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{On homogenization of electromagnetic crystals formed by uniaxial resonant scatterers}
27:
28: \author{Pavel A. Belov}
29: \affiliation{Mobile Communication Division, Telecommunication Network Business, Samsung Electronics Co., Ltd., \\
30: 94-1, Imsoo-Dong, Gumi-City, Gyeong-Buk, 730-350, Korea}
31:
32: %\email{belov@rain.ifmo.ru}
33:
34: \author{Constantin R. Simovski}
35: \affiliation{Photonics and Optoinformatics Department, St. Petersburg State University of Information Technologies, Mechanics and Optics, Sablinskaya 14, 197101, St. Petersburg, Russia}
36:
37: \date{\today}
38: \begin{abstract}
39: Dispersion properties of electromagnetic crystals formed by small
40: uniaxial resonant scatterers (magnetic or electric) are studied
41: using the local field approach. The goal of the study is to
42: determine the conditions under which the homogenization of such
43: crystals can be made. Therefore the consideration is limited by the frequency region
44: where the wavelength in the host medium is larger than the lattice
45: periods. It is demonstrated that together with known restriction for the homogenization
46: related with the large values of the material parameters there is
47: an additional restriction related with their small absolute values.
48: From the other hand, the homogenization becomes allowed in both cases of large and small material parameters
49: for special directions of propagation. Two unusual effects inherent to the crystals under consideration are revealed:
50: flat isofrequency contour which allows subwavelength imaging using canalization regime
51: and birefringence of extraordinary modes which can be used for beam splitting.
52: \end{abstract}
53:
54: \pacs{78.20.Ci, 42.70.Qs, 42.25.Gy}
55:
56: \maketitle
57:
58: \section{Introduction and problem formulation}
59: The problem of homogenization of bulk arrays of small scatterers
60: operating in the applied field as dipoles (elelectric or magnetic)
61: has a long history. One can recall here the classical works of
62: Lorentz, Madelung, Ewald and Oseen. However, in what concerns the
63: homogenization of arrays of small resonant scatterers these
64: classical results were revised in 1970-s taking into account the
65: possible shortening the propagating wave at the resonance and the
66: strong mutual coupling of resonant particles. It was done in the
67: seminal work by Sipe and Kranendonk \cite{Sipe}. In 1990-s the
68: interest to this problem was renewed by extensive studies of
69: bianisotropic composites (see e.g. in \cite{T1} and \cite{T2}).
70: The metal bianisotropic particles (chiral particles and omega
71: particles) have small resonant size at microwaves due to their complex
72: shape (they include a wire ring and straight wire portions).
73: However, the known studies of their homogenization are mainly
74: referred to the non-regular arrays. This can be explained by
75: specific applications of microwave bianisotropic composites (as
76: absorbers or antiradar coverings). The works like \cite{T3}
77: concerning the regular bianisotropic lattices do not consider
78: effects of particles resonance. Briefly, the homogenization
79: problem for resonant scatterers has not been studied enough.
80: However, it is becoming very important now due to the following
81: reasons.
82:
83: The first one is the rapid development of nano-technologies. It
84: becomes possible to prepare the lattices of metallic
85: nano-particles operating at the frequencies rather close to that
86: of the plasmon resonance of the individual particle. Recently, a
87: significant amount of works has been devoted to 1D arrays (chains)
88: of silver or gold particles which were found prospective for
89: subwavelength guiding the light (see e.g. \cite{Weber} and the
90: list of references of this work). It is evident that the 2D and 3D
91: lattices of metal nano-particles provide potentially even more
92: broad scope of optical applications than the chains. If the
93: homogenization of a 3D lattice is possible then one can use the basic
94: knowledge on the continuous media and apply it to the lattices.
95: This approach can be rather instructive and we demonstrate below
96: its example. In the present paper we study the case of microwave
97: scatterers, but this is only an illustration of the theory.
98: Similar results can be obtained for the optical range, too.
99: The electric scatterers of small resonant size are already known in
100: optics, and the possibility to create the small resonant
101: scatterers with magnetic properties was recently shown in work
102: \cite{T4}.
103:
104: The second motivation of the present research is related with the
105: intensive studies of the so-called left-handed media
106: \cite{Veselago}. The left-handed medium (LHM) is an effective
107: continuous medium with simultaneously negative permittivity and
108: permeability. The all-angle negative refraction and backward
109: waves are inherent to such media. The interest to these artificial
110: materials was evoked by seminal work of Pendry \cite{Pendrylens}
111: indicating the opportunity of the subwavelength imaging using a
112: slab of LHM. The most loud realization of LHM is a uniaxial
113: version of this medium studied in works
114: \cite{SmithWSRR},\cite{Shelbyscience},\cite{T5} and others. This
115: structure is composed from two components playing the roles of
116: the building blocks. The first block (responsible for negative
117: permittivity) is a wire medium
118: \cite{Brown,Rotmanps,Pendryw} and the second block (responsible
119: for negative permeability) is a lattice of the
120: split-ring resonators (SRR:s) \cite{PendrySRR}. The SRR:s are
121: small magnetic scatterers experiencing a two-time derivative
122: Lorentz-type resonance. As a result, the permeability of the
123: structure can take negative values within the resonance band of
124: SRR:s. This structure operates, however, as a LHM only in the
125: plane orthogonal to wires (and for an only polarization of the
126: wave). The reason of that is the strong spatial dispersion inherent to
127: wire media at all frequencies \cite{WMPRB}. This effect makes
128: the axial component of the wire medium permittivity
129: depending on the propagation direction. Only for the waves
130: propagating in the orthogonal plane this permittivity component is definitely
131: negative until the so-called plasma frequency and the structure
132: suggested in \cite{SmithWSRR} can be treated as a LHM only in this
133: special case.
134:
135: In order to obtain a variant of LHM operating in three dimensions some
136: attempts to use small resonant electric scatterers together with
137: magnetic ones \cite{LWD} as well as the bianisotropic scatterers \cite{T6,T7,T8} were made.
138: The samples of LHM obtained in \cite{T6,T8} demonstrate high losses in the LHM
139: regime and this makes the known variants of isotropic LHM not very interesting.
140:
141: However, if the goal is to observe negative refraction and
142: backward waves, and to obtain subwavelength images in three
143: dimensions, then the isotropic LHM is not an only solution. These
144: effects can be obtained in anisotropic structures, too. And not
145: only at high frequencies. The so-called indefinite media (in which
146: the principal components of permittivity and permeability tensors
147: have different signs) were studied in works
148: \cite{Smithindef,Smithindefref,Smithindeffoc}. These media offer
149: variety of effects including negative refraction, backward wave
150: effect, near field focusing, high-impedance surface reflection,
151: etc. Anisotropy of the media introduces additional freedom in
152: manipulation by its dispersion and reflection properties
153: \cite{Lind}. Even a uniaxial media with negative permittivity
154: along its axis allows to observe effects of negative refraction
155: and backward wave with respect to the interface \cite{BelovMOTL}.
156: The theoretical results \cite{Smithindef,Smithindefref,Smithindeffoc} do not prove that
157: the structure composed by wire medium and SRR:s will demonstrate
158: these effects in practice. On the contrary, from \cite{WMPRB} and
159: \cite{BelovMOTL} it is evident that these effects (which should
160: exist in a continuous uniaxial medium with negative axial
161: permittivity) are absent in wire media. In the same time, a
162: lattice of uniaxial electric scatterers oriented in parallel
163: allows to obtain the negative axial permittivity for all
164: directions of propagation (i.e. without spatial dispersion). If
165: such a lattice substitutes the wire medium in the structure
166: reported in \cite{SmithWSRR,Shelbyscience} then the effects
167: predicted in \cite{Smithindef,Smithindefref,Smithindeffoc} for
168: continuous indefinite media can be obtained in practice. This is
169: the second reason of the present study.
170:
171: In the current paper dispersion properties of electromagnetic crystals formed by orthorhombic lattices
172: of uniaxial magnetic or electric scatterers are studied.
173: The orientation of scatterers along one of the crystal axes is considered.
174: Geometry of the structure is presented in Fig. \ref{geomsm}.
175: As an example of magnetic scatterers we have chosen the SRR:s \cite{PendrySRR,SmithWSRR,Shelbyscience} (see Fig. \ref{SRR}).
176: The electric dipoles are represented by the short inductively loaded wires (ILW) \cite{LWD} (see Fig. \ref{lwd}).
177: \begin{figure}[h]
178: \centering
179: \epsfig{file=infpc.eps, width=7cm}
180: \caption{Geometry of an electromagnetic crystal. The arrows show directions of dipole moments of the uniaxial scatterers.}
181: \label{geomsm}
182: \end{figure}
183:
184: An analytical model based on dipole approximation and local field approach is introduced.
185: The dipole approximation (magnetic dipoles describing SRR:s and
186: electric dipoles describing ILW:s) restricts the dimensions of
187: inclusions to be much smaller than wavelength in the host media.
188: The local field approach allows to take into account the dipolar
189: interactions between scatterers exactly. It makes possible
190: accurate studies of lattice resonances. The results allows to
191: examine when the structure corresponds to its homogenized model of
192: local uniaxial media and when not.
193:
194: It is well known that the lattices of resonant scatterers (though
195: they do not exhibit the spatial dispersion at all frequencies unlike
196: wire media) can experience spatial dispersion at low frequencies
197: as compared to the spatial resonance of the lattice. This is the
198: case when the wavelength in the medium becomes comparable with
199: lattice period \cite{Sipe}. This results on resonance stop-band
200: \cite{Sipe} and on appearance of complex modes within it
201: \cite{Belovnonres,T7}. This makes the homogenization
202: impossible within a sub-band belonging to the resonance band. In
203: the present work we do not pay attention to the complex modes. The
204: comparison of the original lattice and its homogenized model is
205: made using the technique of isofrequency contours.
206: Such an approach allows to check correspondence between
207: properties of the structures under consideration and their
208: homogenization models.
209:
210: Uniaxial media with negative permittivity (or permeability) along
211: its axis and positive permittivity (or permeability) in the
212: transversal plane has the isofrequency contours of hyperbolic form
213: \cite{Smithindef,Smithindefref,Smithindeffoc,Lind}. These
214: isofrequencies correspond to the negative refraction
215: \cite{Smithindefref,Smithindeffoc}. If both original lattice and
216: its homogenized model possess such isofrequencies then they both
217: possess the negative refraction. More generally, if the
218: homogenized model of the lattice keeps (at least approximately)
219: the same isofrequency contours, then the homogenization is allowed. If
220: the homogenization dramatically change them the homogenization is
221: forbidden. This is the main idea of our approach.
222:
223:
224: \section{Models of individual scatterers}
225:
226: The geometries of the SRRs and ILW are presented in Fig. \ref{SRR}
227: and Fig. \ref{lwd}, respectively. Since the dipoles moments of all
228: scatterers are directed along $x$ (see Fig. \ref{geomsm}) an
229: individual scatterer can be characterized by scalar polarizability
230: $\alpha$ relating the dipole moment with the local field (external
231: field applied to a scatterer).
