cond-mat0402146/abc.tex
1: \documentclass[twocolumn,showpacs,preprintnumbers,amsmath,amssymb,groupedaddress,pre]{revtex4}
2: %\documentclass[preprint,showpacs,preprintnumbers,amsmath,amssymb]{revtex4}
3: 
4: % Some other (several out of many) possibilities
5: %\documentclass[preprint,aps]{revtex4}
6: %\documentclass[preprint,aps,draft]{revtex4}
7: %\documentclass[prb]{revtex4}% Physical Review B
8: 
9: \usepackage{graphicx}% Include figure files
10: \usepackage{dcolumn}% Align table columns on decimal point
11: \usepackage{bm}% bold math
12: \usepackage{epsfig}
13: 
14: \def\e{\begin{equation}}
15: \def\f{\end{equation}}
16: \def\=#1{\overline{\overline #1}}
17: \def\s{\strut\displaystyle}
18: \def\-#1{{\bf #1}}
19: \def\.{\cdot}
20: \def\l#1{\label{eq:#1}}
21: \def\r#1{(\ref{eq:#1})}
22: \def\vec#1{{\bf #1}}
23: 
24: \begin{document}
25: 
26: \title{Boundary conditions for interfaces of electromagnetic (photonic) crystals and generalized Ewald-Oseen extinction principle}
27: 
28: \author{Pavel A. Belov}
29: \affiliation{Queen Mary College, University of London, Mile End
30: Road, London, E1 4NS, United Kingdom} \affiliation{Photonics and
31: Optoinformatics Department, St. Petersburg State University of
32: Information Technologies, Mechanics and Optics, Sablinskaya 14,
33: 197101, St. Petersburg, Russia}
34: 
35: %\email{belov@rain.ifmo.ru}
36: 
37: \author{Constantin R. Simovski}
38: \affiliation{Photonics and Optoinformatics Department, St. Petersburg State University of Information Technologies, Mechanics and Optics, Sablinskaya 14, 197101, St. Petersburg, Russia}
39: 
40: \date{\today}
41: 
42: 
43: \begin{abstract}
44: The problem of plane-wave diffraction on semi-infinite orthorhombic
45: electromagnetic (photonic) crystals of general kind is considered.
46: Boundary conditions are obtained in the form of infinite system of
47: equations relating amplitudes of incident wave, eigenmodes excited
48: in the crystal and scattered spatial harmonics. Generalized
49: Ewald-Oseen extinction principle is formulated on the base of
50: deduced boundary conditions. The knowledge of properties of infinite
51: crystal's eigenmodes provides option to solve the diffraction
52: problem for the corresponding semi-infinite crystal numerically. In
53: the case when the crystal is formed by small inclusions which can be
54: treated as point dipolar scatterers with fixed direction the problem
55: admits complete rigorous analytical solution. The amplitudes of
56: excited modes and scattered spatial harmonics are expressed in terms
57: of the wave vectors of the infinite crystal by closed-form
58: analytical formulae. The result is applied for study of reflection
59: properties of metamaterial formed by cubic lattice of split-ring
60: resonators.
61: \end{abstract}
62: 
63: \pacs{78.20.Ci, 42.70.Qs, 42.25.Fx, 73.20.Mf}
64: \maketitle
65: 
66: \section{Introduction}
67: 
68: Electromagnetic crystals are artificial periodical structures
69: operating at the wavelengths comparable with their periods
70: \cite{Sakoda,PhotJMW,specialis}. At the optical frequencies such
71: structures are called as photonic crystals. The inherent feature of
72: these materials is the existence of frequency bands where the
73: crystal does not support propagating waves. The band gaps are caused
74: by spatial resonances of the crystal lattice and strongly depend on
75: the direction of propagation. It means that electromagnetic crystals
76: are media with spatial dispersion \cite{Agranovich,Agarwal,Birman}.
77: The material parameters: permittivity and permeability for such
78: materials, if they can be introduced at all, depend on the wave
79: vector as well as on the frequency. Notice, that the homogenization
80: approach is not the most convenient way for the description of
81: electromagnetic crystals even at low frequencies. It often requires
82: introduction of additional boundary conditions in order to describe
83: boundary problems correctly, and this involves related complexities.
84: The photonic and electromagnetic crystals are usually studied with
85: the help of numerical methods \cite{Sakoda,PhotJMW,specialis}.
86: Analytical models exist only for a very narrow class of the
87: crystals. Some types of the crystals can be studied analytically
88: under a certain approximation, but the strict analytical solution
89: for a photonic crystal is an exception.
90: 
91: The goal of the present paper is to demonstrate how boundary
92: problems for electromagnetic crystals can be effectively studied
93: using analytical methods. The paper is separated into the two parts.
94: In the first part the boundary conditions for electromagnetic
95: crystals of general kind are deduced in the form of infinite system
96: of equations relating amplitudes of incident wave, excited
97: eigenmodes of the crystal and scattered spatial harmonics. This
98: system can be interpreted as generalization of well-known
99: Ewald-Oseen extinction principle \cite{Ewald,Oseen,BornWolf} which
100: states that the polarization of dielectric is distributed so that it
101: cancels out the incident wave and produces the propagating wave. For
102: the electromagnetic crystals, inherently periodic structures, the
103: generalized Ewald-Oseen principle states that the polarization of
104: dielectric is distributed so that it cancels out the incident wave
105: as well as all spatial harmonics associated with periodicity of the
106: boundary. This principle expressed in the form of infinite system of
107: boundary conditions provides opportunity to solve the boundary
108: problem for semi-infinite crystal of certain kind numerically if the
109: eigenmode problem for corresponding infinite crystal is already
110: solved. In the second part of the paper the proposed approach is
111: applied for the case of electromagnetic crystals formed by small
112: inclusions which can be treated as point dipolar scatterers with
113: fixed direction. In this case the system of boundary conditions
114: admits complete rigorous analytical solution. The amplitudes of
115: excited eigenmodes and scattered spatial harmonics are expressed in
116: terms of wavevectors of eigenmodes using closed-form analytical
117: formulae. These results are unique extension and generalization of
118: known Mahan-Obermair theory \cite{Mahan} for the case then period of
119: the crystal is compared with wavelength. At the end of the paper it
120: is demonstrated how reflection from the semi-infinite cubic lattice
121: of resonant scatterers (split-ring resonators) can be modeled in the
122: regime of strong spatial dispersion observed in such crystals
123: \cite{Belovhomo}.
124: 
125: \section{Proof of generalized Ewald-Oseen extinction principle}
126: 
127: In this section we provide proof of generalized Ewald-Oseen
128: extinction principle for arbitrary semi-infinite electromagnetic
129: crystal with orthorhombic elementary cell. First, let us consider an
130: infinite orthorhombic electromagnetic crystal with geometry
131: schematically presented in Fig.~\ref{gengeom} and characterized by
132: three-periodical permittivity distribution: \e \=\varepsilon(\vec
133: r)=\=\varepsilon(\vec r+\vec a n+\vec b s+\vec c l). \f In this
134: expression and in the further text the two lines over a quantity
135: designate that the quantity is dyadic (tensor of second rank in
136: three-dimensional space). It means that we consider  of the most
137: general kind of electromagnetic crystals formed by dielectrics.
138: \begin{figure}[h]
139: \centering \epsfig{file=gengeom.eps, width=7.5cm} \caption{Geometry
140: of an infinite electromagnetic crystal} \label{gengeom}
141: \end{figure}
142: 
143: In this paper we are using local field approach, an unconventional
144: method for description of fields inside dielectrics. We will operate
145: with local parameters like polarization density $\vec P$ and local
146: electrical field $\vec E_{\rm loc.}$, but not with average electric
147: field $\vec E$ and displacement $\vec D$ as usual. The similar
148: approach was used in \cite{BornWolf} for rigorous derivation of
149: Ewald-Oseen extinction theorem and in \cite{Birman}. The dielectric
150: can be treated as a very dense cubic lattice of point scatterers
151: with ceratin local polarizability. In this formulation the
152: dielectric permittivity $\=\varepsilon(\vec r)$ has to be replaced
153: (see Figure \ref{local} for illustration) by the local
154: \begin{figure}[h]
155: \centering \epsfig{file=local.eps, width=7.5cm}
156: \caption{Illustration for replacement of dielectric permittivity by
157: local polarizability} \label{local}
158: \end{figure}
159: polarizability $\=\alpha(\vec r)$ relating the bulk polarization
160: density $\vec P (\vec r)$ with the local electric field $\vec E_{\rm
161: loc}(\vec r)$: \e
162:  \vec P(\vec r)=\=\alpha(\vec r)\vec E_{\rm loc}(\vec r).
163: \f
164: 
165: The expression for local polarizability in terms of dielectric
166: permittivity has the following form: \e \=\alpha^{-1}(\vec
167: r)=\left[\=\varepsilon(\vec r)-\varepsilon_0\=
168: I\right]^{-1}+\=I/(3\varepsilon_0), \l{alphar} \f where
169: $\varepsilon_0$ is permittivity of free space, $\=I$ is unit dyadic.
170: This expression follows from the Lorentz-Lorenz formula
171: \cite{BornWolf} \e \vec E_{\rm loc} (\vec r) = \vec E(\vec r) + \vec
172: P (\vec r)/(3\varepsilon_0)\f and material equation \e \vec D (\vec
173: r)=\varepsilon_0\vec E(\vec r)+\vec P(\vec r)=\=\varepsilon(\vec r)
174: \vec E(\vec r).\l{me}
175:  \f
176: 
177: \subsection{Dispersion equation}
178: Following local field approach one can write down dispersion
179: equation for the crystal under consideration in the next integral
180: form: \e \vec P (\vec r)=\=\alpha(\vec r) \int\limits_V\=G_3(\vec
181: r-\vec r',\vec q)\vec P (\vec r')d\vec r', \ \forall \ \vec r\in V,
182: \l{disp2} \f where $V=V(\vec a, \vec b, \vec c)$ is volume of the
183: elementary lattice cell, $\=G_3(\vec r, \vec q)$ is the lattice
184: dyadic Green's function: \e \=G_3(\vec r, \vec
185: q)=\sum\limits_{n,s,l}\=G(\vec r-\vec a n-\vec b s-\vec c
186: l)e^{-j(q_xan+q_ybs+q_zcl)}, \l{G3} \f which takes into account
187: cell-to-cell polarization distribution determined by wave vector
188: $\vec q=(q_x,q_y,q_z)^T$: \e \vec P(\vec r+\vec a n+\vec b s+\vec c
189: l)=\vec P(\vec r)e^{-j(q_xan+q_ybs+q_zcl)}, \l{shiftp2} \f $\=G(\vec
190: r)$ is dyadic Green's function of free space: \e \=G(\vec r) =
191: (k^2\=I+\nabla\nabla)\frac{e^{-jkr}}{4\pi\varepsilon_0 r},
192: \l{gfree}\f $n,s,l$ are integer indices, $k$ is wave number of free
193: space. The integral in \r{disp2} is singular if the point
194: corresponding to vector $r$ is located inside of some polarized
195: dielectric. It has to be evaluated in the meaning of principal value
196: by excluding small spherical region around the singular point and
197: tending the radius of this region to zero \cite{YaghjianDGF}.
198: 
199: The dispersion equation \r{disp2} relates distribution of
200: polarization density $\vec P(\vec r)$ and wave vector $\vec q$
201: corresponding to the eigenmodes of the electromagnetic crystal. If
202: the distribution of average electric field $\vec E(\vec r)$ of a
203: crystal eigenmode  is known then the polarization density $\vec
204: P(\vec r)$ can be found directly using material equation \r{me}: \e
205: \vec P(\vec r)=[\=\varepsilon(\vec r)-\varepsilon_0]\vec E(\vec
206: r).\f The reverse operation is possible only for space regions
207: filled by dielectric with $\=\varepsilon(\vec r)\ne \varepsilon_0$.
208: The distribution of electric field in free space regions, if
209: required, have to be calculated using the next integral
210: representation \e \vec E (\vec r)=\int\limits_V\=G_3(\vec r-\vec
211: r',\vec q)\vec P (\vec r')d\vec r'.\f
212: 
213: For our proof of generalized Ewald-Oseen extinction principle we
214: have to transform dispersion equation \r{disp2} into the form
215: corresponding to summation by layers in $x$-direction. The
216: expression for lattice dyadic Green's function $\=G_3(\vec r, \vec
217: q)$ \r{G3} can be rewritten using summation over planes in the form:
218: \e \=G_3(\vec r, \vec q)=\sum\limits_{n=-\infty}^{+\infty}\=G_2(\vec
219: r-\vec a n) e^{-jq_xan}, \l{G3layer} \f where $\=G_2(\vec r)$ is the
220: grid dyadic Green's function: \e \=G_2(\vec
221: r)=\sum\limits_{s,l}\=G(\vec r-\vec b s-\vec c
222: l)e^{-j(q_ybs+q_zcl)}. \f
223: 
224: Applying Poisson summation formula by both indices $s$ and $l$ one
225: can express the grid dyadic Green's function in terms of the spatial
226: Floquet harmonics. This expansion is also called as spectral
227: representation: \e \=G_2(\vec r) = \sum\limits_{s,l}
228: \=\gamma_{s,l}^{{\rm sign}(x-a)} e^{-j(\vec k^{{\rm
229: sign}(x)}_{s,l}\.\vec r)}, \l{genfloquet2}\f where $$
230: \=\gamma^{\pm}_{s,l}=\frac{j}{2bc\varepsilon_0 k_{s,l}} [\-
231: k^{\pm}_{s,l}\times[\- k^{\pm}_{s,l}\times \= I]], \ \vec
232: k^{\pm}_{s,l}=(\pm k^x_{s,l}, k^y_{s}, k^z_{l} )^T,$$
233: $$k^y_{s}=q_y+\frac{2\pi s}{b}, k^z_{l}=q_z+\frac{2\pi l}{c},
234: k^x_{s,l}=\sqrt{k^2-(k^y_{s})^2-(k^z_{l})^2}.$$ The square root in
235: the expression for $k_{s,l}$ should be chosen so that ${\rm
236: Im}(\sqrt{\.})<0$. The sign $\pm$ corresponds to half spaces $x>a$
237: and $x<a$ respectively.
238: 
239: Using \r{G3layer} the dispersion equation \r{disp2} can be rewritten
240: in the following form which will be used later on: \e \vec P (\vec
241: r)=\=\alpha(\vec r) \sum\limits_{n=-\infty}^{+\infty}\int\limits_V
242: \=G_2(\vec r-\vec r'-\vec a n) \vec P (\vec r') e^{-jq_xan} d\vec
243: r'. \l{displayer2} \f
244: 
245: \subsection{Semi-infinite crystal}
246: Now let us consider a semi-infinite crystal (half space $x\ge a$,
247: see Figure \ref{geomsemi}) excited by a plane electromagnetic wave
248: with wave vector $\vec k=(k_x,k_y,k_z)^{T}$ coming from free space:
249: \e \vec E_{\rm inc}(\vec r)=\vec E_{\rm inc} e^{-j(\vec k\.\vec r)}.
250: \l{inc2} \f The origin of our coordinate system is intentionally
251: shifted by one period into the free space since it simplifies rather
252: cumbersome calculations which are presented below and causes
253: exponential convergence of series in the final expressions.
254: 
255: \begin{figure}[h]
256: \centering \epsfig{file=gensemi.eps, width=7.5cm} \caption{Geometry
257: of an semi-infinite electromagnetic crystal} \label{geomsemi}
258: \end{figure}
259: 
260: Due to the periodicity of the semi-infinite structure along $y-$ and
261: $z-$axes the distribution of the excited polarization along these
262: directions is determined by the phase of the incident wave: \e \vec
263: P(\vec r+\vec a m+\vec b s+\vec c l)=\vec P(\vec r+\vec a
264: m)e^{-j(k_ybn+k_zcl)}, \l{shiftpsemi2} \f for any $\vec r\in V$ and
265: $m\ge 1$.
266: 
267: The electric field produced by the polarized semi-infinite crystal
268: has the form \e \vec E_{\rm scat}(\vec r) =
269: \sum\limits_{n=1}^{+\infty} \int\limits_V \=G_2(\vec r -\vec r' -
270: \vec a n)  \vec P(\vec r'+\vec a n) d\vec r'. \l{scat2} \f
271: 
272: The total local electric field is sum of incident and scattered
273: (produced by polarization of crystal) fields. Following the local
274: field approach we can write: \e \vec P(\vec r)= \=\alpha(\vec r)
275: \left[\vec E_{\rm inc}(\vec r)+\vec E_{\rm scat}(\vec r)\right].
276: \l{local} \f
277: 
278: Combining \r{scat2} and \r{local} we obtain an integral equation for
279: the polarization in the semi-infinite crystal excited by an incident
280: wave: \e \vec P(\vec r) =\=\alpha(\vec r) \left[ \vec E_{\rm
281: inc}(\vec r)+\sum\limits_{n=1}^{+\infty} \int\limits_V \=G_2(\vec r
282: -\vec r' - \vec a n)  \vec P(\vec r'+\vec a n) d\vec r'\right].
283: \l{layer2} \f
284: 
285: Now let us suppose that the dispersion equation \r{disp2} is solved
286: under condition that wave vector has the form $\vec
287: q=(q_x,k_y,k_z)^T$ with unknown $x-$component $q_x$, and the set of
288: eigenmodes $\{(\vec P_i(\vec r),q^{(i)}_x)\}$, characterized by
289: x-component of wave vector $q^{(i)}_x$ and polarization distribution
290: $\vec P_i(\vec r)$, is found. In addition, we include into this set
291: only the eigenmodes which either transfer energy into half-space
292: $x\ge a$ [$dq^{(i)}_x/d\omega>0$] or decay in $x$-direction [${\rm
293: Im}(q^{(i)}_x)<0$]. In such a case the polarization of the
294: semi-infinite crystal excited by the incident wave \r{inc2} can be
295: expanded by the eigenmodes of infinite crystal as follows: \e \vec
296: P(\vec r+\vec a m)=\sum\limits_i A_i \vec P_i(\vec r)
297: e^{-jq^{(i)}_xam}, \ \forall \vec r \in V, m\ge 1. \l{expand2} \f
298: 
299: Substituting \r{expand2} into \r{layer2} and using \r{inc2} we
300: obtain:
301: 
302:  \e \sum\limits_i A_i \vec P_i(\vec
303: r)e^{-jq_x^{(i)}am} =\=\alpha(\vec r) \left[
304: \vphantom{\sum\limits_{n=0}^{+\infty} \int\limits_V} \vec E_{\rm
305: inc}e^{-j \vec k(\vec r+\vec a m)}\right. \l{subst1} \f
306: $$
307: \left. +\sum\limits_{n=1}^{+\infty} \sum\limits_i A_i \int\limits_V
308: \=G_2(\vec r -\vec r' - \vec a (n-m)) \vec P_i(\vec r')
309: e^{-jq_x^{(i)}an} d\vec r'\right].
310: $$
311: 
312: Splitting series in the dispersion equation \r{displayer2} one can
313: derive the following auxiliary relation: $$\vec P_i (\vec
314: r)e^{-jq_x^{(i)}am}=
315: $$
316: \e =\=\alpha(\vec r) \sum\limits_{n=1}^{+\infty} \int\limits_V
317: \=G_2(\vec r-\vec r'-\vec a (n-m)) \vec P_i (\vec r')
318: e^{-jq_x^{(i)}an} d\vec r'  \l{aux2} \f
319: $$ +\=\alpha(\vec r) \sum\limits_{n=-\infty}^{0} \int\limits_V
320: \=G_2(\vec r -\vec r'-\vec a (n-m)) \vec P_i (\vec r')
321: e^{-jq_x^{(i)}an} d\vec r'.$$
322: 
323: Substituting \r{aux2} into \r{subst1} and following the fact that
324: ${\rm det} \{\=\alpha(\vec r)\}\ne 0$ we obtain: $$ \sum\limits_i
325: A_i \sum\limits_{n=-\infty}^{0} \int\limits_V \=G_2(\vec r -\vec
326: r'-\vec a (n-m)) \vec P_i (\vec r') e^{-jq_x^{(i)}an} d\vec r' $$\e
327: =\vec E_{\rm inc}e^{-j \vec k(\vec r+\vec a m)}. \l{subst2} \f
328: 
329: Further, substituting \r{genfloquet2} into \r{subst2}, changing the
330: summation order and evaluating the sum of the geometrical
331: progression by index $n$ we get: $$ \sum\limits_{s,l} \left(
332: \sum\limits_i A_i \frac{\=\gamma_{s,l}^{+}  \int\limits_V \vec P_i
333: (\vec r') e^{j(\vec k^{+}_{s,l}\.\vec r')}d\vec
334: r'}{1-e^{j(q_x^{(i)}-k^x_{s,l})a}}\right) e^{-j(\vec
335: k^{+}_{s,l}\.(\vec r+\vec a m))}$$\e =\vec E_{\rm inc}e^{-j \vec
336: k(\vec r+\vec a m)}. \l{system2} \f
337: 
338: 
339: \subsection{Generalized Ewald-Oseen extinction principle}
340: 
341: The left part of equation \r{system2} represents an expansion of the
342: right part into a spatial spectrum of Floquet harmonics. The right
343: part represents an incident spectrum of Floquet harmonics containing
344: only the single incident plane wave \r{inc2} with $\vec k=\vec
345: k^+_{0,0}$. Equating coefficients in the left and right parts of
346: \r{system2} we obtain: \e \=\gamma_{s,l}^{+} \sum\limits_i A_i
347: \frac{\int\limits_V \vec P_i (\vec r') e^{j(\vec k^{+}_{s,l}\.\vec
348: r')}d\vec r'}{1-e^{j(q_x^{(i)}-k^x_{s,l})a}}
349: =\left\{\begin{array}{lcl}\vec E_{\rm inc},\ (s,l)=(0,0)\\ 0,\
350: (s,l)\ne (0,0) \end{array}.\right. \l{ewald2} \f
351: 
352: The values in the right side of \r{ewald2} are the amplitudes of the
353: incident spatial harmonics (all harmonics except fundamental one
354: have zero amplitudes), and the series in the left side are the
355: amplitudes of the spatial harmonics produced by the whole
356: semi-infinite crystal polarization in order to cancel these incident
357: harmonics. It means that equation \r{ewald2} represents the
358: generalization of Ewald-Oseen extinction principle (see \cite{Ewald,
359: Oseen, BornWolf} for classical formulation in the case of
360: dielectrics): {\it the polarization in a semi-infinite
361: electromagnetic crystal excited by a plane wave is distributed in
362: such a way that it cancels the incident wave together with all
363: high-order spatial harmonics associated with periodicity of the
364: boundary} (even if they have zero amplitudes as in the present
365: case). The additional words related to high-order Floquet harmonics
366: is the main and principal difference of Ewald-Oseen extinction
367: principle formulation for electromagnetic crystals as compared to
368: the classical case of isotropic dielectrics.
369: 
370: Substitution of \r{expand2} and \r{genfloquet2} into \r{scat2}
371: allows to express the scattered field in the half space $x<a$ in
372: terms of spatial Floquet harmonics: \e \vec E_{\rm scat}=
373: \sum\limits_{s,l} \vec E^{s,l}_{\rm scat} e^{-j(\vec k^-_{s,l}\.\vec
374: r)},\f  \e \vec E^{s,l}_{\rm scat}= \=\gamma_{s,l}^{-} \sum\limits_i
375: A_i \frac{\int\limits_V \vec P_i (\vec r') e^{j(\vec
376: k^{-}_{s,l}\.\vec r')}d\vec r'}{1-e^{j(q_x^{(i)}+k^x_{s,l})a}}.
377: \l{amplscat2} \f Note, that the formula for the amplitudes of
378: scattered Floquet harmonics \r{amplscat2} contains series which have
379: the same form as \r{ewald2} and differs only by the sign of the $x-$
380: components of wave vectors $\vec k^\pm_{s,l}=(\pm
381: k^x_{s,l},k_y,k_z)^T$ corresponding to the spatial harmonics
382: propagating into the half spaces $x<a$ and $x>a$, respectively.
383: 
384: 
385: If the eigenmodes of the crystal $\{q^{(i)}_x,\vec P_i(\vec r)\}$
386: are known then one can solve the system of linear equations
387: \r{ewald2} and find amplitudes of excited eigenmodes $\{A_i\}$. With
388: use of these amplitudes the scattered field can be found by
389: \r{amplscat2}. This provides a new numerical method which allows to
390: solve problem of plane-wave diffraction by a semi-infinite
391: electromagnetic crystal using knowledge of eigenmodes of the
392: infinite crystal. This fact is very important since at the moment
393: the reflection and dispersion problems for electromagnetic crystals
394: are usually solved by separate numerical approaches. The expressions
395: \r{ewald2} and \r{amplscat2} create a link between these two
396: problems and show how results of dispersion studies can be used in
397: order to describe reflection properties of electromagnetic crystals.
398: 
399: Until this point we considered only one incident wave with wave
400: vector $\vec k=\vec k^+_{0,0}$, but from \r{system2} it is clear
401: that we could consider also other incident spatial harmonics with
402: wave vectors $\vec k^+_{s,l}$ and get the similar results as
403: \r{ewald2}, but the nonzero terms at the right side of equation
404: would correspond to the respective incident spatial harmonic. Using
405: principle of superposition we obtain that if the semi-infinite
406: crystal is excited by whole spectrum of incident spatial harmonics
407: with amplitudes $\vec E^{s,l}_{\rm inc}$ and wave vectors $\vec
408: k^+_{s,l}$ then the following system of linear equations is valid:
409: \e \=\gamma_{s,l}^{+} \sum\limits_i A_i \frac{\int\limits_V \vec P_i
410: (\vec r') e^{j(\vec k^{+}_{s,l}\.\vec r')}d\vec
411: r'}{1-e^{j(q_x^{(i)}-k^x_{s,l})a}} =\vec E^{s,l}_{\rm inc}
412: \l{ewald3} \f
413: 
414: The scattered field is given by \r{amplscat2} as in the case of one
415: incident wave. The equation \r{ewald3} represents cancelation of all
416: incident spatial spectrum by induced polarization of the crystal in
417: accordance to the formulated above generalized Ewald-Oseen
418: extinction principle.
419: 
420: \subsection{Formulation of boundary conditions} In the previous
421: section we have proved generalized Ewald-Oseen extinction principle
422: for the case of semi-infinite electromagnetic crystal described by
423: certain periodic permittivity distribution $\=\varepsilon(\vec r)$
424: excited by a plane wave coming from free space. In order to extend
425: this theory for the case when incident wave comes from homogeneous
426: isotropic dielectric with permittivity $\epsilon$ it is enough to
427: change $\epsilon_0$ in all formulae to $\epsilon$. Physically it
428: means that we have to consider the polarization of the crystal with
429: respect to the host material with permittivity $\varepsilon$, but
430: not free space. In the model of dense cubic lattice of point dipoles
431: it means that the lattice is located inside of this host material.
432: This approach is very unusual since it can lead to results which are
433: strange from first point of view. For example, free space happen to
434: have negative polarization density with respect to dielectrics with
435: $\varepsilon>\varepsilon_0$. This can be simply explained since free
436: space with respect to these dielectrics is like real materials with
437: $\varepsilon<\varepsilon_0$ with respect to free space: they indeed
438: have negative polarization density.
439: 
440: The meaning of polarization density becomes relative automatically
441: when the replacement of $\varepsilon_0$ to $\varepsilon$ is made. In
442: order to avoid the use of this ambiguous polarization in the final
443: formulae it is possible to express the polarization density in terms
444: of the average field using \r{me}. The resulting expressions provide
445: complete set of boundary conditions for interface between
446: semi-infinite electromagnetic crystal and isotropic dielectric:
447: 
448: \e \vec E(\vec r)=\left\{ \begin{array}{lcl} \sum\limits_{s,l}
449: \left( \vec E^{s,l}_{\rm inc}e^{-j \vec k^-_{s,l}\vec r}+
450: \vec E^{s,l}_{\rm scat}e^{-j \vec k^+_{s,l}\vec r}\right),\ x<a\\
451: \sum\limits_{i} A_i \vec E_i e^{-j\vec q_i\vec r},\ x\ge a \\
452: \end{array} \right.
453: \f
454: 
455: \e \=\gamma_{s,l}^{+} \sum\limits_i A_i \frac{\int\limits_V
456: [\=\varepsilon-\varepsilon_0\=I]\vec E_i (\vec r') e^{j(\vec
457: k^{+}_{s,l}\.\vec r')}d\vec r'}{1-e^{j(q_x^{(i)}-k^x_{s,l})a}}=\vec
458: E_{\rm inc}^{s,l}, \l{bc1} \f
459: 
460: \e \=\gamma_{s,l}^{-} \sum\limits_i A_i \frac{\int\limits_V
461: [\=\varepsilon-\varepsilon_0\=I]\vec E_i (\vec r') e^{j(\vec
462: k^{-}_{s,l}\.\vec r')}d\vec r'}{1-e^{j(q_x^{(i)}+k^x_{s,l})a}}=\vec
463: E^{s,l}_{\rm scat}, \l{bc2} \f where we use following notations
464: $$ \=\gamma^{\pm}_{s,l}=\frac{j}{2bc\varepsilon k_{s,l}} [\-
465: k^{\pm}_{s,l}\times[\- k^{\pm}_{s,l}\times \= I]], \ \vec
466: k^{\pm}_{s,l}=(\pm k^x_{s,l}, k^y_{s}, k^z_{l} )^T,$$
467: $$k^y_{s}=q_y+\frac{2\pi s}{b}, k^z_{l}=q_z+\frac{2\pi l}{c},
468: k^x_{s,l}=\sqrt{k^2-(k^y_{s})^2-(k^z_{l})^2},$$ $k$ is wave number
469: in the dielectric with permittivity $\varepsilon$, $\vec
470: q_i=(q_i^x,k_y,k_z)^T$ and the square root in the expression for
471: $k_{s,l}$ should be chosen so that ${\rm Im}(\sqrt{\.})<0$.
472: 
473: The expressions \r{bc1} and \r{bc2} relate amplitudes of incident
474: $\vec E_{\rm inc}^{s,l}$ and scattered $\vec E_{\rm scat}^{s,l}$
475: spatial harmonics corresponding to tangential wave vector $\vec
476: k_t=(k_y,k_z)^T$ and periodicity of the boundary (rectangular
477: lattice with periods $b$ and $c$) with amplitudes $A_i$ of
478: eigenmodes $(\vec E_i,q^x_i)$ excited in the semi-infinite crystal.
479: 
480: 
481: The presented set of boundary conditions is complete: these
482: equations are enough to determine amplitudes of excited eigenmodes
483: and scattered spatial harmonics if the eigenmodes $\{(\vec
484: E_i,q_i^x)\}$ of infinite crystal corresponding to tangential wave
485: vector $\vec k_t$ are known. But this set is not unique. One can
486: immediately suggest to use classical boundary conditions (continuous
487: tangential component of electric field and normal component of
488: electric displacement at any point of the boundary $x=a$) which
489: being expanded into Fourier series will also grant a complete set of
490: linear equations relating amplitudes of incident, scattered and
491: excited modes.
492: 
493: The advantage of equations \r{bc1} and \r{bc2} as compared to any
494: other boundary conditions is such that they have very special form
495: which can admit analytical solution. We will demonstrate it in the
496: next section for the special case of electromagnetic crystals formed
497: by small scatterers which can be treated as point dipole with fixed
498: orientation. But this is not the only case when an analytical
499: solution of  \r{bc1} and \r{bc2} can be obtained. Recently, M.
500: Silveirinha \cite{MarioABC} demonstrated that the method proposed by
501: ourselves can be successfully applied for studies of reflection from
502: semi-infinite wire medium, material with strong low-frequency
503: spatial dispersion \cite{WMPRB}. Unfortunately, we can not provide
504: solution of equations \r{bc1} and \r{bc2} for the general case.
505: However, we can give some recommendations and an example how these
506: equations can be solved using method of characteristic function. We
507: hope that with some modification this method can be used for other
508: special cases as well.
509: 
510: 
511: \section{Lattice of uniaxial dipolar scatterers}
512: 
513: If the scatterers which form electromagnetic crystal are small as
514: compared to the wavelength then sometimes they can be effectively
515: replaced by point dipoles. It is assumed that the dipole moment of
516: such a dipole is determined by local field acting to the scatterer
517: and the field produced by the scatterer is equal to the field
518: created by the dipole. The polarizability which relates the induced
519: dipole moment and the local field acting to the scatterer is the
520: only parameter which depends on the shape of scatterer in such a
521: local field approach.
522: 
523: Generally, the field produced by any scatterer can be presented
524: using expansion by multipoles. The electric and magnetic dipoles are
525: first and second order multipoles. For some scatterers, the electric
526: or magnetic dipole moments dominate over high-order multipoles. It
527: means that some scatterers behave as electrical dipoles, some other
528: ones as magnetic. Below we will consider only such scatterers. The
529: scatterers which has both electric and magnetic dipole moments of
530: the same order or whose quadrupole or other high-order multipoles
531: can not be neglected are out of scope of our consideration.
532: Moreover, below we will consider only scatterers which can be
533: replaced by dipoles with fixed orientation.
534: 
535: The typical example of the scatterer which behaves as electric
536: dipole with fixed orientation at microwave frequencies is a short
537: metallic cylinder or piece of wire which can be loaded by some
538: inductance in order to increase its polarizability \cite{LWD} (see
539: Fig. \ref{scat}.b). At optical frequencies it can be prolate
540: metallic cylinder which has strong plasmonic resonance. The typical
541: magnetic scatterer at microwave frequencies is split-ring resonator
542: \cite{PendryMagnet} (see Fig. \ref{scat}.a) if the bianisotropic
543: properties of this scatterer are neglected or canceled using method
544: suggested in \cite{MarquesSRR}. At optical frequencies the split
545: metallic rings \cite{YenTHZ,LindenTHZ} behave as magnetic scatterers
546: with fixed orientation of dipole moment. Metallic spheres which also
547: can be replaced by point dipoles are out of scope of our
548: consideration since the orientation of their dipole moments depends
549: on direction of external field.
550: 
551: \begin{figure}[h]
552: \centering \epsfig{file=scat.eps, width=7cm} \caption{Geometries of
553: scatterers which can be modeled as dipoles with fixed orientation:
554: a) split-ring-resonator, b) inductively loaded wire.} \label{scat}
555: \end{figure}
556: 
557: 
558: \subsection{History of the problem}
559: 
560: Further in this section we present an analytical solution for
561: problem of plane wave diffraction on semi-infinite electromagnetic
562: crystals formed by point scatterers with known polarizabilites, but
563: before that we have to describe history of this problem. The first
564: attempt to obtain such an analytical solution has been made by G.D.
565: Mahan and G. Obermair in a seminal work \cite{Mahan}. Analytical
566: expressions for reflection coefficients and amplitudes of excited
567: modes for a semi-infinite crystal were obtained in terms of wave
568: vectors of the infinite crystal eigenmodes. However this theory is
569: not free from the drawbacks. Mahan and Obermair treated the
570: interaction between a reference crystal plane of a semi-infinite
571: crystal and its $N$ nearest neighbors exactly, neglecting the other
572: crystal planes. That is why this approximation is called
573: ``near-neighbor approximation''. Such an approach allows to
574: introduce fictitious zero polarization at the imaginary crystal
575: planes in free space over the semi-infinite crystal. This
576: manipulation gave a set of equations which were treated in
577: \cite{Mahan} as additional boundary conditions. It will be shown
578: below that if the interaction between planes is taken into account
579: exactly but not restricted to finite number of neighboring planes
580: then the fictitious polarization of imaginary planes turns out to be
581: non-zero. In works of C.A. Mead \cite{MeadExactly,MeadFormally} it
582: was already shown that the ``near-neighbor approximation'' appears
583: to be not a strict one. The mead states that the serious
584: disagreement appears in the cases then the interaction between
585: crystal planes falls off not sufficiently fast with distance. In
586: other words, the results of Mahan and Obermair are valid only then
587: the high-order spatial Floquet harmonics produced by the planes
588: rapidly decay with distance. Mahan and Obermair considered only the
589: normal incidence of the plane wave. Within such a restriction their
590: approach is valid in the case when the periods of the structure are
591: small as compared with the wavelength in the host medium. The strong
592: disagreement with the exact solution appears when one of high-order
593: Floquet harmonic happens to be propagating one. This fact is
594: illustrated below by a numerical comparison.
595: 
596: The work \cite{Mahan} caused numerous extensions
597: \cite{PhilpottRef,PhilpottSlab,PhilpottPRB}. The Mahan and Obermair
598: approach was generalized for the cases of oblique incidence
599: \cite{PhilpottRef}, both possible polarizations of incident wave
600: \cite{PhilpottPRB}, various lattice structures of the crystal
601: \cite{PhilpottRef}, tensorial polarizability of scatterers
602: \cite{PhilpottPRB} and even diffraction of the finite-size slabs of
603: the crystals was considered \cite{PhilpottSlab}. Note, that all the
604: listed works use the same ``near-neighbor approximation'' and their
605: applicability is restricted as described above. In order to avoid
606: this trouble one needs to use another model for interaction between
607: crystal planes. The simplest one is the so-called ``exp model''
608: suggested by Mead \cite{MeadExactly,MeadExp2} which assumes that
609: interaction can be described by a single decaying exponent. In terms
610: of spatial Floquet harmonics this approach is equivalent to
611: neglecting all high-order Floquet harmonics except the one with the
612: slowest decay. The ``exp model'' as well as the ``near-neighbor
613: approximation'' allow to solve the problem of excitation
614: analytically for both normal \cite{MeadExactly} and oblique
615: \cite{MeadExp2} incidences. The ``exp model'' of Mead gives a set of
616: two equations which correspond to the generalized Ewald-Oseen
617: extinction principle formulated in the present paper. The first
618: equation of Mead is the same as one of equations given by
619: ``near-neighbor approximation''. It describes the fact that the
620: incident electromagnetic wave (fundamental Floquet harmonic) inside
621: the semi-infinite crystal is canceled by induced polarization of the
622: crystal. This fact was pointed out in the papers
623: \cite{Mahan,PhilpottPRB,MeadExactly}. The second equation clearly
624: expresses the fact that induced polarization cancels also the second
625: Floquet harmonic (taken into account in the ``exp model'') of
626: incident wave which has zero amplitude, but unfortunately it was not
627: noted by the authors. The system of these two equations is solved in
628: \cite{MeadExactly} and the amplitudes of excited eigenmodes and an
629: expression for reflection coefficient are obtained.
630: 
631: It is possible to modify the ``exp model'' in order to obtain an
632: exact solution. For that purpose one simply should take into account
633: all Floquet harmonics in the interaction between the crystal planes.
634: This has been done by authors of the present paper and the results
635: are presented below. As it was shown above, it turns out that every
636: incident Floquet harmonic (even if it has zero amplitude) is
637: canceled by the induced polarization following to the generalized
638: Ewald-Oseen extinction principle. It provides an infinite system of
639: equations relating amplitudes of excited eigenmodes. This system can
640: be truncated and then the number of equations in the system turns
641: out to be equal to the number of Floquet harmonics taken into
642: account. Such the finite system can be easily solved analytically
643: for the case when only two Floquet harmonics are taken into account
644: (this is the ``exp model'' of Mead \cite{MeadExactly}), but in the
645: case when one would like to take into account more Floquet harmonics
646: this approach requires numerical calculations. We avoid the
647: truncation of the system of equations and offer a closed-form
648: rigorous analytical solution which is simple and explicit.
649: 
650: Note, that a ``formally closed solution'' for the problem under
651: consideration was proposed by Mead in \cite{MeadFormally}. In this
652: solution there is a contour integral of a certain function given in
653: the form of infinite series. However, the calculation with help of
654: such ``formally closed solution'' requires serious numerical
655: efforts. The main idea of work \cite{MeadFormally} is based on
656: introduction of characteristic analytical function  which allows to
657: determine all parameters entering the expression for the reflection
658: coefficient. It is shown that knowledge of its roots allows to
659: recover this function and obtain analytical expressions for all
660: amplitudes of excited eigenmodes and for the reflection coefficient,
661: consequently. Unfortunately, these roots were not found in
662: \cite{MeadFormally}. That is why the contour integration was used in
663: \cite{MeadFormally} in order to bypass the problem of these roots
664: finding. In fact, as it is shown below, the roots of this
665: characteristic analytical function are determined by the wave
666: vectors of Floquet harmonics and can be easily expressed
667: analytically. This fact is a consequence of generalized Ewald-Oseen
668: extinction principle which is an important point of our theory.
669: 
670: One could directly apply general results \r{ewald2} and
671: \r{amplscat2} to the problem under consideration. It is enough to
672: replace polarization density of eigenmodes by three-dimensional
673: delta function corresponding to the point of location of the dipolar
674: scatterer in the unit cell and one could obtain the system of linear
675: equations \r{ewald2} and \r{amplscat2} which correspond to the
676: boundary conditions in the present case as it is shown below.
677: However, we prefer to re-deduce all the expressions by making the
678: same steps as in the previous section but for the case of point
679: scatterers. We suppose that it is very useful step in order to
680: demonstrate physical background of the computations undertaken in
681: the previous section at a specific example.
682: 
683: \subsection{Dispersion equation}
684: 
685: Let us consider an infinite crystal formed by point dielectric
686: dipoles with some known polarizability $\alpha$ along fixed
687: direction given by unit vector $\vec d$, $\=\alpha=\alpha \vec d\vec
688: d$. The case of magnetic dipoles can be easily obtained from the
689: theory for dielectric ones using duality principle. The scatterers
690: are arranged in the nodes of the three-dimensional lattice with an
691: orthorhombic elementary cell $a\times b\times c$ located in free
692: space, see Figure \ref{geomsm}.
693: 
694: \begin{figure}[h]
695: \centering \epsfig{file=infpc.eps, width=7.5cm} \caption{Geometry of
696: an infinite electromagnetic crystal formed by uniaxial dipolar
697: scatterers} \label{geomsm}
698: \end{figure}
699: 
700: The distribution of dipole moments corresponding to an eigenmode
701: with wave vector $\vec q=(q_x,q_y,q_z)^T$ is described as $\vec
702: p_{n,s,l}=\vec p e^{-j(q_xan+q_ybs+q_zcl)}$, where $n,s,l$ are
703: integer indices of scatterers along the $x-,y-,z-$ axes,
704: respectively, and $\vec p$ is a dipole moment of the scatterer
705: located at the center of coordinate system. Following to the local
706: field approach $\vec p$ can be expressed as $\vec p=\alpha (\vec
707: E_{\rm loc}\.\- d)\- d$, where $\vec E_{\rm loc}$ is a local
708: electric field acting to the scatterer. The local field is produced
709: by all other scatterers which form the infinite crystal and it can
710: be given by the formula \e \vec E_{\rm
711: loc.}={\sum\limits_{n,s,l}}'\=G(\vec R_{n,s,l})\vec p_{n,s,l}, \f
712: where $\=G(\vec r)$ is the three-dimensional dyadic Green function
713: of the free space \r{gfree} and summation is taken over all triples
714: of indices except the zero one. Accordingly the following dispersion
715: equation for the crystal under consideration is obtained (compare
716: with \r{disp2}): \e \alpha^{-1}=\left[{\sum\limits_{n,s,l}}'\=G(\vec
717: R_{n,s,l})e^{-j(q_xan+q_ybs+q_zcl)}\-d\right]\.\-d. \l{gendis} \f
718: 
719: In order to evaluate sums of series in \r{gendis} we use a
720: plane-wise approach \cite{nonresPRE}. According to this approach the
721: dispersion equation takes the following form: \e
722: \alpha^{-1}=\sum\limits_{n=-\infty}^{+\infty} \beta_n e^{-jq_xan}.
723: \l{layerdisp} \f
724: 
725: The coefficients $\beta_n$ describe the interaction between planes
726: and include the information on transverse wave vector components
727: $q_y$, $q_z$ as well as on a geometry of a single plane (lattice
728: periods $b$, $c$). For $n\ne 0$ coefficients $\beta_n$ can be
729: expressed using expansion by Floquet harmonics. For $n=0$ the
730: calculation of coefficient $\beta_0$ (describing interaction inside
731: of a plane and expressed in the form of two-dimensional series
732: without the zero term) requires additional efforts (see
733: \cite{nonresPRE,Belovhomo,Collin} for details).
734: 
735: The electric field produced by a single plane (namely
736: two-dimensional grid $b\times c$ of point dipoles with the
737: distribution $\vec p_{s,l}=\vec p e^{-j(q_ybs+q_zcl)}$) located in
738: the plane $x=0$ is equal to \e \vec E(\vec r) =
739: \frac{j}{2bc\varepsilon_0}\sum\limits_{s,l} [\- k^{{\rm
740: sign}(x)}_{s,l}\times[\- k^{{\rm sign}(x)}_{s,l}\times \vec p]]
741: \frac{e^{-j(\vec k^{{\rm sign}(x)}_{s,l}\.\vec R)}}{k^x_{s,l}},
742: \l{genfloquet} \f where $\vec k^{\pm}_{s,l}=(\pm k^x_{s,l}, k^y_{s},
743: k^z_{l} )^T$, $k^y_{s}=q_y+\frac{2\pi s}{b}$,
744: $k^z_{l}=q_z+\frac{2\pi l}{c}$,
745: $k^x_{s,l}=\sqrt{k^2-(k^y_{s})^2-(k^z_{l})^2}$ and$k$ is wave number
746: of free space. One should choose the square root in the expression
747: for $k_{s,l}$ so that ${\rm Im}(\sqrt{\.})<0$. The sign $\pm$
748: corresponds to half spaces $x>0$ and $x<0$ respectively.
749: 
750: The formula \r{genfloquet} defines an expansion of the field
751: produced by a single grid of dipoles in terms of plane waves and it
752: can be obtained using double Poisson summation formula to series of
753: fields produced by single scatterers in free space. These plane
754: waves have wave vectors $\vec k^{\pm}_{s,l}$. They are also called
755: Floquet harmonics and represent a spatial spectrum of the field
756: (compare with \r{genfloquet2}). Floquet harmonics are widely used in
757: analysis of phased array antennas \cite{Amitay}.
758: 
759: Using \r{genfloquet} we get the following expression for $\beta_n$
760: ($n\ne 0$): \e \beta_n=\sum\limits_{s,l}\gamma^{-{\rm
761: sign}(n)}_{s,l} e^{-jk^x_{s,l}a|n|}, \l{floquet} \f where
762: $\gamma^\pm_{s,l}=[k^2-(\- k^\pm_{s,l}\.\-d)^2]/(2jbc\varepsilon_0
763: k^x_{s,l})$. After substitution of \r{floquet} into \r{layerdisp},
764: changing the order of summation and using the formula for sum of
765: geometrical progression we obtain the dispersion equation in the
766: following form: \e \alpha^{-1}=\beta_0 + \sum\limits_{s,l} \left[
767:  \frac{\gamma^{-}_{s,l}}{e^{j(k^x_{s,l}+q_x)a}-1}+
768:  \frac{\gamma^{+}_{s,l}}{e^{j(k^x_{s,l}-q_x)a}-1}\right].
769: \l{disp} \f This is a transcendental equation expressed by rapidly
770: convergent series. Dispersion properties of the crystal under
771: consideration can be studied with help of numerical solution of the
772: latter equation. The case of dipoles oriented along one of the
773: crystal axes of the orthorhombic crystal has been considered in
774: \cite{Belovhomo}, the dispersion equation was solved and typical
775: dispersion curves and iso-frequency contours for resonant scatterers
776: were presented.
777: 
778: 
779: \subsection{Semi-infinite crystal}
780: Now let us consider a semi-infinite electromagnetic crystal, the
781: half-space $x\ge a$ filled by the crystal formed by point dipoles,
782: see Figure \ref{semigeom}.
783: \begin{figure}[ht]
784: \centering \epsfig{file=semiinfpc.eps, width=7.5cm}
785: \caption{Geometry of a semi-infinite electromagnetic crystal formed
786: by uniaxial dipolar scatterers} \label{semigeom}
787: \end{figure}
788: The structure is excited by a plane electromagnetic wave with the
789: wave vector $\-k=(k_x,k_y,k_z)^T$ and the intensity of electric
790: field $\- E_{\rm inc}$. Let us denote the component of the incident
791: electric field along the direction of dipoles as $E_{\rm inc}=(\-
792: E_{\rm inc}\.\-d)$. The axis $x$ is assumed to be normal to the
793: interface. The tangential (with respect to the interface)
794: distribution of dipole moments in excited semi-infinite crystal is
795: determined by the tangential component of the incident wave vector.
796: It means that $p_{n,s,l}=p_{n}e^{-j(k_ybs+k_zcl)}$, where the
797: polarizations of zero-numbered scatterers from planes with the index
798: $n$ (parallel to the interface) are denoted as $p_n=p_{n,0,0}$. The
799: plane-to-plane distribution $\{p_n\}$ is unknown and it has to be
800: found. Using the local field approach one can write the infinite
801: linear system of equations for this distribution: \e p_m =\alpha
802: \left(E_{\rm inc}e^{-jk_xam}+\sum\limits_{n=1}^{+\infty} \beta_{n-m}
803: p_n \right), \ \forall\ m\ge 1 . \l{layer} \f The distribution
804: $\{p_n\}$ of polarization in the excited semi-infinite crystal can
805: be determined solving the system of equations \r{layer}. The known
806: distribution of polarization allows to determine the scattered field
807: in the half-space $x<a$ with help of the expansion by Floquet
808: harmonics \r{genfloquet}: \e \vec E = \sum\limits_{s,l} \vec E_{s,l}
809: e^{-j(\vec k^-_{s,l}\.\vec R)}, \l{diffr} \f where the amplitudes of
810: Floquet harmonics are following: \e \vec E_{s,l} =
811: \frac{j}{2ab\varepsilon_0 k^x_{s,l}}[\- k^-_{s,l}\times[\-
812: k^-_{s,l}\times \vec d]] \sum\limits_{n=1}^{+\infty} p_n
813: e^{-jk^x_{s,l}an}. \l{ampl} \f
814: 
815: If the crystal supports propagating modes, it is quite difficult to
816: find a solution of \r{layer} numerically. Simple methods such as
817: system truncating (considering a slab with finite thickness instead
818: of half-space like in \cite{flo}) results in nonconvergent
819: oscillating solutions which have nothing to do with actual solution
820: of \r{layer}.
821: 
822: \subsection{Expansion by eigenmodes}
823: In order to solve \r{layer} accurately one has to use an expansion
824: of the polarization by eigenmodes \cite{Mahan}: \e p_n
825: =\sum\limits_{i} A_i e^{-jq_x^{(i)}an}, \l{expansion} \f where $A_i$
826: are amplitudes of eigenmodes and $q_x^{(i)}$ are the $x$-components
827: of their wave vectors. Every eigenmode is assumed to be a solution
828: of the dispersion equation \r{disp} with the wave vector $\vec
829: q_i=(q_x^{(i)},k_y,k_z)^{T}$. In the formula \r{expansion} the
830: summation is taken by eigenmodes which either transfer energy into
831: half-space $x\ge a$ ($\frac{dq_x^{(i)}}{d\omega}>0$) or decay along
832: $x-$axis (${\rm Im}(q_x^{(i)})<0$).
833: 
834: Let us assume that the dispersion equation \r{disp} is solved (for
835: example numerically) and the necessary set of eigenmodes
836: $\{q_x^{(i)}\}$ is found. Then the substitution of \r{expansion}
837: into \r{layer} will replace the set of unknown  polarizations of
838: planes by a set of unknown amplitudes of eigenmodes: \e
839: \alpha^{-1}\sum\limits_{i} A_i e^{-jq_x^{(i)}am} = E_{\rm
840: inc}e^{-jk_xam}+\sum\limits_{n=1}^{+\infty} \beta_{n-m}
841: \sum\limits_{i} A_i e^{-jq_x^{(i)}an}. \l{st} \f Applying the
842: auxiliary relation evidently following from \r{disp}: \e \alpha^{-1}
843: e^{-jq_x^{(i)}am}-\sum\limits_{n=-\infty}^{0} \beta_{n-m}
844: e^{-jq_x^{(i)}an}= \sum\limits_{n=1}^{+\infty} \beta_{n-m}
845: e^{-jq_x^{(i)}an}, \l{aux} \f the equation \r{st} can be transformed
846: as follows: \e \sum\limits_{i} A_i \left(\sum\limits_{n=-\infty}^{0}
847: \beta_{n-m} e^{-jq_x^{(i)}an} \right)=E_{\rm inc}e^{-jk_xam}.
848: \l{eigenamp} \f
849: 
850: It should be noted, that using definition of Mahan and Obermair for
851: the polarization of fictitious planes ($p_n=\sum\limits_{i} A_i
852: e^{-jq_x^{(i)}an}, \ \forall n\le 0$) one can rewrite \r{eigenamp}
853: as \e \sum\limits_{n=-\infty}^{0} \beta_{n-m} p_n=E_{\rm
854: inc}e^{-jk_xam}. \l{fictitious} \f It is evident that the assumption
855: of Mahan and Obermair, requiring all polarizations of fictitious
856: planes to be zeros, contradicts with \r{fictitious}. This fact
857: proves that the ``near-neighbor approximation'' made in \cite{Mahan}
858: is not accurate.
859: 
860: The system of equations \r{eigenamp} can be truncated and solved
861: numerically quite easily in contrast to \r{layer}. As a result, the
862: amplitudes of eigenmodes $\{A_i\}$ can be found and the polarization
863: distribution can be restored using formula \r{expansion}. The
864: amplitudes of scattered Floquet harmonics \r{ampl} can be also
865: expressed in terms of excited eigenmodes amplitudes by means of
866: substitution of \r{expansion} into \r{ampl}, changing the order of
867: summation and evaluating sums of geometrical progressions. The final
868: expression for the amplitudes of scattered Floquet harmonics is
869: following (compare with \r{amplscat2}) : \e \vec E_{s,l} = \frac{[\-
870: k^-_{s,l}\times[\- k^-_{s,l}\times \vec d]]}{2jab\varepsilon_0
871: k^x_{s,l}} \sum\limits_{i}
872: A_i\frac{1}{1-e^{j(q_x^{(i)}+k^x_{s,l})a}}. \l{ampl2} \f
873: 
874: One can stop at this stage and claim that the problem of the
875: semi-infinite electromagnetic crystal excitation is solved. However
876: in this case the solution would require long numerical calculations,
877: such as solving the system \r{eigenamp} and substituting the
878: obtained solution into \r{ampl2}. The possibility to make all
879: described operations analytically in the closed-form is shown below.
880: 
881: \subsection{Analytical solution}
882: Substituting the expansion \r{floquet} into \r{eigenamp} we obtain:
883: $$\sum\limits_{i} A_i \left(\sum\limits_{n=-\infty}^{0}
884: \left[\sum\limits_{s,l}\gamma^+_{s,l} e^{-jk^x_{s,l}a|n-m|}\right]
885: e^{-jq_x^{(i)}an} \right)$$ \e =E_{\rm inc}e^{-jk_xam}. \l{big} \f
886: Changing the order of summation in \r{big}, taking into account that
887: $n-m<0$ and using formula for the sum of geometrical progression we
888: obtain: \e \sum\limits_{s,l}\gamma^+_{s,l}  \left(\sum\limits_{i}
889: A_i \frac{1}{1-e^{j(q_x^{(i)}-k^x_{s,l})a}}\right)
890: e^{-jk^x_{s,l}am}=E_{\rm inc}e^{-jk_xam} \l{syst} \f This is a
891: system of linear equations where unknowns are given by expressions
892: in brackets. It has a unique solution because the determinant of the
893: system has finite nonzero value. Note that $k^x_{s,l}=k_x$ only if
894: $(s,l)=(0,0)$. Thus, the solution of \r{syst} has the following
895: form: \e \sum\limits_{i} A_i
896: \frac{1}{1-e^{j(q_x^{(i)}-k^x_{s,l})a}}=\left\{\begin{array}{lcl}
897: E_{\rm inc}/\gamma^+_{0,0},\ \mbox{if}\  (s,l)=(0,0)\\ 0, \
898: \mbox{if}\ (s,l)\ne (0,0)\\ \end{array}\right. \l{ewald} \f
899: 
900: This equation could be directly obtained from general expression
901: \r{ewald2} by substitution of delta function instead of polarization
902: density of eigenmodes, but as we already mentioned above we
903: intentionally re-deduced it since we suppose that it can help to
904: understand background for deduction of \r{ewald2} for general case.
905: The values at the right side of \r{ewald} are the normalized
906: amplitudes of incident Floquet harmonics, and the series at the left
907: side are the normalized amplitudes of Floquet harmonics produced by
908: the whole semi-infinite crystal polarization which cancel the
909: incident harmonics. Thus, the equation \r{ewald} represents the
910: generalization of the  Ewald-Oseen extinction principle already
911: formulated above for general case of semi-infinite crystals: {\it
912: The polarization in a semi-infinite electromagnetic crystal excited
913: by a plane wave is distributed in such a way that it cancels the
914: incident wave together with all high-order spatial harmonics
915: associated with periodicity of the boundary}.
916: 
917: Note, that the formula for the amplitudes of scattered Floquet
918: harmonics \r{ampl2} contains series that have the same form as
919: \r{ewald}, but another sign in front of $k^x_{s,l}$.
920: 
921: The amplitudes of the excited modes $A_i$ can be found numerically
922: from the infinite set of equations \r{ewald} and substitution of
923: $A_i$ into \r{ampl2} will give us amplitudes of scattered Floquet
924: harmonics. However, it is possible to obtain a closed-form
925: analytical solution of \r{ewald}.
926: 
927: In order to solve the set of equations \r{ewald} one should consider
928: a characteristic function $f(x)$ (see also \cite{MeadFormally}) of
929: the form : \e f(u)=u\sum\limits_{i} A_i \frac{1}{u-e^{jq_x^{(i)}a}}.
930: \l{f} \f Comparing \r{ampl2} and \r{ewald} with \r{f} one can see
931: that the function $f(u)$ has the following properties:
932: \begin{itemize}
933: \item It has poles at $u=e^{jq_x^{(i)}a}$
934: \item It has roots at $u=e^{jk^x_{s,l}a}$, $(s,l)\ne (0,0)$ and $u=0$
935: \item It has a known value $E_{\rm inc}/\gamma^+_{0,0}$ at $u=e^{jk_xa}$
936: \item Its values at $u=e^{-jk^x_{s,l} a}$ are equal to the normalized amplitudes of scattered Floquet harmonics
937: \item Its residues at $u=e^{jq_x^{(i)}a}$ are equal to the normalized amplitudes of excited eigenmodes.
938: \end{itemize}
939: 
940: It is possible to restore the function $f(u)$ using the known values
941: of its poles, roots and a value at one point: \e f(u)=\frac{E_{\rm
942: inc}u}{\gamma^+_{0,0} e^{jk_xa}} \prod\limits_{(s,l)\ne (0,0)}
943: \frac{u-e^{jk^x_{s,l}a}}{e^{jk_xa}-e^{jk^x_{s,l}a}} \prod\limits_i
944: \frac{e^{jk_xa}-e^{jq_x^{(i)}a}}{u-e^{jq_x^{(i)}a}}. \f
945: 
946: The knowledge of the characteristic function $f(u)$ provide us with
947: complete solution of our diffraction problem. The amplitudes of
948: excited eigenmodes with indices $n$ are equal to residues of $f(u)$
949: at $u=e^{jq_x^{(n)}a}$:
950: $$
951: A_n={\rm Res} f(u) \left|_{u=e^{jq_x^{(i)}a}} \vphantom{\frac{a}{a}}
952: \right.=\frac{E_{\rm
953: inc}(1-e^{j(q_x^{(n)}-k_x)a})}{\gamma^+_{0,0}}\times
954: $$
955: \e \times \prod\limits_{(s,l)\ne (0,0)}
956: \frac{e^{jq_x^{(n)}a}-e^{jk^x_{s,l}a}}{e^{jk_xa}-e^{jk^x_{s,l}a}}
957: \prod\limits_{i\ne n}
958: \frac{e^{jk_xa}-e^{jq_x^{(i)}a}}{e^{jq_x^{(n)}a}-e^{jq_x^{(i)}a}},
959: \l{A} \f and the amplitudes of scattered Floquet harmonics with
960: indices $(r,t)$ can be expressed through values of $f(u)$ at
961: $u=e^{-jk^x_{r,t} a}$:
962: $$
963: \-E_{r,t}= \frac{E_{\rm inc}e^{-jk^x_{r,t}a}[\- k^-_{r,t}\times[\-
964: k^-_{r,t}\times \vec d]]}{2jab\varepsilon_0 k^x_{r,t}\gamma^+_{0,0}
965: e^{jk_xa}} \times
966: $$
967: \e \times \prod\limits_{(s,l)\ne (0,0)}
968: \frac{e^{-jk^x_{r,t}a}-e^{jk^x_{s,l}a}}{e^{jk_xa}-e^{jk^x_{s,l}a}}
969: \prod\limits_i
970: \frac{e^{jk_xa}-e^{jq_x^{(i)}a}}{e^{-jk^x_{r,t}a}-e^{jq_x^{(i)}a}}.
971: \l{R} \f
972: 
973: The products in the formulae \r{A} and \r{R} have very rapid
974: convergence. It is enough to take a few terms in order to reach
975: excellent accuracy. The main requirement for truncation of these
976: infinite products is to take into account all terms corresponding to
977: propagating $\mbox{Re}(k^x_{s,l})=0$ and slowly decaying
978: $\mbox{Im}(k^x_{s,l})\ll 2\pi/a$ Floquet harmonics as well as
979: propagating $\mbox{Re}(q_x^{(i)})=0$ and slowly decaying
980: $\mbox{Im}(q_x^{(i)})\ll 2\pi/a$ eigenmodes.
981: 
982: \subsection{Comparison with other theories}
983: 
984: Let us consider the case from work \cite{Mahan} when $\-d=\-y_0$,
985: and $a=b=c$. In this case the formula \r{R} for the fundamental
986: Floquet harmonic ($r=t=0$) can be rewritten in terms of the
987: reflection coefficient: \e R= -e^{-2jk_xa} \prod\limits_{(s,l)\ne
988: (0,0)} \frac{e^{-jk_xa}-e^{jk^x_{s,l}a}}{e^{jk_xa}-e^{jk^x_{s,l}a}}
989: \prod\limits_i
990: \frac{e^{jk_xa}-e^{jq_x^{(i)}a}}{e^{-jk_xa}-e^{jq_x^{(i)}a}}.
991: \l{Rmain} \f
992: 
993: Comparing that result with the final result of the work \cite{Mahan}
994: (the next formula after (C7) on page 841) one can see that the first
995: product in our formula \r{Rmain} \e \Pi=\prod\limits_{(s,l)\ne
996: (0,0)} \frac{e^{-jk_xa}-e^{jk^x_{s,l}a}}{e^{jk_xa}-e^{jk^x_{s,l}a}}
997: \l{Pi} \f is absent  in the result of Mahan and Obermair. This
998: difference is a consequence of the fact that in our study we
999: considered interaction between crystal planes accurately taking into
1000: account all Floquet harmonics for any distance between planes in
1001: contrast to the ``near-neighbor approximation'' used in the approach
1002: of Mahan and Obermair.
1003: 
1004: \begin{figure}[ht]
1005: \vspace{5mm}
1006: \centering
1007: \epsfig{file=prod.eps, width=7.5cm}
1008: \caption{Dependence of $\Pi$ vs. normalized frequency $ka/(2\pi)$}
1009: \label{prod}
1010: \end{figure}
1011: 
1012: The dependence of the product $\Pi$ vs. normalized frequency is
1013: plotted in Figure \ref{prod} for the case of normal incidence
1014: $k_y=k_z=0$ and $k_x=k$. One can see that the value of the product
1015: is nearly equal to the unity for $ka<1.6\pi$, but for $ka>2\pi$ the
1016: value of the product significantly differs from the unity. Thus we
1017: conclude that the theory of Mahan and Obermair is valid in the low
1018: frequency range when periods of the lattice are small compared to
1019: the wavelength. Our theory does not have such a restriction (within
1020: the frame of the dipole model of electromagnetic crystal).
1021: 
1022: The comparison with results of \cite{MeadFormally} shows that
1023: \r{Rmain} is equivalent to formula (46) from \cite{MeadFormally}
1024: with $\Pi=\exp(\Gamma)$ where $\Gamma$ is given by the contour
1025: integral (47) from \cite{MeadFormally}. The calculation of $\Pi$
1026: using \r{Pi} requires taking into account only a few terms in the
1027: infinite products, because they are very rapidly convergent. This is
1028: a significant advantage of our approach as compared to work
1029: \cite{MeadFormally} which requires complicate numerical calculation
1030: of the contour integral.
1031: 
1032: In the long-wavelength limit, the series in equation \r{gendis} for
1033: the cubic lattice can be replaced by the integral taken over the
1034: whole space except unit cell $V$ and we obtain: \e
1035: \alpha^{-1}=\left[\left(\int\limits_{R^3\slash V}\=G(\vec
1036: R)e^{-j(\vec q\. \vec R)} d\vec R\right)\-d\right]\.\-d. \l{disint}
1037: \f
1038: 
1039: The integral in the right-hand side of equation \r{disint} can be
1040: evaluated by means of the same technique that was used while
1041: deducing Ewald-Oseen extinction principle in \cite{BornWolf}. The
1042: result is following: \e \alpha^{-1}=\left[\frac{1}{3}+\frac{(\vec
1043: q\. \vec d)^2-|\vec q|^2}{K^2-|\vec
1044: q|^2}\right]\frac{V}{\varepsilon_0}. \l{disint2} \f
1045: 
1046: The obtained dispersion equation \r{disint2} can be transformed in
1047: the common form: \e \tilde\varepsilon (k^2-q_d^2)=\varepsilon_0
1048: (|\vec q|^2-q_d^2), \l{disuni} \f where \e
1049: \tilde\varepsilon=\varepsilon_0\left(1+\frac{\alpha/(\varepsilon_0
1050: V)}{1-\alpha/(3\varepsilon_0 V)}\right), \l{epseff} \f and
1051: $q_d=(\vec q\. \vec d)$ is the component of the wave vector $\vec q$
1052: along the anisotropy axis.
1053: 
1054: The formula \r{disuni} is classical form of the dispersion equation
1055: for uniaxial dielectrics \cite{BornWolf} with permittivity
1056: $\tilde\varepsilon$ along the anisotropy axis and $\varepsilon$ in
1057: the transverse plane. The expression \r{epseff} is the
1058: Clausius-Mossotti formula for the effective permittivity of cubic
1059: lattices of scatterers.
1060: 
1061: In the long-wavelength limit the formula \r{R} for amplitude of
1062: reflected wave simplifies as follows: \e \-E_R= -\frac{(\vec E_{\rm
1063: inc}\.\vec d) [\- k^-\times[\- k^-\times \vec d]]}{[k^2-(\-
1064: k\.\-d)^2]} \frac{k_x-q_x}{k_x+q_x}, \l{Rlong} \f where $\vec
1065: k^-=(-k_x,k_y,k_z)^T$ is wave vector of reflected wave. The formula
1066: \r{Rlong} represents a compact form of an expression for electric
1067: field amplitude of a wave reflected from an interface between an
1068: isotropic dielectric and an uniaxial dielectric (see, e.g.
1069: \cite{Fedorov}). Note, that in our case the situation is simplified
1070: as compared to the general case, because the incident wave comes
1071: from isotropic dielectric with permittivity $\varepsilon$ which is
1072: equal to the permittivity of uniaxial dielectric in transverse
1073: plane. It means, that an incident wave with normal polarization with
1074: respect to the anisotropy axis transforms at the interface in a
1075: refracted ordinary wave without reflection.
1076: 
1077: Let us consider the reflection problem at the special case when
1078: $\-d=\-y_0$, $k_y=0$ and $\vec E_{\rm inc}\parallel \-y_0$. The
1079: nonzero components of the wave vector for the incident wave can be
1080: expressed in terms of incident angle $\theta_i$ as
1081: $k_z=k\sin\theta_i$ and $k_x=k\cos\theta_i$. From \r{disuni} we
1082: obtain that in this case the transmitted wave has $x$-component of
1083: wave vector equal to $q_x=\sqrt{\tilde\varepsilon
1084: k^2/\varepsilon_0-k_z^2}=\sqrt{\tilde\varepsilon/\varepsilon_0}k\cos\theta_t$,
1085: where $\theta_t$ in angle or refraction. With the result \r{Rlong}
1086: we get the reflection coefficient in the form: \e
1087: R=\frac{k_x-q_x}{k_x+q_x}=\frac{n_1\cos\theta_i-n_2\cos\theta_t}{n_1\cos\theta_i+n_2\cos\theta_t},
1088: \l{Fresnel} \f where $n_1=\sqrt{\varepsilon_0\mu_0}$ and
1089: $n_2=\sqrt{\tilde\varepsilon\mu_0}$ are indices of refraction of the
1090: materials. The formula \r{Fresnel} coincide with the classical
1091: Fresnel equation \cite{BornWolf}. This fact can be treated as an
1092: additional verification of presented theory.
1093: 
1094: \section{Lattice of split-ring-resonators}
1095: 
1096: In this section we apply theory presented in previous sections for
1097: study of reflection from semi-infinite cubic lattice of
1098: split-ring-resonators.
1099: 
1100: The general dispersion equation \r{disp2} of the integral form in
1101: the case of point electric scatterers transforms into transcendental
1102: equation \r{disp}. In the case when $\vec d=\vec y_0$ the dispersion
1103: equation \r{disp} can be rewritten in the following closed form
1104: convenient for numerical calculations (see \cite{Belovhomo} for
1105: details): \e \varepsilon_0\alpha^{-1}(\omega)=C(k,\-q,a,b,c),
1106: \l{disphomo}\f where $C(k,\-q,a,b,c)$ is dynamic interaction
1107: constant of the form:
1108: $$
1109: C(k,\-q,a,b,c)= -\sum\limits_{l=1}^{+\infty}\sum\limits_{{\rm
1110: Re}(p_s)\ne 0} \frac{p_s^2}{\pi b}K_0\left(p_scl\right)\cos(q_zcn)
1111: $$
1112: \e
1113: +\sum\limits_{s=-\infty}^{+\infty}\sum\limits_{l=-\infty}^{+\infty}
1114: \frac{p_s^2}{2jbc k^x_{s,l}} \frac{e^{-j k^x_{s,l} a}-\cos
1115: q_xa}{\cos k^x_{s,l} a-\cos q_xa} \l{cfinal} \f
1116: $$
1117: -\sum\limits_{{\rm Re}(p_s)=0} \frac{p_s^2}{2bc}
1118: \left(\frac{1}{jk^x_{s,0}}+ \sum\limits_{l=1}^{+\infty}
1119: \left[\frac{1}{jk^x_{s,l}}+\frac{1}{jk^x_{s,-l}}\right. \right.
1120: $$
1121: $$
1122: \left.\left. -\frac{c}{\pi l}-\frac{r_sc^3}{8\pi^3 l^3}\right]+1.202
1123: \frac{r_sc^3}{8\pi^3}+ \frac{c}{\pi} \left(\log
1124: \frac{c|p_s|}{4\pi}+\gamma\right)+j\frac{c}{2}
1125: \vphantom{\sum\limits_{n\ne 0}
1126: \left[\frac{1}{jk^x_{s,l}}-\frac{c}{2\pi |l|}\right]}\right)
1127: $$
1128: $$
1129: +\frac{1}{4\pi b^3} \left[ 4\sum\limits_{s=1}^{+\infty}
1130: \frac{(2jkb+3)s+2}{s^3(s+1)(s+2)}e^{-jkbs}\cos(q_ybs) \right.
1131: $$
1132: $$
1133: -(jkb+1)\left(t_+^2\log t^++t_-^2\log t^-+2e^{jkb}\cos (q_yb)\right)
1134: $$
1135: $$
1136: \left. -2jkb\left(t_+\log t^++t_-\log t^-\right) +(7jkb+3)
1137: \vphantom{\sum\limits_{s=1}^{+\infty}
1138: \frac{(2jkb+3)s+2}{s^3(s+1)(s+2)}} \right],
1139: $$
1140: and the following notations are used:
1141: $$
1142: p_s=\sqrt{\left(k^y_s\right)^2-k^2},\qquad r_s=2q_z^2-p_s^2,
1143: $$
1144: $$
1145: t^+=1-e^{-j(k+q_z)c}, \qquad t^-=1-e^{-j(k-q_z)c},
1146: $$
1147: $$
1148: t_+=1-e^{j(k+q_z)c}, \qquad t_-=1-e^{j(k-q_z)c}.
1149: $$
1150: 
1151: The calculations using \r{cfinal} can be restricted to the real part
1152: of dynamic interaction constant $C(k,\-q,a,b,c)$  only, because its
1153: imaginary part is given by much simpler expression (see
1154: \cite{Belovhomo} for details): \e {\rm
1155: Im}\left\{C(k,\-q,a,b,c)\right\}=\frac{k^3}{6\pi}. \l{imc} \f The
1156: series in \r{cfinal} have excellent convergence that ensure very
1157: rapid numerical calculations.
1158: 
1159: The case of magnetic scatterers can be considered using duality
1160: principle. The expression \r{Rmain} can be used for calculation of
1161: reflection coefficient by magnetic field (originally, this equation
1162: represented reflection coefficient by electric field). The
1163: dispersion equation \r{disphomo} have to be rewritten for the case
1164: of magnetic point scatterers in the following form: \e
1165: \mu_0\alpha_m^{-1}(\omega)=C(k,\-q,a,b,c), \l{disphomom}\f where
1166: $\alpha_m(\omega)$ is magnetic polarizability of the scatterers. The
1167: analytical expressions for the magnetic polarizability
1168: $\alpha(\omega)$ of split-ring-resonators with geometry plotted in
1169: Fig.\ref{scat} were derived and validated in \cite{SimSRR}. The
1170: final result reads as follows \cite{Belovhomo}: \e
1171: \alpha(\omega)=\frac{A\omega^2}{\omega_0^2-\omega^2+j\omega\Gamma},
1172: \l{alpha} \f where $A$ is amplitude, $\omega_0$ is resonant
1173: frequency and $\Gamma=A\omega k^3/(6\pi\mu_0)$ is the radiation
1174: reaction factor. The expressions for amplitude $A$ and resonant
1175: frequency $\omega_0$ in terms of dimensions of split-ring-resonators
1176: are available in \cite{SimSRR,Belovhomo}. In the present paper we
1177: will use typical parameters  $A=0.1\mu_0a^3$ and
1178: $\omega_0=1/(a\sqrt{\varepsilon_0\mu_0})$.
1179: 
1180: The dispersion properties of cubic lattice of split-ring-resonators
1181: with such parameters have been extensively studied in
1182: \cite{Belovhomo}. Using the theory of the present paper we will
1183: study reflection properties of such a metamaterial. Let us consider
1184: the case of cubic lattice ($a=b=c$), normal incidence ($k_y=k_z=0$)
1185: and let the magnetic field of incident wave is along directions of
1186: magnetic dipoles $\vec d=\vec y_0$. Numerical solution of dispersion
1187: equation \r{disphomom} with $q_y=k_y=0$ and $q_z=k_z=0$ allows to
1188: get a set of wave vectors of excited eigenmodes $\{q^x_i\}$. These
1189: wave vectors are plotted at the top of Fig. \ref{reflSRR} as
1190: functions of normalized frequency $ka$. The point $ka=1$ corresponds
1191: to the resonant frequency $\omega_0$ os split-ring-resonators. One
1192: can see that the propagating modes (${\rm Im}(q_x)=0$) exist only
1193: for $ka\le 0.978$ and $ka\ge 1.044$. It means that a partial
1194: resonant band gap is observed for $ka\in [0.978,1.044]$. At the
1195: frequencies inside of the band gap all the eigenmodes decay with
1196: distance. Note, that such decaying modes exist at all frequencies,
1197: not only inside of the band gap. The Fig. \ref{reflSRR} shows only
1198: the eigenmodes with slowest decay $|{\rm Im}(q_x)|<1.5\pi/a$. There
1199: is an infinite number of other decaying modes which decay with
1200: distance more rapidly. The contribution of such modes into the
1201: reflection coefficient is negligible as it was shown above. The
1202: decaying modes can be separated into the three classes:
1203: \begin{itemize}
1204: \item evanescent modes, the modes which have ${\rm Re}(q_x)=0$, they
1205: decay exponentially from one crystal plane to the other one;
1206: 
1207: \item staggered modes,  the modes which have ${\rm Re}(q_x)=\pi/a$,
1208: they exponentially decay from one crystal plane to the other one by
1209: absolute value, but the dipoles in the neighboring plane are excited
1210: in out of phase;
1211: 
1212: \item complex modes, the most general case of the decaying modes which have ${\rm Re}(q_x)\ne
1213: 0$, they experience both exponential decay and phase variation from
1214: one crystal plane to the other one.
1215: \end{itemize}
1216: 
1217: \begin{figure}[h]
1218: \centering \epsfig{file=dref.eps, width=8.5cm} \caption{(Color
1219: online) Dependencies of the normalized wave vectors $q_xa/\pi$ of
1220: excited eigenmodes (imaginary and real parts) and the reflection
1221: coefficient $R$ calculated using \r{Rmain} on normalized frequency
1222: $ka$ for semi-infinite cubic lattice of split-ring-resonators with
1223: $A=0.1\mu_0a^3$ and $\omega_0=1/(a\sqrt{\varepsilon_0\mu_0})$.}
1224: \label{reflSRR}
1225: \end{figure}
1226: 
1227: The evanescent modes are the most common type of decaying modes.
1228: They can be observed in dielectrics with negative permittivity, for
1229: example. The staggered modes are limiting case of the complex modes
1230: and can be widely observed in periodical structures in vicinity of
1231: the band gap edges, see for example \cite{Belovnonres,lwPRE}. The
1232: complex modes of general kind are quite exotic for common materials.
1233: 
1234: In the system under consideration we are able to observe all three
1235: kind of mentioned decaying modes. The presence of staggered and
1236: complex modes are evidence of spatial dispersion in this material
1237: reported in \cite{Belovhomo}. The staggered modes exists for $ka\ge
1238: 0.984$, evanescent modes for $ka\ge 1.015$ and complex modes for
1239: $ka\in [0.984,1.015]$ (see Fig. \ref{reflSRR}). One can see that for
1240: a fixed frequency from the range $ka\in [0.978,0.984]$  there are
1241: two staggered modes and in the range $ka\in [1.015,1.044]$ there are
1242: two evanescent modes. Actually, in the range $ka\in [0.984,1.015]$
1243: there are also two complex modes which have the same imaginary parts
1244: but differs by sign of the real part. Thus, we conclude that for
1245: every frequency from the range $ka\in [0.95,1.08]$ that we consider
1246: the incident wave will excite in the crystal a pair of modes with
1247: $|{\rm Im}(q_x)|<1.5\pi/a$ (propagating and staggered, two
1248: staggered, two complex, two evanescent or propagating and
1249: evanescent). Using the usual approach one has to introduce an
1250: additional boundary condition to solve this problem since usual
1251: condition of tangential component continuity is not enough in the
1252: case of excitation of two modes.
1253: 
1254: Using the theory introduced in the previous section it is enough to
1255: substitute obtained wave vectors of eigenmodes into \r{Rmain} the
1256: reflection coefficient is calculated. The reflection coefficient is
1257: plotted at the bottom of Fig. \ref{reflSRR}. One can see that at the
1258: frequencies in vicinity of the bottom and top edges of the band gap
1259: the semi-infinite crystal operate nearly as the electric and
1260: magnetic walls, respectively, as it was predicted in
1261: \cite{Belovnonres}. At the frequency $ka=0.984$ the reflection
1262: coefficient is equal to $+1$ (electric wall), and at $ka=1.044$ it
1263: is $-0.8+0.6j$ (nearly magnetic wall). Note, that the frequency
1264: corresponding to the electric wall effect is not equal to the bottom
1265: edge of the band gap and there are no frequency exactly
1266: corresponding to the magnetic wall effect. The use of the usual
1267: formulae for reflection coefficient from magnetic and
1268: Clausius-Mossotti formulae which do not take into account effects of
1269: spatial dispersion one could get idea that magnetic and magnetic
1270: wall effects have to happen at the edges of the band gap. Our study
1271: demonstrate that if the spatial dispersion is taken into account
1272: accurately then it is not so.
1273: 
1274: Thus, we have demonstrated how the proposed theory can be used for
1275: modeling of reflection from semi-infinite crystals with spatial
1276: dispersion. Our theory can be treated as generalization of results
1277: of Mahan and Obermair \cite{Mahan} which have been widely applied
1278: for modeling of various kinds of reflection problems. We hope that
1279: the present generalization can find much more applications in
1280: modeling of reflection  from spatially dispersive materials since in
1281: has no restriction on the period of the lattice to be smaller than
1282: wavelength and allows to consider electromagnetic crsytals of
1283: general kind.
1284: 
1285: \section{Conclusion}
1286: 
1287: In this paper a new approach for solving problems of plane-wave
1288: diffraction on semi-infinite electromagnetic crystals is proposed.
1289: The boundary conditions for the interface between isotropic
1290: dielectric and electromagnetic crystal of general kind are deduced
1291: in the form of infinite system of equations relating amplitudes of
1292: incident wave, excited eigenmodes and scattered spatial harmonics.
1293: This system of equations represent mathematical content of
1294: generalized Ewald-Oseen extinction principle which is formulated in
1295: this paper: the polarization of the semi-crystal excited by plane
1296: wave is distributed in such way that it cancels the incident wave
1297: together with all high-order spatial harmonics associated with
1298: periodicity of the boundary. In our opinion, the proof of
1299: generalized Ewald-Oseen extinction principle presented in this paper
1300: is an important theoretical fact which helps to understand
1301: interrelation between reflection and dispersion properties of
1302: electromagnetic crystals. If the eigenmodes of the infinite crystal
1303: are known then the system can be solved numerically which provides
1304: new numerical method for solving the diffraction problem under
1305: consideration. We believe in the quite good prospects for the
1306: application of the described method in further studies of dielectric
1307: and even metallic electromagnetic crystals at both microwave and
1308: optical ranges.
1309: 
1310: For the special case when the crystal is formed by small scatterers
1311: which can be effectively replaced by dipoles with fixed orientation
1312: the deduced system of equations is solved analytically using method
1313: of the characteristic function. The closed form expressions for the
1314: amplitudes of excited eigenmodes and scattered spatial harmonics are
1315: provided in terms of rapidly convergent products. These expressions
1316: can be treated as generalization of classical result of Mahan and
1317: Obermair \cite{Mahan} for the case when period of the lattice can be
1318: large as compared to the wavelength. The proposed method is applied
1319: for calculation of reflection coefficient from semi-infinite crystal
1320: formed by resonant magnetic scatterers (split-ring-resonators) at
1321: the frequencies corresponding to the strong spatial dispersion.
1322: 
1323: \section*{Acknowledgements}
1324: The authors would like to thank Sergei Tretyakov, Stanislav
1325: Maslovski and  Mario Silveirinha for useful discussions of the
1326: material presented in this paper.
1327: 
1328: 
1329: \bibliography{abc}
1330: \end{document}
1331: