r)$ and gradient of the field ${\nabla} \mathbf{E}(\mathbf r)$
493 < throughout space. The total gradient of the electric field over volume
494 < due to the all charges within the sphere of radius $R$ is given by
495 < (cf. Jackson equation 4.14):
489 > In section IV.C of the main text, we stated that for quadrupolar
490 > fluids, the self-contribution to the field gradient vanishes at the
491 > singularity. In this section, we prove this statement.  For this
492 > purpose, we consider a distribution of charge $\rho(\mathbf{r})$ which
493 > gives rise to an electric field $\mathbf{E}(\mathbf{r})$ and gradient
494 > of the field $\nabla\mathbf{E}(\mathbf{r})$ throughout space. The
495 > gradient of the electric field over volume due to the charges within
496 > the sphere of radius $R$ is given by (cf. Ref. \onlinecite{Jackson98},
497 > equation 4.14):
498   \begin{equation}
499 < \int_{r<R} {\nabla}\mathbf{E}\;d^3r = -\int_{r=R} R^2 \mathbf{E}\;\hat{n}\; d\Omega
499 > \int_{r<R} \nabla\mathbf{E} d\mathbf{r} = -\int_{r=R} R^2 \mathbf{E}\;\hat{n}\; d\Omega
500   \label{eq:8}
501   \end{equation}
502   where $d\Omega$ is the solid angle and $\hat{n}$ is the normal vector
503   of the surface of the sphere,
293 $\hat{n} = sin[\theta]cos[\phi]\hat{x} + sin[\theta]sin[\phi]\hat{y} +
294 cos[\theta]\hat{z}$
295 in spherical coordinates.  For the charge density ${\rho}(\mathbf r')$, the
296 total gradient of the electric field can be written as, ~\cite{Jackson98}
504   \begin{equation}
505 < \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
505 > \hat{n} = \sin\theta\cos\phi\; \hat{x} + \sin\theta\sin\phi\; \hat{y} +
506 > \cos\theta\; \hat{z}
507 > \end{equation}
508 > in spherical coordinates.  For the charge density $\rho(\mathbf{r}')$, the
509 > total gradient of the electric field can be written as,\cite{Jackson98}
510 > \begin{equation}
511 > \int_{r<R} {\nabla}\mathbf {E}\; d\mathbf{r}=-\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\mathbf{r}'\right) \hat{n}\; d\Omega
512   \label{eq:9}
513   \end{equation}
514   The radial function in the equation (\ref{eq:9}) can be expressed in
# Line 307 | Line 520 | If the sphere completely encloses the charge density t
520   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,
521   \begin{equation}
522   \begin{split}
523 < \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 \\
524 < &= -\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
523 > \int_{r<R} {\nabla}\mathbf{E}\;d\mathbf{r} &=-\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\mathbf{r}'\right) \hat{n}\; d\Omega \\
524 > &= -\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\mathbf{r}
525   '
526   \end{split}
527   \label{eq:11}
# Line 322 | Line 535 | Using equation (\ref{eq:12}) we get,
535   \end{split}
536   \label{eq:12}
537   \end{equation}
538 < Using equation (\ref{eq:12}) we get,
538 > where $Y_{l,l+1,m}(\theta, \phi)$ is a vector spherical
539 > harmonic.\cite{Arfkan} Using equation (\ref{eq:12}) we get,
540   \begin{equation}
541   {\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}},
542   \label{eq:13}
543   \end{equation}
544 < 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,
544 > Using Clebsch-Gordan coefficients $C(l+1,1,l|m_1,m_2,m)$, the vector
545 > spherical harmonics can be written in terms of spherical harmonics,
546   \begin{equation}
547 < 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}.
547 > 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}.
548   \label{eq:14}
549   \end{equation}
550   Here $\hat{e}_{m_2}$ is a spherical tensor of rank 1 which can be expressed
# Line 340 | Line 555 | The normal vector $\hat{n} $ is then expressed in term
555   \end{equation}
556   The normal vector $\hat{n} $ is then expressed in terms of spherical tensor of rank 1 as shown in below,
557   \begin{equation}
558 < \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).
558 > \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).
559   \label{eq:16}
560   \end{equation}
561   The surface integral of the product of $\hat{n}$ and
562 < ${Y_{l+1}}^{m_1}(\theta, \phi)$ gives,
562 > $Y_{l+1}^{m_1}(\theta, \phi)$ gives,
563   \begin{equation}
564   \begin{split}
565 < \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 \\
566 < &=  \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 \\
565 > \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 \\
566 > &=  \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 \\
567   &=   \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),
568   \end{split}
569   \label{eq:17}
570   \end{equation}
571 < where ${Y_{l}}^{-m} = (-1)^m\;{{Y_{l}}^{m}}^* $ and
572 < $ \int {{Y_{l}}^{m}}^*\;{Y_{l'}}^{m'}\;d\Omega =
571 > where $Y_{l}^{-m} = (-1)^m\;{Y_{l}^{m}}^* $ and
572 > $ \int {Y_{l}^{m}}^* Y_{l'}^{m'}\;d\Omega =
573   \delta_{ll'}\delta_{mm'} $.
574   Non-vanishing values of equation \ref{eq:17} require $l = 0$,
575   therefore the value of $ m = 0 $. Since the values of $ m_1$ are -1,
# Line 363 | Line 578 | modified,
578   modified,
579   \begin{equation}
580   \begin{split}
581 < \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(
582 < 1, 1, 0|0,0,0)\;{\hat{e}_{0}}{\hat{e}_{0}} ]\; d^3r' \\
583 < &= -\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)\\
584 < &= - \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).
581 > \int_{r<R} {\nabla}\mathbf{E}\;d\mathbf{r} = &- \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(
582 > 1, 1, 0|0,0,0)\;{\hat{e}_{0}}{\hat{e}_{0}} ]\; d\mathbf{r}' \\
583 > &= -\sqrt{\frac{4\pi}{{3}}}\;\frac{1}{\epsilon_o}\int \rho(r')\;d\mathbf{r}'\left({\hat{e}_{-1}}{\hat{e}_{1}}+{\hat{e}_{1}}{\hat{e}_{-1}}-{\hat{e}_{0}}{\hat{e}_{0}}\right)\\
584 > &= - \sqrt{\frac{4\pi}{{3}}}\;\frac{1}{\epsilon_o}\;C_\mathrm{total}\;\left({\hat{e}_{-1}}{\hat{e}_{1}}+{\hat{e}_{1}}{\hat{e}_{-1}}-{\hat{e}_{0}}{\hat{e}_{0}}\right).
585   \end{split}
586   \label{eq:19}
587   \end{equation}
588 < 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.
589 < For quadrupolar fluids the total charge within a sphere is zero, therefore
590 < $ \int_{r<R} {\nabla}\mathbf{E}\;d^3r = 0 $.  Hence the quadrupolar
588 > In the last step, the charge density was integrated over the sphere,
589 > yielding a total charge $C_\mathrm{total}$.Equation (\ref{eq:19})
590 > gives the total gradient of the field over a sphere due to the
591 > distribution of the charges.  For quadrupolar fluids the total charge
592 > within a sphere is zero, therefore
593 > $ \int_{r<R} {\nabla}\mathbf{E}\;d\mathbf{r} = 0 $.  Hence the quadrupolar
594   polarization produces zero net gradient of the field inside the
595   sphere.
596  
379
597   \bibliography{dielectric_new}
598   \end{document}
599   %

Diff Legend

Removed lines
+ Added lines
< Changed lines
> Changed lines