232:
233: \subsection{Split-Ring Resonators}
234: The SRR considered in \cite{PendrySRR,SmithWSRR,Shelbyscience} is a pair of two coplanar broken rings (see Fig.\ref{SRR}).
235: \begin{figure}[h]
236: \centering
237: \epsfig{file=SRR.eps, width=3.5cm}
238: \caption{Geometry of a Split-Ring-Resonator}
239: \label{SRR}
240: \end{figure}
241: Since the two loops are not identical the analytical models for
242: such SRR:s are rather cumbersome \cite{MarquesSRR,SimSRR}. In
243: fact, such SRR can not be described as a purely magnetic
244: scatterer, because it exhibits bianisotropic properties and has
245: resonant electric polarizability \cite{MarquesSRR,SimSRR} (see
246: also discussion in \cite{APSWSRR}). However, the electric
247: polarizability and bianisotropy of SRR is out of the scope of this
248: paper. We neglect these effects and consider an ordinary SRR as a
249: magnetic scatterer. The analytical expressions for the magnetic
250: polarizability $\alpha(\omega)$ of SRRs with geometry plotted in
251: Fig.\ref{SRR} were derived and validated in \cite{SimSRR}. The
252: final result reads as follows: \e
253: \alpha(\omega)=\frac{A\omega^2}{\omega_0^2-\omega^2+j\omega\Gamma},
254: \qquad A=\frac{\mu_0^2\pi^2r^4}{L+M}, \l{alpha} \f where
255: $\omega_0$ is the resonant frequency of magnetic polarizability:
256: $$
257: \omega_0^2=\frac{1}{(L+M)C_r},
258: $$
259: $L$ is inductance of the ring (we assume that both rings have the same inductance):
260: $$
261: L=\mu_0 r\left[\log\left(\frac{32R}{w}\right)-2\right],
262: $$
263: $M$ is mutual inductance of the two rings:
264: $$
265: M=\mu_0 r\left[(1-\xi)\log\left(\frac{4}{\xi}\right)-2+\xi\right], \qquad \xi=\frac{w+d}{2r},
266: $$
267: $C_r$ is the effective capacitance of the SRR:
268: $$
269: C_r=\varepsilon_0 \frac{r}{\pi} {\rm arccosh}
270: \left(\frac{2w}{d}\right),
271: $$
272: $\Gamma$ is the radiation reaction factor:
273: $$
274: \Gamma=\frac{A\omega k^3}{6\pi\mu_0},
275: $$
276: $r$ is the inner radius of the inner ring, $w$ is the width of the
277: rings, $d$ is distance between the edges of the rings (see
278: Fig.\ref{SRR}), $\varepsilon_0$ and $\mu_0$ are permittivity and
279: permeability of the host media, and
280: $k=\omega\sqrt{\varepsilon_0\mu_0}$ is the wave number of the host
281: medium. The presented formulae are valid within the frame of the
282: following approximations: $w,d\ll r$ and the splits of the rings
283: are large enough compared to $d$. Also, we assume that SRR is
284: formed by ideally conducting rings (no dissipation losses).
285:
286: The magnetic polarizability \r{alpha} takes into account the
287: radiation losses and satisfies to the basic Sipe-Kranendonk
288: condition \cite{Sipe,Belovcond,Belovnonres} which in the present
289: case has the following form: \e {\rm
290: Im}\left(\alpha^{-1}(\omega)\right)=\frac{k^3}{6\pi\mu_0}.
291: \l{sipe} \f
292:
293: In the following analysis we operate with the inverse polarizability
294: $\alpha^{-1}(\omega)$, thus, we rewrite \r{alpha} in the following
295: form: \e
296: \alpha^{-1}(\omega)=A^{-1}\left(\frac{\omega_0^2}{\omega^2}-1\right)+j\frac{k^3}{6\pi\mu_0}.
297: \l{invalph} \f
298:
299: \subsection{Inductively Loaded Short Wires}
300:
301: A typical resonant electric scatterer is an inductively loaded short wire, as
302: shown in Fig. \ref{lwd}.
303: \begin{figure}[h]
304: \centering \epsfig{file=lwd.eps, width=7.5cm} \caption{Geometry of
305: the inductively loaded wire dipole} \label{lwd}
306: \end{figure}
307:
308: The electric polarizability $\alpha_e$ of loaded wires following
309: the known model \cite{LWD} has the form:
310: \e
311: \alpha_e^{-1}=
312: \frac{3}{l^2 C_{\rm wire}}
313: \left(\frac{1-\omega^2/\omega_0^2}{4-\omega^2/\omega_0^2}\right)+
314: j\frac{k^3}{6\pi\varepsilon_0} \l{alpe} \f where
315: $C_{\mbox{wire}}=\pi l\varepsilon_0/\log(2l/r_0)$ is the
316: capacitance of the wire, $\omega_0=\sqrt{L C_{\rm wire}}$ is the
317: resonant frequency, $L$ is the inductance of the load, $l$ is the
318: half length of the wire and $r_0$ is the wire radius.
319:
320: It is clear, that at the frequencies near the resonance the polarizability of LSW has the form
321: \e \alpha_e^{-1}(\omega)\approx
322: A_e^{-1}\left(\frac{\omega_0^2}{\omega^2}-1\right)+j\frac{k^3}{6\pi\varepsilon_0},
323: \l{invalphe} \f with $A_e=l^2 C_{\rm wire}$, which is similar to
324: \r{invalph}. Moreover, if $A_e/\varepsilon_0=A/\mu_0$ then using
325: duality principle the magnetic dipole with polarizability $\alpha$
326: \r{invalph} can be transformed to the electric dipole with
327: polarizability $\alpha_e$ \r{invalph}, and vice versa. This means
328: that it is enough to consider only one type of resonant
329: scatterers. In the present paper we have chosen magnetic ones to
330: be principal. The electric scatterers can be easily obtained using
331: duality principle from the magnetic scatterers with
332: $A=\mu_0A_e/\varepsilon_0$.
333:
334: \section{Homogeneous media approach}
335:
336: Let us consider an orthorhombic lattice with periods $a\times
337: b\times c$ formed by magnetic uniaxial scatterers directed along
338: $x$ (see Fig. \ref{geomsm}) and described by polarizability
339: \r{alpha}. For electric scatterers (ILW:s) the problem is dual to
340: the present one. In the long wavelength limit the lattices of
341: scatterers are usually described as homogeneous media with certain
342: material parameters. The lattice under study can be modelled as a
343: resonant uniaxial magnetic. The permeability of such a magnetic is
344: a dyadic of the form:
345: $$
346: \=\mu=\mu \-x_0\-x_0+\mu_0(\-y_0\-y_0+\-z_0\-z_0).
347: $$
348:
349: The permeability $\mu$ ($x$-component of the tensor) can be
350: calculated though the individual polarizability of a single
351: scatterer using the Clausius-Mossotti formula \cite{Collin}: \e
352: \mu=\mu_0\left(1+\frac{\alpha(\omega)/(\mu_0V)}{1-C_s(a,b,c)
353: \alpha (\omega)/\mu_0}\right), \l{CM} \f where $V=abc$ is a volume
354: of the lattice elementary cell and $C_s(a,b,c)$ is the static
355: interaction constant of the lattice. The following expression for
356: this interaction constant is available in \cite{Collin}, p.758:
357: $$
358: C_s(a,b,c)=\frac{1}{4\pi}\sum\limits_{(m,n,l)\ne (0,0,0)}\frac{2(am)^2-(bn)^2-(cl)^2}{\left[(am)^2+(bn)^2+(cl)^2\right]^{5/2}}
359: $$
360: \e =\frac{1.202}{\pi a^3}-\frac{4\pi}{a^3}\sum\limits_{(n,l)\ne
361: (0,0)} \sum\limits_{m=1}^{+\infty} m^2 K_0 \left(\frac{2\pi
362: m}{a}\sqrt{(bn)^2+(cl)^2}\right)_, \l{cs} \f where $K_0(x)$ is the
363: modified Bessel function of 3d kind (the McDonald function). In
364: the case of a cubical lattice $a=b=c$ the interaction constant is
365: equal to the classical value $C_s=1/(3V)$.
366:
367: Notice, that the radiation losses contribution in expression
368: \r{invalph} should be skipped while substituting into formula
369: \r{CM}. This makes permeability purely real number as it should be
370: for lossless regular arrays \cite{Sipe,Belovnonres}. This
371: manipulation is based on the fact that the far-field radiation of
372: the single scatterer is compensated by the electromagnetic
373: interaction in a regular three-dimensional array, so that there is
374: no radiation losses for the wave propagating in the lattice. The
375: mathematical proof of this fact for the general dimensions of
376: lattice is presented in the Appendix.
377:
378: The typical dependence of magnetic permeability $\mu$ on frequency
379: is presented in Fig. \ref{eps} for cubic lattice ($a=b=c$) of SRR:s with parameters chosen so that $A=0.1\mu_0a^3$ and $\omega_0=1/(a\sqrt{\varepsilon_0\mu_0})$.
380: The resonant frequency shift from $ka=1$ to $ka=0.984$ is clearly observed.
381: While $ka<0.984$ the structure is paramagnetic ($\mu>1$). For $ka$ within $[0.984,1.0352]$ range the permeability is negative
382: ($\mu<0$). For $ka>1.0352$ the medium is diamagnetic ($0<\mu<1$).
383:
384: \begin{figure}[h]
385: \centering \epsfig{file=eps.eps, width=8.5cm}
386: \caption{Dependencies of relative permeability $\mu/\mu_0$ and normalized polarizability $\alpha/(\mu_0 a^3)$
387: vs. normalized frequency $ka$ for cubic lattice ($a=b=c$) of SRR:s with $A=0.1\mu_0a^3$ and $\omega_0=1/(a\sqrt{\varepsilon_0\mu_0})$.}
388: \label{eps}
389: \end{figure}
390:
391: The dispersion equation for the uniaxial magnetic medium has the
392: following form (see e.g. \cite{Collin,Lind,BelovMOTL}): \e \mu_0
393: (q_y^2+q_z^2)=\mu (k^2-q_x^2). \l{dispunis} \f Thus, the
394: isofrequency surfaces for such material have form of a spheroid if
395: $\mu>0$ (the spheroid is prolate for $\mu<1$ and oblate for
396: $\mu>1$) or a hyperboloid if $\mu<0$. Both types of isofrequency
397: surfaces have symmetry axis along $OX$. The media is isotropic in
398: the $YZ$ plane, and we can restrict our consideration by the $XY$
399: plane without loss of generality. The typical isofrequency
400: contours in this plane are shown in Figs. \ref{homisodown} and
401: \ref{homisoup}. The magnetic under consideration has the same
402: parameters as in Fig. \ref{eps}. The ranges of wave vector
403: components $q_x$ and $q_y$ are restricted by $\pm \pi/a$ and $\pm
404: \pi/b$, respectively, having in mind that the exact dispersion
405: diagram of the lattice corresponds to the first Brillouin zone and
406: we will compare the homogenized model with the exact theory.
407:
408: \begin{figure}[h]
409: \centering \epsfig{file=homisodown.eps, width=8.5cm}
410: \caption{Isofrequency contours in $XY$ plane for uniaxial magnetic with permeability as in Fig. \ref{eps}
411: for frequencies near the resonance of the permeability.}
412: \label{homisodown}
413: \end{figure}
414:
415: \begin{figure}[h]
416: \centering \epsfig{file=homisoup.eps, width=8.5cm}
417: \caption{Isofrequency contours in $XY$ plane for uniaxial magnetic with permeability as in Fig. \ref{eps}
418: for the frequencies where the permeability is close to zero.}
419: \label{homisoup}
420: \end{figure}
421:
422: While the frequency is below the resonance ($ka<0.984$) the
423: isofrequency contour has the form of an ellipse prolate along $OY$
424: ($\mu>1$). For frequencies above the resonance but less than the
425: frequency at which the permeability turns to zero
426: ($0.984<ka<1.0352$) the isofrequency contours are hyperbolas
427: ($\mu<0$), see Fig. \ref{eps}). If the frequency is above the
428: frequency at which the permeability passes zero ($ka>1.0352$) then
429: the isofrequency contour becomes an ellipse oblate along $OY$
430: ($0<\mu<1$). All the isofrequency contours pass through points
431: $q_x=\pm k$. In particular, all the ellipses have the same
432: semi-axes along $x$ (equal to $k$). Notice, that the solution
433: $q_x=k$ corresponds also to the arbitrary values of $q_y,q_z$ if
434: $\mu\rightarrow\infty$. This implies the propagation of all waves
435: along the optical axis $x$ with same phase velocity which is equal
436: to that of host medium. Strictly speaking, for an infinite value
437: as well as for finite large values of $\mu$ the homogenization is
438: forbidden. But we will show using the local field method that the
439: homogenization is allowed even at the frequency so that
440: $\mu\rightarrow \infty$ if one restricts by a special case of
441: propagation. In general, it turns out that at all frequencies
442: there are special cases of propagation for which the
443: homogenization is allowed!
444:
445: The hyperbolic form of isofrequency contour is a unique feature
446: inherent to resonant uniaxial magnetics
447: \cite{Smithindef,Smithindefref,Smithindeffoc,Lind,BelovMOTL}. It
448: allows to achieve negative refraction at all incident angles for
449: p-polarization in the case if the interface is normal to the
450: optical axis. In the case of resonant uniaxial dielectric the same
451: effect happens for s-polarization. If the uniaxial medium is
452: two-component and has both negative axial permittivity and
453: permeability, the negative refraction should be observed for both
454: p- and s- polarizations
455: \cite{Smithindef,Smithindefref,Smithindeffoc,Lind,BelovMOTL}.
456:
457: Below we will compare Figs. \ref{homisodown} and \ref{homisoup}
458: with those calculated for an original lattice of SRR:s using exact approach.
459:
460: \section{Dispersion equation for electromagnetic crystals formed by uniaxial scatterers}
461:
462: Following the local field approach the dipole moment $M$ of a zero
463: numbered scatterer is determined by the magnetic field $\-H_{\rm
464: loc.}$ acting to this scatterer: $M=\alpha H^x_{\rm loc.}$, where
465: $H^x_{\rm loc.}=(\-H_{\rm loc.}\.\-x_0)$. This local field is a sum of
466: the magnetic fields $\-H_{m,n,l}$ produced at the coordinate
467: origin by all other scatterers with indexes $(m,n,l)\ne (0,0,0)$:
468: \e \-H_{\rm loc.}=\sum\limits_{(m,n,l)\ne(0,0,0)} \-H_{m,n,l}.
469: \l{sum} \f
470:
471: The magnetic field produced by a single scatterer with index $(m,n,l)$ is given by dyadic Green's function $\=G(\-R)$:
472: \e
473: \-H_{m,n,l}=\mu_0^{-1}\=G(\-R_{m,n,l})\-M_{m,n,l},
474: \l{Ggen}
475: \f
476: where
477: $$
478: \=G(\-R)=\left(k^2\=I+\nabla\nabla\right)\frac{e^{-jkR}}{4\pi R}.
479: $$
480:
481: We consider uniaxial scatterers oriented along $\-x_0$ direction,
482: so it is enough to use only $\-x_0\-x_0$ component of dyadic
483: Green's function. So, we replace \r{Ggen} by the scalar
484: expression: \e H^x_{m,n,l}=\mu_0^{-1}G(\-R_{m,n,l})M_{m,n,l},
485: \l{Green} \f where
486: $$
487: G(\-R)=\left(k^2+\frac{\partial^2}{\partial x^2}\right)\frac{e^{-jkR}}{4\pi R}
488: $$
489: $$
490: =\left[(1+jkR)\frac{2x^2-y^2-z^2}{R^4}+k^2\frac{y^2+z^2}{R^2}\right]\frac{e^{-jkR}}{4\pi R}.
491: $$
492:
493: To study eigenmodes of the system we introduce the phase
494: distribution of dipole moments determined by the unknown wave
495: vector $\-q=(q_x,q_y,q_z)^T$ as follows: \e
496: M_{m,n,l}=Me^{-j(q_xam+q_ybn+q_zcl)}. \l{distr} \f
497:
498: Collecting together expressions \r{sum}, \r{Green} and \r{distr}
499: we obtain dispersion equation relating the wave vector $\-q$ with
500: the frequency $\omega$:
501: $$
502: M=\alpha\mu_0^{-1}\sum\limits_{(m,n,l)\ne(0,0,0)} G(\-R_{m,n,l}) Me^{-j(q_xam+q_ybn+q_zcl)}.
503: $$
504:
505: It can be rewritten in a more appropriate form: \e
506: \left[\mu_0\alpha^{-1}(\omega)-C(k,\-q)\right]M=0, \l{disp} \f
507: where \e C(k,\-q,a,b,c)=\sum\limits_{(m,n,l)\ne(0,0,0)}
508: G(\-R_{m,n,l}) e^{-j(q_xam+q_ybn+q_zcl)}. \l{C} \f
509:
510: We call $C$ as the dynamic interaction constant of the lattice
511: using the analogy with the classical interaction constant from the
512: theory of artificial dielectrics and magnetics \cite{Collin}.
513:
514: Dispersion equation \r{disp} has two different types of solutions.
515: The first ones are ordinary waves with zero dipole moments
516: ($M=0$). They are plain waves propagating in the host media which
517: have zero component of magnetic field along direction of dipoles.
518: They do not interact with lattice (do not excite magneto-dipole
519: moments). The waves of second type are extraordinary waves. They
520: excite magneto-dipole moments ($M \ne 0$) strongly interacting
521: with each other. The dispersion equation for extraordinary modes
522: transforms from \r{disp} to \e
523: \mu_0\alpha^{-1}(\omega)-C(k,\-q,a,b,c)=0. \l{disper} \f
524:
525: The solution of this dispersion equation allows to study
526: dispersion diagrams for the crystal under consideration. The main
527: problem is the calculation of the dynamic interaction constant $C$
528: given by \r{C}. This question is closely related with such
529: concepts as static interaction constant \r{cs} and the
530: triply-periodic dyadic Green's function. The static interaction
531: constant can be obtained from \r{C} by letting $k=q_x=q_y=q_z=0$
532: and choosing appropriate order of summation for obtained
533: conditionally convergent series \cite{Nijboer}. The plane-wise
534: summation method \cite{Nijboer,deWette} or Poisson summation
535: formulae based technique \cite{Collin,Kharadly} are usually
536: applied for calculation of the static interaction constant. The triply-periodic
537: dyadic Green's function represents the field produced by a phased
538: lattice of point dipoles. If the zero-numbered term is added to
539: series \r{C} (simultaneously one should move the observation point
540: in \r{Ggen} from the node of the lattice to avoid singularity)
541: formula \r{C} will give a co-polarized component of the dyadic
542: Green's function. The triply-periodic dyadic Green's function is
543: usually evaluated with help of classical Ewald's method
544: \cite{Ewald,Marioaccel,Mario3d}. However, the other methods of
545: summation (with improved convergence rate) exist as well
546: \cite{Nicorovichi3d,Borji}. All the listed above methods can be
547: applied for evaluation of dynamic interaction constant \r{C}. The
548: Ewald's method require an appropriate choice of a splitting
549: parameter \cite{splitpar}, which is a sophisticated manipulation.
550: Also, it does not show the energy balance in the lattice (as well
551: as other methods mentioned above). Therefore, we have chosen an
552: original method of summation. Our approach combines the plane-wise
553: summation \cite{Nijboer,deWette} and the Poisson summation
554: technique with singularity cancellation \cite{Collin}. The details
555: of the evaluation of $C$ which includes the energy balance
556: condition as an intermediate step are presented in Appendix.
557:
558: \section{Dispersion properties of the crystal}
559: The dispersion equation \r{disper} with interaction constant
560: $C(k,\vec q)$ given by formula \r{cfinal} from Appendix is solved
561: numerically. The parameters of the structure are the same as
562: those of the homogenized structure: cubic lattice ($a=b=c$) of
563: SRR:s, $A=0.1\mu_0a^3$ and
564: $\omega_0=1/(a\sqrt{\varepsilon_0\mu_0})$. The dispersion diagram
565: for the crystal is presented in Fig. \ref{dc}. The points
566: $\Gamma=(0,0,0)^T$, X $=(\pi/a,0,0)^T$, Y $=(0,\pi/a,0)^T$, L
567: $=(0,\pi/a,\pi/a)^T$, K $=(\pi/a,0,\pi/a)^T$ and R
568: $=(\pi/a,\pi/a,\pi/a)^T$ of the first Brilloun zone are
569: illustrated at the sketch in the left bottom corner of the plot.
570: The dotted lines represent dispersion curves for ordinary modes of
571: the crystal which coincide with light lines. The incomplete
572: resonant stop band for extraordinary modes (similar to discussed
573: in \cite{Belovnonres}) is observed in vicinity of the resonance of
574: individual inclusions.
575:
576: \begin{figure}[h]
577: \centering \epsfig{file=dc.eps, width=8.5cm}
578: \caption{Dispersion diagram for cubic lattice ($a=b=c$) of SRR:s with $A=0.1\mu_0a^3$ and $\omega_0=1/(a\sqrt{\varepsilon_0\mu_0})$.}
579: \label{dc}
580: \end{figure}
581:
582:
583:
584: \begin{figure}[h]
585: \centering \epsfig{file=dispcomp.eps, width=8.5cm}
586: \caption{Dispersion curve for $(0 1 0)$ direction. Exact solution (solid line), prediction by homogenization model
587: (dashed line) and light line (dotted line).}
588: \label{dispcomp}
589: \end{figure}
590:
591: The dispersion curve for $(0 1 0)$ direction (branch $\Gamma$Y in
592: Fig. \ref{dc}) is shown in Fig. \ref{dispcomp} for comparison with
593: result predicted by homogenization model. For this direction of
594: propagation the agreement with the homogenized model is fine
595: except narrow frequency range $ka\approx 0.98$. This region is not
596: visible in Fig. \ref{dispcomp} but it is clear from Fig. \ref{dc}
597: that this is the lower edge of the stop-band for waves propagating
598: in the transverse plane ($YZ$). This frequency range in the
599: homogenization model corresponds to high propagation constant
600: $q_y>\pi/a$ and high positive permeability $\mu>\pi^2/(ka)^2$.
601: This means that the homogenization in the case $\mu>\pi^2/(ka)^2$,
602: strictly speaking, describes the dispersion of the lattice in a
603: wrong manner. This is an expected result which corresponds to the
604: known predictions of the classical theory \cite{Sipe}. Below we
605: will consider this frequency range in details.
606:
607: The frequency band $0.9803<ka<1.044$ corresponds to the negative
608: axial permittivity of the homogenized model of the lattice.
609: Negative axial permittivity means the imaginary propagation
610: constant for the transverse plane and this result nicely
611: corresponds to the stop-band for the $YZ$ plane predicted by the
612: exact theory. So, the homogenization within $0.9803<ka<1.044$ is
613: allowed.
614:
615:
616: \begin{figure}[h]
617: \centering \epsfig{file=isoyz.eps, width=7.5cm}
618: \caption{Isofrequency contours in $YZ$ plane for frequencies near of the bottom edge of stop band.}
619: \label{isoyz}
620: \end{figure}
621:
622: The isofrequency contours in $YZ$ plane for frequencies near the
623: bottom $ka=0.96\dots 0.9803$ and top $ka=1.04\dots 1.10$ edges of
624: the transverse stop band are presented in Fig. \ref{isoyz}. The
625: behavior of isofrequency contours shown in Fig. \ref{isoyz} is
626: typical for general electromagnetic crystals at the frequencies
627: near the stop band edges \cite{NotomiPRB,Allanglediag,canal}.
628: While the frequency is rather far below the stop band ($ka<0.979$)
629: the isofrequency contours have form of circles and the agreement
630: with the homogenized model is fine. The same behavior is observed
631: above the stop band ($ka>1.044$). The circles for $ka>1.044$ are
632: smaller than those for $ka<0.979$ which nicely corresponds to the
633: smaller effective permeability (see Fig. \ref{eps}). However,
634: within the narrow frequency range $0.979<ka<0.9803$ the
635: isofrequency contours acquire a form which is different from a
636: circle. This anisotropy in the transverse plane gives the evidence
637: of spatial dispersion. Notice, that in this band in the lattice
638: there are two evanescent modes whose wave vectors lie in the
639: transverse plane (see also \cite{Belovnonres}). Strictly speaking,
640: the crystal can not be homogenized at these frequencies. And
641: these frequencies correspond to high positive $\mu$ of the
642: homogenized lattice. It was already noticed above that it is the
643: expected result.
644:
645:
646: \begin{figure}[h]
647: \centering \epsfig{file=isoup.eps, width=8.5cm}
648: \caption{Isofrequency contours in $XY$ plane for frequencies near the top edge of stop band.}
649: \label{isoup}
650: \end{figure}
651:
652: Significant disagreement between the exact solution and the result
653: of homogenization was also obtained at the frequencies near the
654: top edge of the stop band. The isofrequency contours in $XY$ plane
655: for this frequency range are presented in Fig. \ref{isoup}. They
656: dramatically differs from prediction given by homogenization model
657: shown in Fig. \ref{homisoup}. Following to homogenization
658: approach, the isofrequency contours should have hyperbolic form at
659: the frequencies corresponding to negative effective permittivity
660: and elliptic one in the case of positive permittivity (see Fig.
661: \ref{homisoup}). The exact modeling reveals that this switching
662: between hyperbolic and elliptic types of isofreqency contours
663: happens in the different manner. While $ka<1.0435$ the
664: isofrequency contours has a form which is similar to hyperbolic
665: one but they are already distorted. At the higher frequencies
666: $ka>1.0435$ the `hyperbolic' contours continue to distort, and
667: simultaneously the `elliptic' contours (the second branch of the
668: same isofrequency) appear in vicinity of $\Gamma$ point. The
669: `hyperbolic' contours pass through points $q_x=\pm k$ while
670: $ka<1.0455$, but 'elliptic' ones do not. For $ka>1.046$ the
671: situation changes to the opposite one. The `elliptic' contours
672: acquire fixed size along $OX$ axis and starts to pass through
673: points $q_x=\pm k$. On the other hand, the `hyperbolic' contours
674: start to collapse around X-point and completely disappear for
675: $ka>1.051$. This way the `hyperbolic' contours transform to
676: 'elliptic' ones passing through the regime where both types of
677: contours co-exist at the same frequencies. At any frequency only
678: one of these contours passes through points $q_x=\pm k$.
679:
680: Thus, in the region $1.043<ka<1.051$ the homogenization gives
681: wrong results for the waves propagating in the $XY$ plane because
682: of the two-mode regime observed in original structure. This region
683: corresponds to the small absolute values of $\mu$ ($|\mu|<0.2$ in
684: our case). Strictly speaking, the homogenization in the region of
685: small $|\mu|$ turns out to be forbidden. In our opinion, this is a
686: qualitatively new result. However, as it follows from Fig.
687: \ref{isoyz} the homogenized model makes correct prediction for the
688: waves propagating in the $YZ$ plane in the band $1.043<ka<1.051$.
689: One can conclude that the homogenization at these frequencies
690: (forbidden in its strict meaning) is however allowed for a case of
691: transversal propagation.
692:
693: The described regime of the co-existence of `hyperbolic' and
694: `elliptic' isofrequency contours at a fixed frequency means the
695: bi-refringence for extraordinary modes and three-refringence in
696: the case of the refraction (one ordinary wave and two
697: extraordinary ones). An extraordinary mode corresponding to the
698: `hyperbolic' contour refracts negatively, and the other one
699: (corresponding to the `elliptic' contour) experiences positive
700: refraction. This property can find different applications (beam
701: splitting, etc).
702:
703:
704: \section{Canalization regime and subwavelength imaging}
705:
706: Above, we pointed out that near the bottom edge of the stop-band
707: (frequencies corresponding to high positive $\mu$) the homogenized
708: model wrongly predicts the dispersion of the waves propagating in
709: the $YZ$ plane. Now let us show that the homogenized model gives
710: the qualitatively correct predictions in this frequency region if
711: consideration is restricted by the propagation in the $XY$ plane.
712: The isofrequency contours in $XY$ plane for frequencies near
713: bottom edge of stop band are presented in Fig.
714: \ref{isodown}. The behavior of contours is in the good agreement
715: with the predictions of the homogenized model (see Fig.
716: \ref{homisodown}). The difference is noticed only near the edges
717: of the lowest Brillouin zone. So, the homogenization (forbidden in
718: its strict meaning for $ka\approx0.98$) is still allowed for a
719: case of oblique propagation with respect to the optical axis.
720:
721: \begin{figure}[h]
722: \centering \epsfig{file=isodown.eps, width=8.5cm}
723: \caption{Isofrequency contours in $XY$ plane for frequencies near the bottom edge of stop band.} \label{isodown}
724: \end{figure}
725:
726: At the frequencies near $ka=0.989$ the isofrequency contours are
727: practically flat. It means that all eigenmodes at such frequencies
728: have the same axial component $q_x=\pm k$ of the wave vector.
729: Moreover, they all have the same group velocity (the group
730: velocity is normal to the isofrequency contour). This makes the
731: eigenmode to be the so-called transmission line mode like that of
732: the wire medium \cite{WMPRB}. For both lattice of uniaxial
733: scatteres and wire medium this isofrequency corresponds to the
734: infinite material parameter of the homogenized model. The
735: difference is that in the present case the flat isofrequency
736: contour exists at a single frequency $ka=0.989$ in contrast to
737: wire medium which support transmission line modes in a very wide
738: frequency range.
739:
740: The flat isofrequency contour we found can be used for the
741: implementation of the so-called canalization regime described in
742: our recent paper \cite{canal}. The similar regimes are called also
743: as self-guiding \cite{Chigrin}, directed diffraction \cite{Chien},
744: self-collimation \cite{Li} and tunneling \cite{Kuo}. In
745: \cite{canal} we have shown that not only all plane waves are
746: collimated into a strictly parallel beam at the flat isofrequency.
747: All evanescent waves impinging the medium at this frequency will
748: be also transformed into the plane wave with $q_x=k$ transporting
749: the energy along the optical axis. Therefore this regime allows to
750: create subwavelength images of the sources and transmit their near field
751: to unrestricted distances.
752:
753: The canalization regime for a slab of the medium possessing the
754: flat isofrequency does not involve negative refraction and
755: amplification of evanescent modes which are usually used for that
756: purpose \cite{Pendrylens,Allanglediag,Subwavelength}. Its main
757: feature is transformation of the spatial spectrum of the incident
758: field into a collimated beam directed across the slab. All spatial
759: harmonics of the source refract into such the eigenmodes at the
760: front interface. These eigenmodes all propagate normally to the
761: interface with same velocity and deliver the input distribution of
762: electric field to the back interface. Their refraction at the back
763: interface forms the image. The problem of the reflection from a
764: slab (and inner reflections in the slab) can be solved using the
765: Fabry-P{\'e}rot resonance. The Fabry-P{\'e}rot resonance holds for
766: all incidence angles including the complex angles. The reason of
767: that is simple: after the refraction all the incident waves
768: acquire the same longitudinal component of the wave vector
769: $q_x=k$. Thus, in the canalization regime there is no image
770: deterioration by the finite thickness of the lens (there are no
771: waves travelling along the interfaces).
772:
773:
774: \section{Conclusion}
775:
776: In the present paper we have studied dispersion properties of the
777: electromagnetic crystals formed by uniaxial resonant scatterers
778: (magnetic and electric ones). The structures are modelled using
779: the local field approach. The main tricky point of this theory is
780: evaluating the dynamic interaction constant of the lattice. This
781: constant has been calculated using a special analytical method
782: based on the plane-wise summation approach, Poisson summation
783: formula, singularity cancellation technique and convergence
784: acceleration of slowly convergent series. As a result, a
785: transcendental dispersion equation has been obtained in the form
786: suitable for rapid and efficient numerical calculations. The
787: comparison of exact solution provided by this equation with
788: homogenization model allows to show that the structure, strictly
789: speaking, can not be homogenized not only at the frequencies which correspond to the very
790: high values of effective permeability or permittivity (this was
791: well known earlier) but also at the frequencies corresponding to
792: small absolute values of them.
793:
794: However, if one is interested in a special cases of propagation then the
795: homogenization can be allowed in both these frequency bands. For
796: the propagation in the plane comprising the optical axis the
797: homogenization is allowed in the region of large material
798: parameters. For the propagation in the plane orthogonal to the
799: optical axis the homogenization is allowed in the region of small
800: material parameters.
801:
802: During this study we have found two interesting properties of the
803: crystals under consideration. At a single frequency near the
804: bottom edge of stop band the isofrequency contour is flat and this
805: frequency corresponds to the infinite permeability or
806: permittivity. This fact makes possible to use the crystals for
807: subwavelength imaging. The two-mode regime is observed at frequencies near
808: the top edge of stop band. This corresponds to the
809: bi-refringence for extraordinary waves and to three-refringence of
810: the incident wave in the general case of arbitrary polarization,
811: which can be used for beam splitting.
812:
813: The dispersion theory presented in this paper is a powerful tool
814: for dispersion analysis of three-dimensional electromagnetic
815: crystals. In the present form the theory is restricted to the case
816: of simple (uniaxial) scatterers, but it can be extended to the
817: case of electric or magnetic scatterers with arbitrary dyadic
818: response. This will be done in our future publications.
819: In this case it will be possible to develop an
820: analytical theory for a lattice of the isotropic resonant
821: scatterers (e.g. metallic spheres in the optical range) in more
822: accurate manner than the known low-frequency approximations
823: \cite{Kharadly,Collin,Ewald,smit} allow to do. This can be actual
824: for the optics of metal nano-particles which is developing fast.
825:
826:
827: \bibliography{EC}% Produces the bibliography via BibTeX.
828:
829: \appendix
830: \section{Evaluation of dynamic interaction constant}
831:
832: For calculation of the dynamic interaction constant $C(k,\-q)$ \r{C} we
833: apply a method based on plane-wise summation, Poisson summation
834: formulae and singularity cancellation technique. This method was
835: applied in \cite{WMJEWA} for calculation of the two-dimensional
836: dynamic interaction constant for theory of doubly-periodic wire lattices.
837: The series in \r{C} are divergent in classical meaning, but the
838: physical reasoning of necessary type of summation is clear enough.
839: Due to existence of losses in real space one should add
840: infinitesimal imaginary part to the wave vector $k$ of free space
841: and tend it to zero in order to get the correct result.
842:
843: We split series \r{C} (remember that the zero term is excluded from summation) onto
844: three parts:
845: $$
846: \sum\limits_{l\ne 0}\sum\limits_{m=-\infty}^{+\infty}\sum\limits_{n=-\infty}^{+\infty}+
847: \left.{\sum\limits_{n\ne 0}\sum\limits_{m=-\infty}^{+\infty}}\right|_{l=0}+
848: \left.{\sum\limits_{m\ne 0}}\right|_{l=n=0.}
849: $$
850:
851: \begin{figure}[h]
852: \centering
853: \epsfig{file=areas.eps, width=8.5cm}
854: \caption{Splitting areas}
855: \label{split}
856: \end{figure}
857:
858: These parts are denoted as $C_{1,2,3}$ respectively, and
859: $C=C_1+C_2+C_3$. The splitting areas are shown in Fig.\ref{split}.
860: The term $C_{1}$ describes contribution into the local field from
861: all plane grids which are parallel to the $OXY$ plane except the
862: grid located at this plane. The term $C_{2}$ corresponds to the
863: contribution of the dipole linear chains parallel to the $OX$ axis
864: and located at $OXY$ plane except the chain located at this axis.
865: The term $C_{3}$ is the contribution from all dipoles of the chain
866: located at the $OX$ axis except the dipole located at the origin
867: of coordinate system. Notice, that $C_2+C_3$ gives the interaction
868: constant of the planar grid.
869:
870: For evaluation of the term $C_{1}$ it is possible to use the Poisson summation
871: formula for double series which leads to the expression with rapidly (exponentially) convergent series.
872: The term $C_{2}$ can be calculated using the ordinary Poisson summation formula together with the singularity
873: cancellation technique \cite{Collin}. It is impossible to apply Poisson summation formula for
874: evaluation of the term $C_{3}$ since it contain non-complete series.
875: Convergence of these series can be accelerated using the dominant part extraction \cite{Kantorovich}.
876:
877: \subsection{Contribution of parallel planar grids}
878:
879: The double Fourier transformation of any function $f(x,y)$ is
880: defined as follows:
881: $$
882: F(p,q)={\bf L}_{x,y}\left\{f(x,y)\right\}=\int\limits_{-\infty}^{+\infty}\int\limits_{-\infty}^{+\infty} f(x,y)e^{-j(px+qy)} dx dy.
883: $$
884:
885: The Poisson summation formula for double series has the following form:
886: \cite{Collin}: \e \sum\limits_{m=-\infty}^{+\infty}
887: \sum\limits_{n=-\infty}^{+\infty} f(am,bn)=
888: \frac{1}{ab}\sum\limits_{m=-\infty}^
889: {+\infty}\sum\limits_{n=-\infty}^{+\infty} F\left(\frac{2\pi
890: m}{a},\frac{2\pi n}{b}\right)_. \l{pdouble} \f
891:
892: The double Fourier transform of the Hertz potential of a dipole
893: reads as: \e {\bf L}_{x,y}\left\{\frac{1}{4\pi}\frac{ e^{
894: -jk\sqrt{x^2+y^2+z^2}}}{\sqrt{x^2+y^2+z^2}}\right\}=
895: \frac{1}{2}\frac{e^{-|z|\sqrt{p^2+q^2-k^2}}}{ \sqrt{p^2+q^2-k^2}},
896: \l{tr} \f where the sign of square root should be chosen so that
897: ${\rm Im}\left(\sqrt{p^2+q^2-k^2}\right)\ge 0$.
898:
899: Applying the shift and differential properties of Fourier
900: transformation to the Fourier-image of the Hertz potential \r{tr}
901: we obtain the transformation rule:
902: $$
903: {\bf L}_{x,y}\left\{
904: \left[\left(k^2+\frac{\partial^2}{\partial^2 x}\right)
905: \frac{ e^{-jk\sqrt{x^2+y^2+z^2}}}{\sqrt{x^2+y^2+z^2}}
906: \right] \frac{e^{-j(q_xx+q_yy)}}{4\pi}
907: \right\}
908: $$
909: \e
910: =\frac{k^2-(q_x+p)^2}{2}
911: \frac{e^{-|z|\sqrt{(q_x+p)^2+(q_y+q)^2-k^2}}}{ \sqrt{(q_x+p)^2+(q_y+q)^2-k^2}},
912: \l{dtc}
913: \f
914:
915:
916: Using the Poisson summation formula \r{pdouble} together with
917: \r{dtc} we obtain the term $C_1$ in the form: \e
918: C_1=\sum\limits_{l\ne
919: 0}\sum\limits_{m=-\infty}^{+\infty}\sum\limits_{n=-\infty}^{+\infty}
920: \frac{jp_m^2}{2ab}
921: \frac{e^{-j\left(|cl|k_z^{(mn)}+q_zcl\right)}}{k_z^{(mn)}},
922: \l{floq} \f where
923: $$
924: k_x^{(m)}=q_x+\frac{2\pi m}{a},\qquad k_y^{(n)}=q_y+\frac{2\pi n}{b},
925: $$
926: $$
927: p_m=\sqrt{\left(k_x^{(m)}\right)^2-k^2},
928: $$
929: $$
930: k_z^{(mn)}=-j\sqrt{\left(k_x^{(m)}\right)^2+\left(k_y^{(n)}\right)^2-k^2}.
931: $$
932:
933: Here we choose ${\rm Im}(\sqrt{\.})\ge 0$, so that ${\rm
934: Im}\left(k_z^{(mn)}\right)\le 0$. The representation \r{floq} can
935: be treated as an expansion of the fields produced by parallel
936: dipole grids in terms of the Floquet modes. The wave vectors of these
937: modes are
938: $$
939: \-k^{(mn)}=\left(k_x^{(m)},k_y^{(n)},k_z^{(mn)}\right)^T.
940: $$
941:
942: The series by $l$ index in \r{floq} are geometrical progressions and their summation can be made directly:
943: $$
944: \sum\limits_{l\ne 0}e^{-j\left(|cl|k_z^{(mn)}+q_zcl\right)}
945: =-\frac{e^{-j k_z^{(mn)}c}-\cos q_zc}{\cos k_z^{(mn)}c-\cos q_zc}.
946: $$
947: It allows to rewrite expression \r{floq} for the term $C_1$ as \e
948: C_1=\sum\limits_{m=-\infty}^{+\infty}\sum\limits_{n=-\infty}^{+\infty}
949: \frac{p_m^2}{2jab k_z^{(mn)}} \frac{e^{-j k_z^{(mn)}c}-\cos
950: q_zc}{\cos k_z^{(mn)}c-\cos q_zc}. \l{c1} \f These series possess
951: exponential convergence. It is clearly seen if the second factor
952: of the term under sign of the sum in \r{c1} is rewritten as
953: $$
954: -\left[\frac{1}{e^{j(k_z^{(mn)}+q_z)cl}-1}+\frac{1}{e^{j( k_z^{(mn)}-q_z )cl}-1}\right].
955: $$
956: This makes \r{c1} suitable for rapid numerical calculations.
957:
958: \subsection{Contribution of parallel chains from OXY-plane}
959:
960: The ordinary Fourrier transformation has the form
961: $$
962: F(p)={\bf L}_x\left\{f(x)\right\}=
963: \int\limits_{-\infty}^{+\infty} f(x)e^{-jpx}dx.
964: $$
965:
966: Poisson's summation formula for single series reads:
967: \e
968: \sum\limits_{m=-\infty}^{+\infty} f(am)=
969: \frac{1}{a}\sum\limits_{m=-\infty}^{+\infty}
970: F\left(\frac{2\pi m}{a}\right).
971: \l{Poisson}
972: \f
973:
974: The Fourrier transform for the Hertz's potential of a dipole is following:
975: \e
976: {\bf L}_x\left\{\frac{1}{4\pi}\frac{ e^{-jk\sqrt{x^2+y^2+z^2}}}{\sqrt{x^2+y^2+z^2}}\right\}
977: =
978: \frac{1}{2\pi}K_0\left(\sqrt{p^2-k^2}\sqrt{y^2+z^2}\right)_.
979: \l{tr2}
980: \f
981:
982: Thus, applying shift and differential properties of Furrier
983: transformation to the image of Hertz's potential \r{tr2} we get
984: the following transformation rule:
985: $$
986: {\bf L}_x\left\{\frac{1}{4\pi}\left(k^2+\frac{\partial ^2}{\partial x^2}\right)\frac{ e^{ -jk\sqrt{x^2+y^2+z^2}}}{\sqrt{x^2+y^2+z^2}}\right\}
987: $$
988: \e
989: =
990: \frac{1}{2\pi}(k^2-p^2)K_0\left(\sqrt{p^2-k^2}\sqrt{y^2+z^2}\right)_.
991: \l{dtc2}
992: \f
993:
994: Using Poisson's summation formulae \r{Poisson} together with \r{dtc2} we obtain the term $C_2$ in the form:
995: \e
996: C_2=-\sum\limits_{n\ne 0}\sum\limits_{m=-\infty}^{+\infty}
997: \frac{p_m^2}{2\pi a}K_0\left(p_m|bn|\right)e^{-jq_ybn}.
998: \l{pc2}
999: \f
1000:
1001: If arguments of McDonald's functions in \r{pc2} have nonzero real
1002: part then the series by index $n$ have very good convergence, but
1003: if these are getting imaginary then McDonald's functions transform
1004: to Hankel's functions and the mentioned series become slowly
1005: convergent. Therefore we separate the part of \r{pc2} which has
1006: good convergence: \e
1007: C'_2=-\sum\limits_{n=1}^{+\infty}\sum\limits_{{\rm Re}(p_m)\ne 0}
1008: \frac{p_m^2}{\pi a}K_0\left(p_mbn\right)\cos (q_ybn). \l{c'} \f
1009: The residuary part of \r{pc2} ($C'_2=C'_2+C''_2$)
1010: \e C''_2=-\sum\limits_{n\ne
1011: 0}\sum\limits_{{\rm Re}(p_m)=0} \frac{p_m^2}{2\pi
1012: a}K_0\left(p_m|bn|\right)e^{-jq_ybn}, \l{c''} \f which has slow
1013: convergence should be calculated with help of special method. Note, that there is only finite
1014: number of indexes $m$ such that ${\rm Re}(p_m)=0$. It means that
1015: in \r{c''} the summation by index $m$ includes only finite number of terms.
1016: For example, at the low frequency limit, when period $a$ is large
1017: compared with wavelength in the host media, the equation ${\rm
1018: Re}(p_m)=0$ has only one solution $m=0$ if $q_x<k$.
1019:
1020: We will calculate the sum of the series \r{c''} as the limit with
1021: $z$ tending to zero: \e C''_2= \lim\limits_{z\to
1022: 0}\sum\limits_{n\ne 0}\sum\limits_{{\rm Re}(p_m)=0}
1023: \frac{-p_m^2}{2\pi
1024: a}K_0\left(p_m\sqrt{(bn)^2+z^2}\right)e^{-jq_ybn}. \l{clim} \f
1025: Introducing the auxiliary parameter $z$ makes possible to
1026: complement series \r{clim} by zeroth terms and then to use the
1027: Poisson summation formula by index $n$ (see \r{dtc} for necessary
1028: Fourier transform). The result is as follows: \e
1029: C''_2=\lim\limits_{z\to 0} \sum\limits_{{\rm Re}(p_m)=0}
1030: \frac{-p_m^2}{2ab} \left(\sum\limits_{n=-\infty}^{+\infty}
1031: \frac{e^{-j|z|k_z^{(mn)}}}{jk_z^{(mn)}}
1032: -\frac{b}{\pi}K_0\left(p_m|z|\right) \right)_. \l{hc2} \f
1033:
1034: The term $K_0\left(p_m|z|\right)$ in \r{hc2} plays role of the
1035: zero term which is subtracted from the complete series (already
1036: transformed using Poisson summation formula) in order to get
1037: series \r{clim} without zero term. This term contains singularity
1038: if $z$ tends to zero. This singularity disappears in \r{hc2}
1039: during substraction from the complete series which experience the same singularity.
1040: In order to cancel out these singularities analytically we apply the method of a dominant
1041: series. Namely, we split the series from \r{hc2} onto dominant and
1042: correction parts:
1043: $$
1044: \sum\limits_{n=-\infty}^{+\infty}
1045: \frac{e^{ -j|z|k_z^{(mn)}}}{jk_z^{(mn)}}=
1046: 2b \sum\limits_{n=1}^{+\infty}
1047: \frac{e^{ -2\pi |z|n/b}}{2\pi n}
1048: $$
1049: $$+
1050: \sum\limits_{n\ne 0}
1051: \left[
1052: \frac{e^{ -j|z|k_z^{(mn)}}}{jk_z^{(mn)}}-
1053: b\frac{e^{ -2\pi |z||n|/b}}{2\pi |n|}\right]
1054: + \frac{e^{ -j|z|k_z^{(m0)}}}{jk_z^{(m0)}}.
1055: $$
1056:
1057: The dominant series can be evaluated using the tabulated formula
1058: (see \cite{Collin}, Appendix): \e \sum\limits_{n=1}^{+\infty}
1059: \frac{e^{ -n\alpha}}{n}= -\log \left(1-e^{-\alpha}\right).
1060: \l{sumn} \f
1061:
1062: The whole singularity is included into the dominant series.
1063: The correction series have no singularity when $z$ tends to zero.
1064: Using this fact the formula \r{hc2} can be rewritten as:
1065: $$
1066: C''_2= \sum\limits_{{\rm Re}(p_m)=0}
1067: \frac{-p_m^2}{2ab}
1068: \left( \frac{1}{jk_z^{(m0)}}+ \sum\limits_{n\ne 0}
1069: \left[\frac{1}{jk_z^{(mn)}}-\frac{b}{2\pi |n|}\right]\right.
1070: $$
1071: \e
1072: \left.-\frac{b}{\pi}\lim\limits_{z\to 0}\left[\log \left(1-e^{- 2 \pi |z|/b}\right)+K_0\left(p_m|z|\right)\right]
1073: \vphantom{\sum\limits_{n\ne 0} \left[\frac{1}{jk_z^{(mn)}}-\frac{b}{2\pi |n|}\right]}\right)
1074: \l{limK}
1075: \f
1076:
1077: The logarithmic singularity occurring here is compensated
1078: by that arising from the term with the McDonald function.
1079: The small-argument expression for the McDonald function reads:
1080: $$
1081: K_0\left(\alpha \right)\to -\left[\gamma+\log(\alpha/2)\right],
1082: $$
1083: where $\gamma=0.577$ is Euler's constant.
1084:
1085: Thus, the value of the limit in \r{limK} is as follows:
1086: \e
1087: \lim\limits_{z\to 0}
1088: \left[\.\right]
1089: =
1090: -\left(\log \frac{b|p_m|}{4\pi}+\gamma+j\frac{\pi}{2}\right).
1091: \l{limfin}
1092: \f
1093:
1094:
1095: The series in \r{limK} with index $n\in [-\infty,+\infty]$ except
1096: $n=0$ have convergence $1/n^2$. Such convergence rate makes
1097: calculations not rapid enough. the convergence can be accelerated
1098: by extraction of the dominant series. In order to get the
1099: convergence $1/n^4$ it is enough to extract series of order
1100: $1/n^2$ and $1/n^3$: \e \sum\limits_{n=1}^{+\infty}
1101: \left[\frac{1}{jk_z^{(m,n)}}-\frac{b}{2\pi n}\right]=
1102: \sum\limits_{n=1}^{+\infty}
1103: \left[\frac{1}{jk_z^{(m,n)}}-\frac{b}{2\pi n} \right. \l{dominant}
1104: \f
1105: $$
1106: \left.
1107: +\frac{q_yb^2}{4\pi^2 n^2}-\frac{l_mb^3}{16\pi^3 n^3}\right]-
1108: \frac{q_yb^2}{24}+1.202 \frac{l_mb^3}{16\pi^3},
1109: $$
1110: where $l_m=2q_y^2-p_m^2$ and we have taken into account, that
1111: $$
1112: \sum\limits_{n=1}^{+\infty}\frac{1}{n^2}=\frac{\pi^2}{6},\qquad
1113: \sum\limits_{n=1}^{+\infty}\frac{1}{n^3}=1.202.
1114: $$
1115:
1116: Collecting the terms corresponding to $+n$ and $-n$ together
1117: in \r{limK} we obtain: \e \sum\limits_{n\ne 0}
1118: \left[\frac{1}{jk_z^{(mn)}}-\frac{b}{2\pi |n|}\right]=
1119: \sum\limits_{n=1}^{+\infty}
1120: \left[\frac{1}{jk_z^{(m,n)}}+\frac{1}{jk_z^{(m,-n)}}-\frac{b}{\pi
1121: n}\right] \l{transf} \f
1122: $$
1123: =\sum\limits_{n=1}^{+\infty}
1124: \left[\frac{1}{jk_z^{(m,n)}}+\frac{1}{jk_z^{(m,-n)}}-\frac{b}{\pi n}-\frac{l_mb^3}{8\pi^3 n^3}\right]+1.202 \frac{l_mb^3}{8\pi^3}.
1125: $$
1126:
1127: The property $k_z^{(m,-n)}(q_x,q_y)=k_z^{(m,n)}(q_x,-q_y)$ makes function
1128: $$
1129: \frac{1}{jk_z^{(m,n)}}+\frac{1}{jk_z^{(m,-n)}}=
1130: $$
1131: $$
1132: \left[\frac{\frac{b}{2\pi |n|}}{\sqrt{\left(\frac{q_yb}{2\pi n}+1\right)^2+\frac{p_m^2}{4\pi^2 n^2/b^2}}}
1133: +
1134: \frac{\frac{b}{2\pi |n|}}{\sqrt{\left(\frac{q_yb}{2\pi n}-1\right)^2+\frac{p_m^2}{4\pi^2 n^2/b^2}}}
1135: \right]
1136: $$
1137: even with respect to variable $q_yb/(2\pi n)$. It means that
1138: being expanded into Taylor series it will contain only even power
1139: terms. Thus, the transformed series \r{transf} being expanded as
1140: the series of the order $1/n$ will contain only odd-power terms.
1141: In \r{transf} we have already extracted the dominant series of the
1142: order $1/n^3$. So, we conclude that series \r{transf} have convergence of the order $1/n^5$
1143: which is better than it was estimated when we started to extract dominant series in \r{dominant}
1144:
1145: Collecting the parts of the term $C_2$ given by \r{c'}, \r{limK},
1146: \r{limfin} and \r{transf} we obtain the final formula for $C_2$
1147: possessing convergence which is appropriate for rapid and effective numerical calculations:
1148: $$
1149: C_2=-\sum\limits_{n=1}^{+\infty}\sum\limits_{{\rm Re}(p_m)\ne 0}
1150: \frac{p_m^2}{\pi a}K_0\left(p_mbn\right)cos(q_ybn),
1151: $$
1152: $$
1153: -\sum\limits_{{\rm Re}(p_m)=0}
1154: \frac{p_m^2}{2ab}
1155: \left(\frac{1}{jk_z^{(m0)}}+ \sum\limits_{n=1}^{+\infty}
1156: \left[\frac{1}{jk_z^{(m,n)}}+\frac{1}{jk_z^{(m,-n)}}\right.
1157: \right.
1158: $$
1159: \e
1160: \left.\left.
1161: -\frac{b}{\pi n}-\frac{l_mb^3}{8\pi^3 n^3}\right]+1.202 \frac{l_mb^3}{8\pi^3}+ \frac{b}{\pi} \left(\log \frac{b|p_m|}{4\pi}+\gamma\right)+j\frac{b}{2}
1162: \vphantom{\frac{1}{jk_z^{(m0)}}+ \sum\limits_{n=1}^{+\infty}
1163: \left[\frac{1}{jk_z^{(m,n)}}+\frac{1}{jk_z^{(m,-n)}}\right.}\right)
1164: \l{c2}
1165: \f
1166:
1167:
1168: \subsection{Contribution of the line located at OX-axis}
1169:
1170: The term $C_3$ has form of the series \cite{Tretlines,Weber}:
1171: \e
1172: C_3=\frac{1}{2\pi a^3} \sum\limits_{m\ne 0}
1173: \left(\frac{1}{|m|^3}+\frac{jka}{m^2}\right)
1174: e^{-j(k|am|+q_xam)}.\l{hc3}
1175: \f
1176:
1177: These series have convergence which is not enough for effective direct numerical calculations.
1178: We will use convergence acceleration technique presented in \cite{Kantorovich}
1179: in order to evaluate these series. The dominant series can be extracted in the following way:
1180: $$
1181: \sum\limits_{m=1}^{+\infty} \left(\frac{1}{m^3}+\frac{jka}{m^2}\right)
1182: e^{-jsm}
1183: =\sum\limits_{m=1}^{+\infty} \left(\frac{1}{m^3}+\frac{jka}{m^2}
1184: -\frac{jka}{m(m+1)}\right.
1185: $$
1186: $$
1187: \left. -\frac{jka+1}{m(m+1)(m+2)}\right) e^{-jsm}
1188: + jka \sum\limits_{m=1}^{+\infty} \frac{e^{-jsm}}{m(m+1)}
1189: $$
1190: \e
1191: +(jka+1) \sum\limits_{m=1}^{+\infty} \frac{e^{-jsm}}{m(m+1)(m+2)}
1192: \l{extract}
1193: \f
1194:
1195: The first series in the right-hand side of \r{extract} (that
1196: containing the expression in prances) can be simplified up to
1197: $$
1198: \sum\limits_{m=1}^{+\infty} \frac{(2jka+3)m+2}{m^3(m+1)(m+2)}e^{-jsm}.
1199: $$
1200: These series have convergence $1/m^4$ which is convenient for rapid
1201: calculations. The other series in the right-hand side of
1202: \r{extract} can be evaluated in the closed form using the formula \r{sumn}:
1203: $$
1204: \sum\limits_{m=1}^{+\infty} \frac{e^{-jsm}}{m(m+1)}=
1205: \sum\limits_{m=1}^{+\infty} \frac{e^{-jsm}}{m}-\sum\limits_{m=1}^{+\infty} \frac{e^{-jsm}}{m+1}
1206: $$
1207: $$
1208: =-(1-e^{js})\log\left(1-e^{-js}\right)+1,
1209: $$
1210: $$
1211: \sum\limits_{m=1}^{+\infty} \frac{e^{-jsm}}{m(m+1)(m+2)}
1212: $$
1213: $$
1214: =
1215: \frac{1}{2}\left[\sum\limits_{m=1}^{+\infty} \frac{e^{-jsm}}{m(m+1)}
1216: -\sum\limits_{m=1}^{+\infty} \frac{e^{-jsm}}{(m+1)(m+2)}\right]
1217: $$
1218: $$
1219: =-\frac{1}{2}\left[(1-e^{js})^2\log\left(1-e^{-js}\right)+e^{js}-\frac{1}{2}\right].
1220: $$
1221:
1222: After these manipulations the formula \r{extract} transforms as
1223: follows:
1224: $$
1225: C_3=\frac{1}{4\pi a^3} \left[
1226: 4\sum\limits_{m=1}^{+\infty} \frac{(2jka+3)m+2}{m^3(m+1)(m+2)}e^{-jkam}\cos(q_xam)
1227: \right.
1228: $$
1229: \e
1230: -(jka+1)\left(t_+^2\log t^++t_-^2\log t^-+2e^{jka}\cos (q_xa)\right)
1231: \l{c3}
1232: \f
1233: $$
1234: \left.
1235: -2jka\left(t_+\log t^++t_-\log t^-\right)
1236: +(7jka+3)
1237: \vphantom{\sum\limits_{m=1}^{+\infty} \frac{(2jka+3)m+2}{m^3(m+1)(m+2)}}
1238: \right],
1239: $$
1240: where
1241: $$
1242: t^+=1-e^{-j(k+q_x)a}, \qquad t^-=1-e^{-j(k-q_x)a},
1243: $$
1244: $$
1245: t_+=1-e^{j(k+q_x)a}, \qquad t_-=1-e^{j(k-q_x)a}.
1246: $$
1247:
1248: The expression \r{c3} looks more cumbersome as compared to the
1249: initial formula \r{hc3}, but it is much more convenient for rapid
1250: calculations. The estimations show that in order to get accuracy
1251: of $0.01\%$ one needs to take more than $200$ terms in expression
1252: \r{hc3} and only $10$ terms in \r{c3}.
1253:
1254: \subsection{Energy conservation}
1255:
1256: In this subsection we evaluate the imaginary part of $C$ and
1257: consider the problem of the energy balance in a 1D array (chain)
1258: of dipoles, in a 2D array (grid) and in a 3D array (lattice).
1259:
1260: Let us return to formula \r{hc3} and find the imaginary part of
1261: the interaction constant of the dipole chain:
1262: \e
1263: {\rm Im} (C_3)=\frac{2}{2\pi a^3} \sum\limits_{m=1}^{\infty} \cos
1264: q_xam \left(\frac{\sin kam}{m^3}-ka \frac{\cos kam}{m^2}\right).
1265: \l{cimc3}
1266: \f
1267: To calculate these series we used the auxiliary formulas \e
1268: \sum\limits_{m=1}^{+\infty} \frac{\cos
1269: sm}{m^2}=\frac{(\pi-s')^2}{4}-\frac{\pi^2}{12}, \l{aux1}\f
1270: \e \sum\limits_{m=1}^{+\infty} \frac{\sin sm}{m^3}=\frac{s'^3-3\pi
1271: s'^2+2\pi^2s'}{12}. \l{aux2}\f These formulas can be easily
1272: obtained from the relation \r{sumn} rewritten for the case
1273: $\alpha=js$
1274: $$
1275: \sum\limits_{m=1}^{+\infty} \frac{e^{-jsm}}{m}=
1276: -\log \left(1-e^{-js}\right)
1277: $$
1278: \e =-\left(\log \left|2\sin \frac{s}{2}\right|+j\frac{\pi
1279: -s'}{2}\right), \l{sumnj} \f where $s'=2\pi\{s/(2\pi)\}$ and we
1280: use notation $\{x\}$ for fractional part of variable $x$. To
1281: derive \r{aux1} and \r{aux2} one should integrate \r{sumnj} by
1282: parameter $s$.
1283:
1284: Note, that real part of $C_3$ contain series \e
1285: \sum\limits_{m=1}^{+\infty} \frac{\sin sm}{m^2}, \quad {\rm and}
1286: \quad \sum\limits_{m=1}^{+\infty} \frac{\cos sm}{m^3}, \l{series}
1287: \f which can be expressed in terms on second and third order repeating integrals of tangent function.
1288: These integrals can not be evaluated in elementary functions,
1289: but they are suitable for numerical calculations. We prefer to use
1290: the acceleration technique leading to the result \r{c3} for evaluation of $C_3$,
1291: rather than to use numerical integration.
1292: Some other recommendations for calculation of series \r{series} can be found in \cite{Collin} together with their
1293: expansions into Taylor series.
1294:
1295: After substitution of \r{aux1} and \r{aux2} into \r{cimc3} and some algebra
1296: the following compact form for ${\rm Im} (C_3)$ can be obtained:
1297: \e {\rm Im}
1298: (C_3)=\frac{k^3}{6\pi}+\frac{1}{4a}\sum\limits_{|k_x^{(m)}|<k}p_m^2.
1299: \l{imc3} \f
1300:
1301: It is easy to obtain the imaginary parts of $C_2$ and $C_1$.
1302: The imaginary part of formula \r{c2} reads as: \e {\rm Im} (C_2)=
1303: \sum\limits_{{\rm Im} (k_z^{(mn)})=0}\frac{p_m^2}{2ab
1304: k_z^{(mn)}}-\sum\limits_{|k_x^{(m)}|<k}\frac{p_m^2}{4a}. \l{imc2}
1305: \f The imaginary part of formulae \r{c1} reads: \e {\rm Im}
1306: (C_1)=-\sum\limits_{{\rm Im} (k_z^{(mn)})=0}\frac{p_m^2}{2ab
1307: k_z^{(mn)}}. \l{imc1} \f Collecting together \r{imc3}, \r{imc2}
1308: and \r{imc1} we obtain that \e {\rm Im}(C)=\frac{k^3}{6\pi}.
1309: \l{imc} \f This relation makes dispersion equation \r{disp} real
1310: valued for the case of propagating modes.
1311:
1312: Now, let us discuss the energy balance in the chain using the
1313: result \r{imc3}. If the dipoles are arranged in a periodical
1314: linear array $x=am$ phased by wave vector with the $x-$component
1315: $q_x$ (as in \cite{Tretlines}) then the structure radiates
1316: cylindrical waves. The number of these waves depends on the
1317: relation between the wavelength, chain period and phase constant
1318: $q_x$. In the regime of the guided mode $q_x>k$ this number is
1319: zero since $|k_x^{(m)}|>k$ for all $m$. Using the Sipe-Kranendonk
1320: condition \r{sipe} for the imaginary part of the polarizability's
1321: inverse value one can obtain a purely real valued dispersion equation for
1322: the guided mode in the chain \cite{Tretlines,Weber}:
1323: $$
1324: \mu_0\alpha^{-1}(\omega)-C_3(\omega,q_x,a)=0
1325: $$
1326:
1327: However, the arrangement of the dipoles into an array changes the
1328: radiation losses of the individual scatterers. The effective
1329: polarizability of the scatterer in the linear array becomes as
1330: follows:
1331: $$
1332: \alpha_1=\left(\alpha^{-1}-\mu_0^{-1}C_3\right)^{-1}.
1333: $$
1334: The Sipe-Kranendonk condition in the general case of radiated
1335: waves should be replaced by \e {\rm
1336: Im}\left(\alpha_1^{-1}\right)=\sum\limits_{|k_x^{(m)}|<k}\frac{-p_m^2}{4a\mu_0}.
1337: \l{sipeline} \f The expression \r{sipeline} follows from the
1338: formulae \r{sipe} and \r{imc3}. This relation expresses the
1339: balance between the radiation losses of the individual scatterer
1340: of the chain and the contribution of the chain unit cell into the
1341: radiated waves ($|k_x^{(m)}|<k$).
1342:
1343: Now, consider a 2D grid of dipoles located at the nodes with
1344: coordinates $x=am$ and $y=bn$ and phased by real $q_x$ and $q_y$,
1345: respectively. The effective polarizability of a scatterer in this
1346: planar grid is \e
1347: \alpha_2=\left(\alpha_1^{-1}-\mu_0^{-1}C_2\right)^{-1}=\left(\alpha^{-1}-\mu_0^{-1}(C_2+C_3)\right)^{-1}.
1348: \l{alpha2} \f
1349:
1350: The formulae \r{alpha2} and \r{imc2} allow to formulate an
1351: analogue of the Sipe-Kranendonk condition for the planar grid: \e
1352: {\rm Im}\left(\alpha_2^{-1}\right)=\sum\limits_{{\rm Im}
1353: (k_z^{(mn)})=0}\frac{-p_m^2}{2ab\mu_0 k_z^{(mn)}}. \l{sipegrid} \f
1354: The terms $-p_m^2/(4a\mu_0)$
1355: corresponding to the cylindrical waves in \r{sipeline} are cancelled
1356: out by the respective terms from \r{imc2} and replaced by terms
1357: $p_m^2/(2ab \mu_0 k_z^{(mn)})$. The last ones correspond to the
1358: radiated plane-waves (Floquet harmonics with indexes $(m,n)$
1359: produced by the grid). The condition ${\rm Im} (k_z^{(mn)})=0$ for
1360: the finite sum in \r{sipegrid} is the radiation condition for
1361: these Floquet harmonics. Formula \r{sipegrid} expresses the
1362: balance between the radiation losses of the dipole and the
1363: contribution of the grid unit cell into radiation.
1364:
1365: In the surface wave regime, when ${\rm Im} (k_z^{(mn)})\ne 0$ for all
1366: $m,n$, using the Sipe-Kranendonk condition \r{sipe} one can
1367: obtain a real valued dispersion equation for the surface wave propagating
1368: along the grid:
1369: $$
1370: \mu_0\alpha^{-1}(\omega)-\tilde C_2(\omega,q_x,q_y,a,b)=0,
1371: $$
1372: where
1373: $$
1374: \tilde C_2(\omega,q_x,q_y,a,b)=C_2(\omega,q_x,q_y,a,b)+C_3(\omega,q_x,a).
1375: $$
1376:
1377: Finally, let us consider a 3D lattice with nodes $x=am$, $y=bn$
1378: and $z=cl$ phased by real $q_x$, $q_y$ and $q_z$ respectively. The
1379: effective polarizability of the scatterer in this lattice is \e
1380: \alpha_3=\left(\alpha_2^{-1}-C_1\right)^{-1}=\left(\alpha^{-1}-C\right)^{-1}.
1381: \l{alpha3} \f
1382:
1383: From \r{imc1},\r{alpha3} and \r{sipegrid} we easily obtain \e {\rm
1384: Im}\left(\alpha_3^{-1}\right)=0. \l{sipelattice} \f The terms
1385: $p_m^2/(2ab\mu_0 k_z^{(mn)})$ in \r{sipegrid} are cancelled by
1386: respective terms of \r{imc1}. Physically, it means that radiation
1387: losses of the scatterer in this lattice are zero. The lattice does
1388: not radiate power because it fills the whole space and the
1389: radiation losses of the single scatterer are compensated by the
1390: electromagnetic interaction in the lattice (as well as in the
1391: waveguide regimes of the chain and of the grid).
1392:
1393:
1394: \subsection{Final formula}
1395:
1396: Collecting together results \r{c1}, \r{c2}, \r{hc3} we obtain the
1397: final expression for the dynamic interaction constant:
1398:
1399: $$
1400: C(k,\-q,a,b,c)=
1401: -\sum\limits_{n=1}^{+\infty}\sum\limits_{{\rm Re}(p_m)\ne 0}
1402: \frac{p_m^2}{\pi a}K_0\left(p_mbn\right)\cos(q_ybn)
1403: $$
1404: \e
1405: +\sum\limits_{m=-\infty}^{+\infty}\sum\limits_{n=-\infty}^{+\infty}
1406: \frac{p_m^2}{2jab k_z^{(mn)}}
1407: \frac{e^{-j k_z^{(mn)}c}-\cos q_zc}{\cos k_z^{(mn)}c-\cos q_zc}
1408: \l{cfinal}
1409: \f
1410: $$
1411: -\sum\limits_{{\rm Re}(p_m)=0}
1412: \frac{p_m^2}{2ab}
1413: \left(\frac{1}{jk_z^{(m0)}}+ \sum\limits_{n=1}^{+\infty}
1414: \left[\frac{1}{jk_z^{(m,n)}}+\frac{1}{jk_z^{(m,-n)}}\right.
1415: \right.
1416: $$
1417: $$
1418: \left.\left.
1419: -\frac{b}{\pi n}-\frac{l_mb^3}{8\pi^3 n^3}\right]+1.202 \frac{l_mb^3}{8\pi^3}+ \frac{b}{\pi} \left(\log \frac{b|p_m|}{4\pi}+\gamma\right)+j\frac{b}{2}
1420: \vphantom{\sum\limits_{n\ne 0} \left[\frac{1}{jk_z^{(mn)}}-\frac{b}{2\pi |n|}\right]}\right)
1421: $$
1422: $$
1423: +\frac{1}{4\pi a^3} \left[
1424: 4\sum\limits_{m=1}^{+\infty} \frac{(2jka+3)m+2}{m^3(m+1)(m+2)}e^{-jkam}\cos(q_xam)
1425: \right.
1426: $$
1427: $$
1428: -(jka+1)\left(t_+^2\log t^++t_-^2\log t^-+2e^{jka}\cos (q_xa)\right)
1429: $$
1430: $$
1431: \left.
1432: -2jka\left(t_+\log t^++t_-\log t^-\right)
1433: +(7jka+3)
1434: \vphantom{\sum\limits_{m=1}^{+\infty} \frac{(2jka+3)m+2}{m^3(m+1)(m+2)}}
1435: \right],
1436: $$
1437: where we use following notations (introduced above and collected
1438: here):
1439: $$
1440: k_x^{(m)}=q_x+\frac{2\pi m}{a},\qquad k_y^{(n)}=q_y+\frac{2\pi n}{b},
1441: $$
1442: $$
1443: p_m=\sqrt{\left(k_x^{(m)}\right)^2-k^2},\qquad
1444: l_m=2q_y^2-p_m^2,
1445: $$
1446: $$
1447: k_z^{(mn)}=-j\sqrt{\left(k_x^{(m)}\right)^2+\left(k_y^{(n)}\right)^2-k^2}.
1448: $$
1449: $$
1450: t^+=1-e^{-j(k+q_x)a}, \qquad t^-=1-e^{-j(k-q_x)a},
1451: $$
1452: $$
1453: t_+=1-e^{j(k+q_x)a}, \qquad t_-=1-e^{j(k-q_x)a}.
1454: $$
1455:
1456: The calculations using \r{cfinal} can be restricted to the
1457: real part only, because its imaginary part is predefined by \r{imc}.
1458: The series in \r{cfinal} have excellent
1459: convergence that ensure very rapid numerical calculations.
1460:
1461: \subsection{Low frequency limit case}
1462:
1463: It is useful to consider the low frequency limit (when $k$, $q_x$,
1464: $q_y$ and $q_z$ are small as compared with $1/a$, $1/b$ and $1/c$)
1465: and show that the result for $C$ transits to the known one for
1466: this case. Following to definition \r{hc3} for term $C_3$ we
1467: conclude, that
1468: $$
1469: C_3=\frac{1}{\pi a^3} \sum\limits_{m=1}^{+\infty} \frac{1}{m^3}=\frac{1.202}{\pi a^3}.
1470: $$
1471: The expression \r{c2} for $C_2$ reduces to
1472: $$
1473: C_2=-\frac{8\pi}{a^3}\sum\limits_{m=1}^{+\infty}\sum\limits_{n=1}^{+\infty} m^2 K_0\left(\frac{2\pi m}{a} bn\right).
1474: $$
1475: Note, that both $C_3$ and $C_2$ turn out to be independent on $k$ and $\-q$.
1476: The formula \r{c1} for $C_1$ splits into the two terms: the first one which depend on $k$ and $\-q$ (where we have expanded trigonometric functions into Taylor series) and some additional constant:
1477: $$
1478: C_1=-\frac{1}{abc} \frac{k^2-q_x^2}{k^2-q_x^2-q_y^2-q_z^2}
1479: -\frac{4\pi}{a^2b}\sum\limits_{m=1}^{+\infty}\frac{m}{e^{2\pi mc/a}-1}
1480: $$
1481: $$
1482: - \frac{8\pi}{a^3} \sum\limits_{m=1}^{+\infty}\sum\limits_{n=1}^{+\infty}
1483: \frac{m^2/\sqrt{(bm/a)^2+n^2}}{e^{2\pi\sqrt{(bm/a)^2+n^2}c/b}-1}
1484: $$
1485: Finally, we get
1486: \e
1487: C=-\frac{1}{abc} \frac{k^2-q_x^2}{k^2-q_x^2-q_y^2-q_z^2}+C_s(a,b,c),
1488: \l{clow}
1489: \f
1490: where $C_s(a,b,c)$ is static interaction constant \r{cs}, and obtain alternative representation for $C_s(a,b,c)$:
1491: $$
1492: C_s(a,b,c)=
1493: \frac{1.202}{\pi a^3}
1494: -\frac{8\pi}{a^3} \left[\sum\limits_{m=1}^{+\infty}\sum\limits_{n=1}^{+\infty}
1495: \frac{m^2/\sqrt{(bm/a)^2+n^2}}{e^{2\pi\sqrt{(bm/a)^2+n^2}c/b}-1}\right.
1496: $$
1497: \e \left. +\frac{a}{2b}\sum\limits_{m=1}^{+\infty}\frac{m}{e^{2\pi
1498: mc/a}-1} +\sum\limits_{m=1}^{+\infty}\sum\limits_{n=1}^{+\infty}
1499: m^2 K_0\left(\frac{2\pi m}{a} bn\right)\right]_. \l{csnew} \f
1500: The static interaction constant expressed as \r{csnew} is equivalent
1501: to \r{cs}. The expression \r{csnew} can be obtained from \r{cs} applying
1502: Poisson summation formula by index $n$ and then by direct
1503: summation by index $l$ (in the same manner as it was done above during evaluation of term $C_1$).
1504: Both formulae \r{csnew} and \r{cs} are extremely effective for rapid numerical
1505: calculations due to excellent convergence of the series.
1506: The difference between \r{cs} and \r{csnew} is that \r{cs} contains triple series in contrast to
1507: \r{csnew} which comprises only double ones.
1508: Noteworthy, that convergence of series in \r{cs} is higher than in \r{csnew}.
1509:
1510: Formula \r{clow} being substituted into \r{disp} reduces the
1511: dispersion equation for an electromagnetic crystal to the known
1512: dispersion equation of a continuous uniaxial magnetic \r{dispunis}
1513: with magnetic permittivity of the form \r{CM}.
1514: This fact is an important verification of introduced dispersion theory.
1515:
1516: \end{document}
1517: