ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/multipole/dielectric.tex
Revision: 4218
Committed: Thu Sep 11 18:08:53 2014 UTC (10 years ago) by gezelter
Content type: application/x-tex
File size: 15816 byte(s)
Log Message:
Changes

File Contents

# User Rev Content
1 gezelter 4195 \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    
21     \begin{document}
22    
23 gezelter 4218 \title{Real space electrostatics for multipoles. III. Dielectric Properties}
24 gezelter 4195
25 gezelter 4218 \author{Madan Lamichhane}
26     \affiliation{Department of Physics, University
27     of Notre Dame, Notre Dame, IN 46556}
28    
29     \author{Kathie E. Newman}
30     \affiliation{Department of Physics, University
31     of Notre Dame, Notre Dame, IN 46556}
32    
33     \author{Thomas Parsons}
34     \affiliation{Department of Chemistry and Biochemistry, University
35     of Notre Dame, Notre Dame, IN 46556}
36    
37 gezelter 4195 \author{J. Daniel Gezelter}
38     \email{gezelter@nd.edu.}
39     \affiliation{Department of Chemistry and Biochemistry, University
40     of Notre Dame, Notre Dame, IN 46556}
41    
42     \date{\today}% It is always \today, today,
43     % but any date may be explicitly specified
44    
45 gezelter 4218 \begin{abstract}
46     Note: This manuscript is a work in progress.
47    
48     We report on the dielectric properties of the shifted potential
49     (SP), gradient shifted force (GSF), and Taylor shifted force (TSF)
50     real-space methods for multipole interactions that were developed in
51     the first two papers in this series. We find that some subtlety is
52     required for computing dielectric properties with the real-space
53     methods, particularly when using the common fluctuation formulae.
54     Three distinct methods for computing the dielectric constant are
55     investigated, including the standard fluctuation formulae,
56     potentials of mean force between solvated ions, and direct
57     measurement of linear solvent polarization in response to applied
58     fields and field gradients.
59     \end{abstract}
60    
61 gezelter 4195 \maketitle
62    
63 gezelter 4218 \section{Introduction}
64 gezelter 4195
65 gezelter 4218 Over the past several years, there has been increasing interest in
66     pairwise methods for correcting electrostatic interactions in computer
67     simulations of condensed molecular
68     systems.\cite{Wolf99,Zahn02,Kast03,Beckd.A.C._Bi0486381,Ma05,Fennell06}
69     These techniques were developed from the observations and efforts of
70     Wolf {\it et al.} and their work towards an $\mathcal{O}(N)$
71     Coulombic sum.\cite{Wolf99} Wolf's method of cutoff neutralization is
72     able to obtain excellent agreement with Madelung energies in ionic
73     crystals.\cite{Wolf99} One of the most difficult tests of any new
74     electrostatic method is the fidelity with which that method can
75     reproduce the bulk-phase polarizability or equivalently, the
76     dielectric properties of a fluid. Of particular interest is the
77     static dielectric constant, $\epsilon$. Using the Ewald sum under
78     tin-foil boundary conditions, $\epsilon$ can be calculated for systems
79     of non-polarizable substances via
80     \begin{equation}
81     \epsilon = 1 + \frac{\langle M^2\rangle}{3\epsilon_0k_\textrm{B}TV},
82     \label{eq:staticDielectric}
83     \end{equation}
84     where $\epsilon_0$ is the permittivity of free space and $\langle
85     M^2\rangle$ is the fluctuation of the system dipole
86     moment.\cite{Allen:1989fp} The numerator in the fractional term in equation
87     (\ref{eq:staticDielectric}) is the fluctuation of the simulation-box
88     dipole moment, identical to the quantity calculated in the
89     finite-system Kirkwood $g$ factor ($G_k$):
90     \begin{equation}
91     G_k = \frac{\langle M^2\rangle}{N\mu^2},
92     \label{eq:KirkwoodFactor}
93     \end{equation}
94     where $\mu$ is the dipole moment of a single molecule of the
95     homogeneous system.\cite{Neumann:1983mz,Neumann:1983yq,Neumann84,Neumann85} The
96     fluctuation term in both equation (\ref{eq:staticDielectric}) and
97     (\ref{eq:KirkwoodFactor}) is calculated as follows,
98     \begin{equation}
99     \begin{split}
100     \langle M^2\rangle &= \langle\bm{M}\cdot\bm{M}\rangle
101     - \langle\bm{M}\rangle\cdot\langle\bm{M}\rangle \\
102     &= \langle M_x^2+M_y^2+M_z^2\rangle
103     - (\langle M_x\rangle^2 + \langle M_x\rangle^2
104     + \langle M_x\rangle^2).
105     \end{split}
106     \label{eq:fluctBoxDipole}
107     \end{equation}
108     This fluctuation term can be accumulated during a simulation; however,
109     it converges rather slowly, thus requiring multi-nanosecond simulation
110     times.\cite{Horn04} In the case of tin-foil boundary conditions, the
111     dielectric/surface term of the Ewald summation is equal to zero.
112 gezelter 4195
113 gezelter 4218 \section{Dielectric constants for real-space electrostatics}
114    
115 gezelter 4195 The new real-space methods require some care when computing dielectric
116     properties because the interaction between molecular dipoles depends
117     on the level of approximation being utilized. Consider the cases of
118     Stockmayer (dipolar) soft spheres that are represented either by two
119     closely-spaced point charges or by a single point dipole (see
120     Fig. \ref{fig:stockmayer}).
121     \begin{figure}
122     \includegraphics[width=3in]{DielectricFigure}
123     \caption{With the real-space electrostatic methods, the effective
124     dipole tensor, $\mathbf{T}$, governing interactions between
125     molecular dipoles is not the same for charge-charge interactions as
126     for point dipoles.}
127     \label{fig:stockmayer}
128     \end{figure}
129     In the case where point charges are interacting via an electrostatic
130     kernel, $v(r)$, the effective {\it molecular} dipole tensor,
131     $\mathbf{T}$ is obtained from two successive applications of the
132     gradient operator to the electrostatic kernel,
133     \begin{equation}
134     \mathbf{T}_{\alpha \beta}(r) = \nabla_\alpha \nabla_\beta \left(v(r)\right) = \delta_{\alpha \beta}
135     \left(\frac{1}{r} v^\prime(r) \right) + \frac{r_{\alpha}
136     r_{\beta}}{r^2} \left( v^{\prime \prime}(r) - \frac{1}{r}
137     v^{\prime}(r) \right)
138     \end{equation}
139     where $v(r)$ may be either the bare kernel ($1/r$) or one of the
140     modified (Wolf or DSF) kernels. This tensor describes the effective
141 gezelter 4196 interaction between molecular dipoles ($\mathbf{D}$) in Gaussian
142 gezelter 4195 units as $-\mathbf{D} \cdot \mathbf{T} \cdot \mathbf{D}$.
143    
144     When utilizing the new real-space methods for point dipoles, the
145     tensor is explicitly constructed,
146     \begin{equation}
147     \mathbf{T}_{\alpha \beta}(r) = \delta_{\alpha \beta} v_{21}(r) +
148     \frac{r_{\alpha} r_{\beta}}{r^2} v_{22}(r)
149     \end{equation}
150     where the functions $v_{21}(r)$ and $v_{22}(r)$ depend on the level of
151     the approximation. Although the Taylor-shifted (TSF) and
152     gradient-shifted (GSF) models produce to the same $v(r)$ function for
153     point charges, they have distinct forms for the dipole-dipole
154     interactions.
155    
156     Neumann and Steinhauser showed~\cite{Neumann:1983mz,Neumann:1983yq}
157     that the relative dielectric permittivity $\epsilon$ is given by the
158     general fluctuation formula,
159     \begin{equation}
160     \frac{\epsilon - 1}{\epsilon +2} = \frac{4 \pi}{3} \frac{\left<M^2\right>}{3 V k_B
161     T} \left( 1 - \frac{3}{4 \pi} \frac{\epsilon -1}{\epsilon + 2}
162     \tilde{T}(0) \right)
163     \end{equation}
164     where $\left<M^2\right>$ is the mean fluctuation of the square of the
165     net dipole moment of the simulation cell, $V$ is the volume of the
166     cell, and $\tilde{T}(0) = \tilde{T̃}_{xx}(0) = \tilde{T̃}_{yy}(0) =
167     \tilde{T̃}_{zz}(0)$ is the $\mathbf{k} = 0$ limit of Fourier transform
168     of the diagonal term of the effective molecular dipolar tensor,
169     \begin{equation}
170     \tilde{T}(0) = \int_V \mathbf{T}(\mathbf{r}) d\mathbf{r}
171     \end{equation}
172     where the integral is carried out over the relevant geometry for the
173     interaction. For the real-space methods, the integration can be done
174     easily up to the imposed cutoff radius, while for the Ewald sum, the
175     results also depend on details of the $k$-space
176     calculation.\cite{Neumann:1983yq}
177    
178     For molecules composed of point charges interacting via the bare
179     $(1/r)$ kernel, the simple cutoff (SC) and minimum image (MI)
180     approaches give $\tilde{T}(0) = 0$ which means that the
181     Clausius-Mosotti equation,
182     \begin{equation}
183     \frac{4\pi}{3}\frac{\left< M^2 \right>}{3 V k_B T} = \frac{\epsilon
184     -1}{\epsilon+2}
185     \end{equation}
186     is the relevant fluctuation formula for the relative permittivity.
187    
188     Zahn {\it et al.}\cite{Zahn02} showed that for the damped shifted
189     force charge-charge kernel, $\tilde{T}(0) = 4 \pi / 3$. This was later
190     generalized by Izvekov {\it et al.} for all point-charge kernels which
191     have forces that go to zero at a cutoff radius and which maintain a
192     pole of first order at $r=0$.\cite{Izvekov:2008wo} When $\tilde{T}(0)
193     = 4 \pi/ 3$, the expression for the dielectric constant reduces to the
194     widely-used {\it conducting boundary} formula,
195     \begin{equation}
196     \frac{4\pi}{3}\frac{\left< M^2 \right>}{3 V k_B T} = \frac{\epsilon -
197     1}{3}
198     \end{equation}
199     It is convenient to define a quantity $Q = \frac{3}{4 \pi}
200     \tilde{T}(0)$, and to combine all of the fluctuation formulae
201     together:
202     \begin{equation}
203     \frac{4 \pi}{3} \frac{\left< M^2 \right>}{3 V k_B T} =
204     \left\{\begin{array}{ll}
205     \frac{\epsilon-1}{\epsilon+2} \left[1- \frac{\epsilon-1}{\epsilon+2}
206     Q\right]^{-1} & \mathrm{General~Case} \\
207     \frac{\epsilon-1}{\epsilon+2} & Q \rightarrow 0 \mathrm{~limit} \\
208     \frac{\epsilon-1}{3} & Q \rightarrow 1 \mathrm{~limit}
209     \end{array}\right.
210     \end{equation}
211    
212     The Clausius-Mossotti $(Q\rightarrow 0)$ approach is subject to noise
213     and error magnification,\cite{Allen:1989fp} and the conducting
214     boundary approach $(Q \rightarrow 1)$ has become widely used as a
215     result. If the electrostatic method being used has $Q < 1$, the
216     relative dielectric permittivity $\epsilon$ can be estimated once the
217     conducting boundary fluctuation result has been found,
218     \begin{equation}
219     \epsilon = \frac{(Q+2) (\epsilon_\text{CB}-1)+3} {(Q-1) (\epsilon_\text{CB}-1)+3}
220     \end{equation}
221     Note that this expression becomes quite numerically sensitive when the
222     value of $Q$ deviates significantly from 1.
223    
224     We have derived a set of expressions for the value of $Q$ for the new
225     real space methods that are shown in Table \ref{tab:Q}.
226    
227     \begin{table}
228     \caption{Expressions for the dielectric correction factor ($Q$) for the
229     real-space electrostatic methods in terms of the damping parameter
230     ($\alpha$) and the cutoff radius ($r_c$). The Ewald-Kornfeld result
231     derived in Refs. \onlinecite{Adams:1980rt,Adams:1981fr,Neumann:1983yq} is shown for comparison using the Ewald
232     convergence parameter ($\kappa$) and the real-space cutoff value ($r_c$). }
233     \label{tab:Q}
234     \begin{tabular}{l|c|c|c|c|}
235     & \multicolumn{2}{c|}{damped} & \multicolumn{2}{c|}{undamped} \\ \cline{2-3} \cline{4-5}
236     Method & $Q_\mathrm{charges}$ & $Q_\mathrm{dipoles}$ & $Q_\mathrm{charges}$ & $Q_\mathrm{dipoles}$ \\
237     \hline
238     Spherical Cutoff (SC) & $\mathrm{erf}(r_c \alpha) - \frac{2 \alpha r_c}{\sqrt{\pi}} e^{-\alpha^2 r_c^2}$ & $\mathrm{erf}(r_c \alpha) - \frac{2 \alpha r_c}{\sqrt{\pi}} e^{-\alpha^2 r_c^2}$ &0 &0\\
239     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} $ &0 &0\\
240     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} $ &1 &1\\
241     Taylor-shifted (TSF) & 1 & 1 & 1 & 1\\
242     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
243     \end{tabular}
244     \end{table}
245 gezelter 4218
246 gezelter 4195 \newpage
247    
248 gezelter 4218 \appendix
249     \section{Point-multipolar interactions with a spatially-varying electric field}
250 gezelter 4195
251 gezelter 4218 We can treat objects $a$, $b$, and $c$ containing embedded collections
252     of charges. When we define the primitive moments, we sum over that
253     collections of charges using a local coordinate system within each
254     object. The point charge, dipole, and quadrupole for object $a$ are
255     given by $C_a$, $\mathbf{D}_a$, and $\mathsf{Q}_a$, respectively.
256     These are the primitive multipoles which can be expressed as a
257     distribution of charges,
258     \begin{align}
259     C_a =&\sum_{k \, \text{in }a} q_k , \label{eq:charge} \\
260     D_{a\alpha} =&\sum_{k \, \text{in }a} q_k r_{k\alpha}, \label{eq:dipole}\\
261     Q_{a\alpha\beta} =& \frac{1}{2} \sum_{k \, \text{in } a} q_k
262     r_{k\alpha} r_{k\beta} . \label{eq:quadrupole}
263     \end{align}
264     Note that the definition of the primitive quadrupole here differs from
265     the standard traceless form, and contains an additional Taylor-series
266     based factor of $1/2$. In Paper 1, we derived the forces and torques
267     each object exerts on the others.
268    
269     Here we must also consider an external electric field that varies in
270     space: $\mathbf E(\mathbf r)$. Each of the local charges $q_k$ in
271     object $a$ will then experience a slightly different field. This
272     electric field can be expanded in a Taylor series around the local
273     origin of each object. A different Taylor series expansion is carried
274     out for each object.
275    
276     For a particular charge $q_k$, the electric field at that site's
277     position is given by:
278     \begin{equation}
279     E_\gamma + \nabla_\delta E_\gamma r_{k \delta}
280     + \frac {1}{2} \nabla_\delta \nabla_\varepsilon E_\gamma r_{k \delta}
281     r_{k \varepsilon} + ...
282     \end{equation}
283     Note that the electric field is always evaluated at the origin of the
284     objects, and treating each object using point multipoles simplifies
285     this greatly.
286    
287     To find the force exerted on object $a$ by the electric field, one
288     takes the electric field expression, and multiplies it by $q_k$, and
289     then sum over all charges in $a$:
290    
291     \begin{align}
292     F_\gamma &= \sum_{k \textrm{~in~} a} q_k \lbrace E_\gamma + \nabla_\delta E_\gamma r_{k \delta}
293     + \frac {1}{2} \nabla_\delta \nabla_\varepsilon E_\gamma r_{k \delta}
294     r_{k \varepsilon} + ... \rbrace \\
295     &= C_a E_\gamma + D_{a \delta} \nabla_\delta E_\gamma
296     + Q_{a \delta \varepsilon} \nabla_\delta \nabla_\varepsilon E_\gamma +
297     ...
298     \end{align}
299    
300     Similarly, the torque exerted by the field on $a$ can be expressed as
301     \begin{align}
302     \tau_\alpha &= \sum_{k \textrm{~in~} a} (\mathbf r_k \times q_k \mathbf E)_\alpha \\
303     & = \sum_{k \textrm{~in~} a} \epsilon_{\alpha \beta \gamma} q_k
304     r_{k\beta} E_\gamma(\mathbf r_k) \\
305     & = \epsilon_{\alpha \beta \gamma} D_\beta E_\gamma
306     + 2 \epsilon_{\alpha \beta \gamma} Q_{\beta \delta} \nabla_\delta
307     E_\gamma + ...
308     \end{align}
309    
310     The last term is essentially identical with form derived by Torres del
311     Castillo and M\'{e}ndez Garrido,\cite{Torres-del-Castillo:2006uo} although their derivation
312     utilized a traceless form of the quadrupole that is different than the
313     primitive definition in use here. We note that the Levi-Civita symbol
314     can be eliminated by utilizing the matrix cross product in an
315     identical form as in Ref. \onlinecite{Smith98}:
316     \begin{equation}
317     \left[\mathsf{A} \times \mathsf{B}\right]_\alpha = \sum_\beta
318     \left[\mathsf{A}_{\alpha+1,\beta} \mathsf{B}_{\alpha+2,\beta}
319     -\mathsf{A}_{\alpha+2,\beta} \mathsf{B}_{\alpha+1,\beta}
320     \right]
321     \label{eq:matrixCross}
322     \end{equation}
323     where $\alpha+1$ and $\alpha+2$ are regarded as cyclic permuations of
324     the matrix indices. In table \ref{tab:UFT} we give compact
325     expressions for how the multipole sites interact with an external
326     field that has exhibits spatial variations.
327    
328     \begin{table}
329     \caption{Potential energy $(U)$, force $(\mathbf{F})$, and torque
330     $(\mathbf{\tau})$ expressions for a multipolar site embedded in an
331     electric field with spatial variations, $\mathbf{E}(\mathbf{r})$.
332     \label{tab:UFT}}
333     \begin{tabular}{r|ccc}
334     & Charge & Dipole & Quadrupole \\ \hline
335     $U$ & $C \phi(\mathbf{r})$ & $-\mathbf{D} \cdot \mathbf{E}(\mathbf{r})$ & $- \mathsf{Q}:\nabla \mathbf{E}(\mathbf{r})$ \\
336     $\mathbf{F}$ & $C \mathbf{E}(\mathbf{r})$ & $+\mathbf{D} \cdot \nabla \mathbf{E}(\mathbf{r})$ & $+\mathsf{Q} : \nabla\nabla\mathbf{E}(\mathbf{r})$ \\
337     $\mathbf{\tau}$ & & $\mathbf{D} \times \mathbf{E}(\mathbf{r})$ & $+2 \mathsf{Q} \times \nabla \mathbf{E}(\mathbf{r})$
338     \end{tabular}
339     \end{table}
340    
341    
342     \newpage
343    
344 gezelter 4195 \bibliography{multipole}
345    
346     \end{document}