Open access peer-reviewed chapter

Dyakonov Surface Waves: Anisotropy-Enabling Confinement on the Edge

Written By

Carlos J. Zapata-Rodríguez, Slobodan Vuković, Juan J. Miret, Mahin Naserpour and Milivoj R. Belić

Submitted: 10 May 2017 Reviewed: 17 January 2018 Published: 02 May 2018

DOI: 10.5772/intechopen.74126

From the Edited Volume

Surface Waves - New Trends and Developments

Edited by Farzad Ebrahimi

Chapter metrics overview

1,139 Chapter Downloads

View Full Metrics

Abstract

The title “Dyakonov surface waves: anisotropy enabling confinement on the edge” plainly sets the scope for this chapter. The focus here is on the formation of bounded waves at the interface of two distinct media, at least one of them exhibiting optical anisotropy, which are coined as Dyakonov surface waves (DSWs) in recognition to the physicist who reported their existence for the first time. First, the general aspects of the topic are discussed. It also treats the characterization of bounded waves in isotropic-uniaxial multilayered structures, allowing not only the derivation of the dispersion equation of DSWs but also that of surface plasmons polaritons (SPPs), for instance. Furthermore, the interaction of such surfaces waves, with the possibility of including guided waves in a given planar layer and external sources mimicking experimental setups, can be accounted for by using the transfer matrix formalism introduced here. Finally, special attention is devoted to hyperbolic media with indefinite anisotropy-enabling hybridized scenarios integrating the prototypical DSWs and SPPs.

Keywords

  • surface waves
  • anisotropy
  • transfer matrix formulation

1. Introduction

The planar interface of two dissimilar materials plays a relevant role in many optical phenomena. In recent years, particularly, evanescent waves have been used in newly developing technologies such as near-field spectroscopy. The electromagnetic surface wave, which is intimately tied to the interface, travels in a direction parallel to the interface but, on either side of the interface, its amplitude is imperceptible after a certain distance from the interface. The notion of an electromagnetic surface wave made a significant appearance in 1907 when Zen-neck [1] authored a theoretical paper exploring the possibility of a wave guided by the interface of the atmosphere and either Earth or a large body of water. His focus was on radio waves, a region of the electromagnetic spectrum far from the optical regime in which we are particularly interested, but the principles involved are the same, owing to the scale invariance of the Maxwell postulates.

Yet, nearly a century later, a unique type of wave, the surface plasmon polariton (SPP) wave, that dominates the nanotechnology scene, at least at optical frequencies, resulted in wonderful developments with the creation of extremely sensitive bio/chemical sensors, and improvements in this mature technology continue to this day [2]. Even in this highly developed application, the two partnering materials which meet at the interface may be simple: one is a typical metal, a plasmonic material at optical frequencies, and the other is a homogeneous, isotropic, dielectric material. While the interface of a plasmonic material and a polarizable material supports SPPs, a variety of other types of surface waves can be supported by the interface of two polarizable materials. Since polarizable materials such as dielectric materials are less dissipative, in general, than plasmonic materials such as metals, the advantage of these materials for long-range propagation of surface waves is apparent.

The interface of two homogeneous dielectric materials of which at least one is anisotropic may support surface-wave propagation of another type, even though the real parts of all components of the permittivity dyadics of both materials are positive. Interest in surface waves guided by the interface of two dielectric materials began to take after Dyakonov in 1988 [3] explored the propagation of a surface wave guided by the interface of a uniaxial dielectric material and an isotropic dielectric material. The Dyakonov surface wave (DSWs) is the focus of this chapter.

In this chapter, we perform a thorough analysis of DSWs taking place in semi-infinite anisotropic media. Basic concepts related to the propagation of electromagnetic waves in homogeneous media will be introduced, including isotropic and anisotropic materials. Birefringent metal-dielectric (MD) lattices will be also considered as a contribution of meta-materials in the development of DSWs [4]. Special emphasis will be put when the effective-medium approaches induce unsatisfactory results, which occur in most experimental configurations. Practical cases will be analyzed including dissipative effects due to Ohmic losses of the metal.

Advertisement

2. Wave propagation in bulk media and interfaces

In this section we introduce the basic concepts related to the propagation of electromagnetic waves in homogeneous media, including isotropic and anisotropic materials. We describe in detail complex multilayered structures. For that purpose, we introduce a transfer matrix formulation that applies to isotropic and uniaxial media simultaneously. Finally, we discuss the conditions that give rise to surface waves at the interface of two isotropic media; the case of dealing with anisotropic media is considered in Section 5. Moreover, we obtain the dispersion equation for SPPs, which appears at the interface between a dielectric and a metal.

2.1. Wave propagation in isotropic media

In this section, we consider the propagation of electromagnetic waves in linear, homogeneous, and isotropic dielectrics. Under these conditions, the relative permittivity ϵ relating E and D is a scalar constant. Considering that the medium is free of electric charges and currents, and taking into account the medium equation B=μ0H, Maxwell’s equations can be written as

×H=ϵϵ0E∂t,E1a
×E=μ0H∂t,E1b
·E=0,E1c
·H=0.E1d

Now, each of the scalar components of E and H satisfies the wave equation 2uc2t2u=0, where u represents any one of the six scalar components of the electromagnetic field and c is the speed of the waves in the medium.

When the electromagnetic wave is plane and monochromatic, all components of the electric and magnetic fields are harmonic functions in time and space at the time frequency ω and spatial frequency k=kxkykz, respectively. Particularly they may be set as

Ert=E0expik·rexpiωt,E2a
Hrt=H0expik·rexpiωt,E2b

where E0r and H0r are the complex amplitudes of the electric and magnetic fields. If we substitute the vector wave fields of Eqs. (2) in Maxwell’s equations (1), with the help of the transformations ik and t, we might attain a simplified expression of k×k×E0, enabling to obtain the following wave equation: M·E0=0, where Mkkk2I+k02ϵI. Here I is the 3 × 3 identity matrix, and k is the modulus of the wavevector k and k0=ω/c0. In order to obtain the dispersion equation, we look for nontrivial solutions of the electric field E0 by imposing that detM=0. Its solution leads to k2=ω2/c2. The electric field amplitude E0 can be written as a linear combination of the following vectors

ê1=0kzky,E3a
ê2=ky2+kz2kykxkzkx.E3b

We point out that ê1 is associated with TEx modes and that ê1.ê2=0. Although not demonstrated here, ê2 is related to TMx-polarized plane waves. Note also that Eqs. (1c) and (1d) lead to the following orthogonality relations-hips: k·e^1=k·e^2=0. As a result, the vectors e^1e^2k form an orthogonal trihedron.

2.2. Wave propagation in uniaxial media

Uniaxial crystals are media with certain symmetries that make them have two equal principal refractive indices: nx=nyno (ordinary index) and nzne (extraordinary index). The crystal is to be a positive uniaxial if ne>no and negative uniaxial if ne<no. The z axis of a uniaxial crystal is called the optic axis. We call it the ordinary polarization direction if the wave has an eigenindex of refraction no and extraordinary, if the wave has an eigenindex of refraction ne. In Table 1 we show the values of no and ne for some natural birefringent materials (uniaxial crystals).

Birefringent materialnoneΔn=neno
Crystal quartz1.5471.5560.009
MgF1.37861.39040.0118
YVO41.99292.21540.2225
Rutile (TiO2)2.652.950.3
E7 liquid crystal1.5201.7250.205
Calomel (Hg2Cl2)1.962.620.68

Table 1.

Birefringence Δn of some natural materials [5].

Let us now consider the propagation of wave planes in uniaxial media. To obtain the eigenvalues associated with plane-wave propagation, we proceed in a similar way as in Section 2.1 for isotropic media, but taking into account that now the relative permittivity ϵ=ϵxxx+yy+ϵzzz is a tensor. Therefore, we solve detM=0, where the matrix Mkkk2I+k02ϵ. Then, after solving the abovementioned determinant, we obtain two solutions. The first of them is kx2+ky2+kz2=ϵxk02 which corresponds to the dispersion equation of the ordinary plane waves. The electric field for this kind of plane waves is proportional to the vector êo=kykx0. As a consequence, ordinary plane waves are TEz-polarized waves. The second solution gives us the dispersion equation of the extraordinary plane waves:

kx2+ky2ϵz+kz2ϵx=k02.E4

In this case, the electric field is proportional to the vector e^e=kxkzkykzkz2ϵxk02.

2.3. Matrix formulation for multilayered media

In this section we look at the case of multilayered media composed of different nonmagnetic materials which are separated by planar parallel interfaces, displaced on x=xi as shown in Figure 1. In particular, we deal with either uniaxial or isotropic materials, including metals. We make a detailed description of the electromagnetic fields inside a given medium “i”, which lies in xi1<x<xi. Our objective is the analysis of the appropriate conditions for the propagation of surface waves at the abovementioned interfaces.

Figure 1.

Schematic arrangement of the multilayered media. The amplitudes Ai and Ai (Bi and Bi) correspond to waves propagating along the positive (negative) x-axis. These amplitudes characterize a given state of polarization: For uniaxial media we may deal with ordinary (for example, Aoi) and extraordinary (Aei) waves. For isotropic materials we have TE-polarized (ATEi) and TM-polarized (ATMi) waves.

2.3.1. Electromagnetic fields in a uniaxial elementary layer

Let us first consider a multilayered media of uniaxial materials. For simplicity, let us take into account only relative permittivities of the form ϵi=ϵxixx+yy+ϵzizz for a given medium “i.” As we demonstrated in Section 2.2, the dispersion equation for extraordinary waves propagating in bulk uniaxial media is given by Eq. (4). Due to the boundary conditions at the interfaces between different media, the components of the wave vector ky and kz are conserved, but not its projection upon the x-axis. More specifically, if we rename kxikoi for ordinary waves, we can write

koi=ϵxik02ky2+kz2,E5

and for extraordinary waves (here kxikei)

kei=ϵzik02ky2+ϵzikz2ϵxi,E6

taken from Eq. (4).

The total electric field of the elementary layer “i” is set as

Etoti=Eixexpikyy+ikzziωt.E7

The part of the electric field that varies along with the spatial coordinate x can be written as

Eix=Aoia^oiexpikoixxi+Boib^oiexpikoixxi+Aeia^eiexpikeixxi+Beib^eiexpikeixxi,E8

where the amplitude Aoi (and Aei) corresponds to propagating ordinary (and extraordinary) waves, and Boi (and Bei) is related with counter-propagating ordinary (and extraordinary) waves. Note that all these amplitudes have zero dephase at x=xi. Finally, the vectors a^oi and a^ei are rewritten as:

a^oi=kykoi0,E9a
a^ei=keikzkykzkz2ϵxik02.E9b

In the case of counter-propagating waves, we take into consideration that kxi=koi for ordinary waves and kxi=kei for extraordinary waves. This fact leads us to introduce the vector fields

b^oi=kykoi0,E10a
b^ei=keikzkykzkz2ϵxik02.E10b

The vector fields a^oi and b^oi can be given in units of k0, whereas a^ei and b^ei can be expressed in units of k02.

For convenience, the field function Ei can be set in terms of the wave amplitudes Aoi, Aei and Boi and Bei with zero dephase at x=xi1, namely:

Eix=Aoia^oiexpikoixxi1+Boib^oi+1expikoixxi1+Aeia^eiexpikeixxi1+Beib^eiexpikeixxi1.E11

The complete set of amplitudes is Aqi, Aqi and Bqi and Bqi where q=oe satisfies the following relationships:

Aqi=Aqiexpikqixixi1,E12a
Bqi=Bqiexpikqixixi1.E12b

For completeness we calculate the magnetic field in a given elementary layer “i.” By using the Maxwell’s equation Htoti=iωμ01×Etoti, and considering that the magnetic field may be written as Htoti=Hixexpikyy+ikzziωt, we finally obtain the following expression for the variation of the field along the x direction, namely:

ωμ0Hix=Aoic^oiexpikoixxi+Boid^oiexpikoixxi+Aeic^eiexpikeixxi+Beid^eiexpikeixxi.E13

In the previous equations, the new vector fields are set as

c^oi=koikzkykzkz2ϵxik02,E14a
c^ei=ϵxik02kykei0,E14b
d^oi=koikzkykzkz2ϵxik02,E14c
d^ei=ϵxik02kykei0.E14d

We point out that the vector fields c^oi and d^oi can be given in units of k02, whereas c^ei and d^ei can be expressed in units of k03.

2.3.2. Boundary conditions for anisotropic layered media

Once we have a complete description of the wave fields in every elementary layer of our metamaterial, we have to impose some boundary conditions at the interfaces x=xi. The components of the electric field Etoti and the magnetic field Htoti lying on the planes x=xi must be continuous. Particularly these boundary conditions may be expressed in a matrix form as

Divi=Di+1vi+1.E15

We defined the following matrix, explicitly given as

Di=koikoikykzkykz00kz2ϵxik02kz2ϵxik02kykzkykzϵxik02keiϵxik02keikz2+ϵxik02kz2+ϵxik0200.E16

On the other hand, we introduced the amplitude column vectors

vi=AoiBoiAeiBei,vi=AoiBoiAeiBei.E17

The matrix formulation can also be used to relate the amplitude vector vi with zero-phase shift at x=xi with the amplitude vector vi exhibiting zero dephase at x=xi1, previously established in Eq. (12a), (12b). For that purpose, we introduce the propagation matrix Pi, which takes into account the amplitude dephasing between the boundaries of each layer. Explicitly we may write

vi=Pi·vi,E18

being

Pi=eikoiwi0000eikoiwi0000eikeiwi0000eikeiwi,E19

where wi=xixi1 denotes the width of the elementary layer “i.” We note that equivalent matrix formulations for anisotropic multilayered media can be found elsewhere [6, 7].

2.3.3. Electromagnetic fields in layered isotropic media

At this point, once we have described the electromagnetic fields in uniaxial media, let us study a multilayered media composed of isotropic materials. Considering an isotropic medium of relative permittivity ϵi, once again, the projection of the wave vector along the positive x direction is set as

kTEi=kTMi=ϵik02ky2+kz2.E20

Formally, kTMi applies to TM-polarized waves and kTEi corresponds to TE-polarized waves. The total electric field of the elementary layer “i” can be set, again, as given in Eq. (7). The part of the electric field that varies along with the spatial coordinate x can be written using the amplitude ATEi (and ATMi), which corresponds to propagating TE (and TM) waves and BTEi (and BTMi) that is related with counter-propagating TE (and TM) waves. This finally reads as

Eix=ATEi a^TEiexpikTEixxi+BTEi b^TEiexpikTEixxi+ATMi a^TMiexpikTMixxi+BTMi b^TMiexpikTMixxi.E21

Note that all these amplitudes have zero dephase at x=xi. Finally, the vectors a^TEi and a^TMi are given in Eq. (3a), (3b) by e^1 and e^2, respectively, which we rewrite as

a^TEi=b^TEi=0kzky,E22a
a^TMi=ky2+kz2kykTMikzkTMi,E22b
b^TMi=ky2+kz2kykTMikzkTMi.E22c

In Eq. (22a) and (22c) we have included the field vectors b^TEi and b^TMi which are associated with counter-propagating waves. Again, we point out that the vector fields a^TEi and b^TEi can be given in units of k0, and the vectors a^TMi and b^TMi can be expressed in units of k02. Similarly as performed in the previous section, the field function Ei can also be set in terms of the wave amplitudes ATEi, ATEi, BTMi, and BTMi with zero dephase at x=xi1. As in the previous section, the amplitudes Aqi, Aqi, Bqi, and Bqi, where q=TETM, are related by Aqi=Aqiexpikqiwi and Bqi=Bqiexpikqiwi. Following the same procedure seen earlier, we calculate the magnetic field in every elementary layer “i.” We finally obtain the following expression for the variation of the field along the x direction, namely

ωμ0Hix=ATEi c^TEiexpikTEixxi+BTEi d^TEiexpikTEixxi+ATMi c^TMiexpikTMixxi+BTMi d^TMiexpikTMixxi.E23

In the previous equation we introduced the vector fields

a^TEi=b^TEi=0kzky,E24a
a^TMi=c^TEi=ky2kz2kykTEikzkTEi,E24b
b^TMi=d^TEi=ky2+kz2kykTEikzkTEi,E24c
c^TMi=d^TMi=ϵik020kzky.E24d

Note that c^TMi=ϵik02a^TEi. We conclude that the amplitudes ATEi and BTEi are associated with TE-polarized waves along the x-axis, that is, TEx waves, and ATMi and BTMi are field amplitudes of TMx waves.

2.3.4. Application of the boundary conditions

At a given interface x=xi, the electromagnetic fields should accomplish the continuity boundary conditions. Note that we can write the four equations derived from the boundary conditions in the following matrix form:

Divi=Di+1vi+1.E25

In the previous matrix equation, we introduced the element

Di=kzkzkykTMikykTMikykykzkTMikzkTMikykTEikykTEiϵik02kzϵik02kzkzkTEikzkTEiϵik02kyϵik02ky.E26

Finally, the amplitude vectors now are represented as

vi=ATEiBTEiATMiBTMi,vi=ATEiBTEiATMiBTMi.E27

The matrix formulation can also be used to relate the amplitude vector vi with zero-phase shift at x=xi with the amplitude vector vi exhibiting zero dephase at x=xi1. For that purpose, we introduce the propagation matrix for the TEx and TMx modes

Pi=eikTEiwi0000eikTEiwi0000eikTMiwi0000eikTMiwi,E28

which takes into account the amplitude phase shift due the finite width wi=xixi1 of each layer. Explicitly we will write vi=Pi·vi, which is formally the same as Eq. (18), previously derived for uniaxial media.

Advertisement

3. Surface modes in isotropic media

The main purpose of this chapter is the analysis of Dyakonov surface waves, which originally was formulated for an isotropic medium and a uniaxial crystal. However, this analysis is developed in Section 5. Here we introduce the most well-known surface waves arisen at the interface between isotropic media of different dielectric constants. In addition, these surface waves will play a relevant role when dealing with metal-dielectric multilayered structures.

The so-called surface plasmon polaritons are waves that propagate along the surface of a conductor, usually a metal [2, 8, 9]. These are essentially light waves that are trapped on the surface, evanescently confined in the perpendicular direction and caused by their interaction with the free electrons of the conductor, the latter oscillating in resonance with the electromagnetic field. To describe these wave fields, we use the matrix formalism applied in the vicinity of a single interface between two isotropic media with different dielectric permittivities.

Let us consider the propagation of bound waves on the interface between two semi-infinite media, which are denoted as medium 1 and medium 2 with dielectric permittivities ϵ1 i ϵ2, respectively. This interface is located at x1=0. For medium 1, the electric and magnetic fields varying along the x-axis are given by Eqs. (22a), (22b), (22c) and (24a), (24b), (24c), (24d) respectively.

As we are only interested in bound states; the elements of the remaining field vectors read as

v1=0BTE10BTM1andv2=ATE20ATM20.E29

The values of the amplitudes ATE1, ATM1, BTE2, and BTM2 are identically zero in case of lack of interaction with external sources, as we assume here. Since we are dealing with bound states, the wavenumbers kTE1, kTM1, kTE2,, and kTM2 are purely imaginary. The application of the boundary conditions, D1v1=D2v2, gives us the following two equations:

0=ϵ2kTM1+ϵ1kTM22ϵ1kTM1ATM2,E30a
BTM1=ϵ2kTM1ϵ1kTM22ϵ1kTM1ATM2.E30b

To accomplish Eq. (30a), the following equation must be satisfied, kTM2/kTM1=ϵ2/ϵ1. This is the dispersion equation of the TMx-polarized surface modes. As kTM1 and kTM2 have a vanishing real part and a positive imaginary part, this requires that ϵ2/ϵ1<0, that is, one of the relative permittivities must be negative. Once we have obtained the relationship between the relative permittivities and the purely imaginary wavenumbers, we may obtain a new ligature involving the field amplitudes, BTM1=ϵ2/ϵ1ATM2, which have been derived from Eq. (30b). We may rewrite the dispersion equation, making use of the definition of kTMi given in Eq. (20), resulting:

kSPP=k0ϵ1ϵ2ϵ1+ϵ2,E31

where kSPP=ky2+kz2 is the SPP wavenumber. In Figure 2 we represent Eq. (31) giving the dispersion relation for TM-bounded modes at frequencies lower than the surface plasmon frequency:

ωSPP=ωp1+ϵ1.E32

Figure 2.

SPP dispersion relation at the interface between a lossless Drude metal (ϵ2ω=1ωp2/ω2) and a dielectric: air in blue (ϵ1=1) and GaAs in red (ϵ1=11.55). Note that kp=ωp/c.

At ω0 the SPP wavenumber tends to zero; however, when ωωSPP we find that kSPP. In addition, radiative modes may arise at higher frequencies, typically ωωp; in such cases, the wave field is not confined near the interface and it will lose its energy by radiation.

Advertisement

4. Multilayered plasmonic lattices

Wave propagation in periodic media can be treated as the motion of electrons in crystalline solids. In fact, formulation of the Kronig-Penney model used in the elementary band theory of solids is mathematically identical to that of the electromagnetic radiation in periodic layered media. Thus, some of the physical concepts used in material physics such as Bloch waves, Brillouin zones, and forbidden bands can also be used here. A periodic layered medium is equivalent to a one-dimensional lattice that is invariant under lattice translation.

Here we will treat the propagation of electromagnetic radiation in a simple periodic layered medium that consists of alternating layers of transparent nonmagnetic materials with different electric permittivities. The layers are set in a way that the x-axis points along the perpendicular direction of the layers. The permittivity profile is given by ϵ1, for x0<x<x1, and ϵ2, for x1<x<x2. In addition, the relative permittivity satisfies the condition of periodicity, ϵx=ϵx+Λ, where w1=x1x0 w2=x2x1 is the thickness of the layers of permittivity ϵ1 ϵ2 and Λ=w1+w2 represents the period of the structure.

According to the Floquet theorem, solutions of the wave equation for a periodic medium may be set in the form Ei+2x+Λ=Eix·expiKΛ. The constant K is known as the Bloch wavenumber. The problem is thus that of determining K and Eix. Finally, the equation

cosKqΛ=12trq.E33

represents the dispersion relation for TE and TM modes, written in a compact way. Representing KTE (KTM), the Bloch wavenumber K associated with the mode TE (TM), and writing

trTE=2cosw1kTE1cosw2kTE2kTE12+kTE22kTE1kTE2sinw1kTE1sinw2kTE2,E34a
trTM=2cosw1kTM1cosw2kTM2ϵ22kTM12+ϵ12kTM22ϵ1ϵ2kTM1kTM2sinw1kTM1sinw2kTM2.E34b

we find the dispersion equation of a binary periodic medium for each polarization.

Neglecting losses in the materials, regimes where trq<2 correspond to real Kq and thus to propagating Bloch waves, when trq>2; however, Kq=/Λ+iKqi, where m is an integer and Kqi is the imaginary part of Kq, which gives an evanescent behavior to the Bloch wave. These are the so-called forbidden bands of the periodic medium. The band edges are set for trq=2.

In Figure 3, the transverse wavenumber reads as

kt=ky2+kz2.E35

Figure 3.

Exact dispersion curves derived from Eq. (33) for a lossless Drude metal/GaAs composite medium with λ=1.55μ m, ωp=13.94fs1 and ϵ1=11.55, for (a) TM modes and (b) TE modes. GaAs layer is always 12 times thicker than the metallic layer.

For ultra-thin metallic layers, for instance, wm=3 nm, the curves resembles ellipses and circumferences for TM and TE modes, respectively. In the particular case of TM modes, a secondary curve surges, in relation with the excitation of SPPs. For increasing values of the metal width, some deviations are evident. Reaching a given value of wm, a large band gap centered at kt=0 surges. Particularly for TM modes, at KTM=0, the two curves collapse. This fact may be understood as that the symmetric and antisymmetric surface modes in the metallic layer approach for SPP in a single metal-dielectric interface, giving ktkSPP [see Eq. (31)]. For illustration, the wavenumber of the SPP propagating on the interface of our Drude metal/GaAs materials yields kSPP=0.746·2π/Λ, assuming that Λ=325nm (associated with the period of a multilayer with metal thickness wm=25 nm).

4.1. Effective medium approach

For near-infrared and visible wavelengths, nanolayered metal-dielectric compounds enable a simplified description of the medium by using the long-wavelength approximation, which involves a homogenization of the structured metamaterial [10, 11]. The effective medium approach (EMA), as Rytov exposed in his seminal paper [12], involves representing MD multilayered metamaterial as an uniaxial plasmonic crystal, whose optical axis is normal to the layers (in our case, the x-axis is the optical axis), a procedure that requires the metallic elements to have a size of a few nanometers. This is caused by the fact that transparency of noble metals is restrained to a propagation distance not surpassing the metal skin depth. In Ref. to this point, recent development of nanofabrication technology makes it possible to create such subwavelength structures. Under this condition, the plasmonic lattice behaves as a uniaxial crystal characterized by a relative permittivity tensor ϵ=ϵxx+ϵyy+zz, where

ϵ=ϵ1ϵ21fϵ2+fϵ1,E36

gives the permittivity along the optical axis, and ϵ=1fϵ1+fϵ2 corresponds to the permittivity in the transversal direction. In the previous equations, f=w2/w1+w2, denotes the filling factor of medium “2” providing the metal rate in a unit cell. The dispersion equations given by the EMA are kx2+kt2=ϵk02, corresponding to TE (o-) waves, where kx represents the Bloch wavenumber KTE, and for TM (e-) waves, we have

kx2ϵ+kt2ϵ=k02,E37

being now kx, the Bloch wavenumber KTM.

The validity of the EMA is related on the assumption that the period Λ is much shorter than the wavelength, that is, Λλ0. Apparently, Eq. (33) is in good agreement with the EMA in the vicinity of kx=0 for TMx waves only. In contrast, propagation along the x-axis, where ky=kz=0, results in large discrepancies. Even small-filling factors of the metallic composite lead to enormous birefringences. Such metamaterials enlarge the birefringence of the effective-uniaxial crystal at least in one order of magnitude in comparison with values shown in Table 1. However, the size of birefringence displayed by extraordinary waves is reduced if w2 increases. On the other hand, the isotropy of the isofrequency curve is practically conserved for ordinary waves.

4.2. Hyperbolic media

As we have seen in Section 3.3, nanolayered metal-dielectric compounds behave like plasmonic crystals enabling a simplified description of the medium by using the long-wavelength approximation [10, 11, 12]. Under certain conditions, the permittivity of the medium set in the form of a second-rank tensor includes elements of opposite signs, leading to a metamaterial of extreme anisotropy [13, 14]. This class of nanostructured media with hyperbolic dispersion is promising metamaterials with a plethora of practical applications from biosensing to fluorescence engineering [15].

Type I hyperbolic media refers to a special kind of uniaxially anisotropic media, that can be described by a permittivity tensor where element ϵ is negative and ϵ is positive. In this case, Eq. (37) leads to a two-sheet hyperboloid. Type II hyperbolic media lead to positive ϵ and negative ϵ, and Eq. (37) gives us a one-sheet hyperboloid [16]. The fulfillment of hyperbolic dispersion allows wave propagation over a wide spatial spectrum that would be evanescent in an ordinary isotropic dielectric. At the optical range, hyperbolic media can be manufactured with metal-dielectric multilayers or metallic nanowires. Multilayered hyperbolic metamaterials at an optical range take advantage of the wide frequency band in which metals exhibit negative permittivity and support plasmonic modes.

The system under analysis is a periodic binary medium, where we take as medium 1 a transparent dielectric medium that is ideally nondispersive. In our three numerical simulations we take a lossless Drude metal where its permittivity is ϵ2=1Ω2 and dielectric media with permittivities: (a) ϵ1=1, (b) ϵ1=2.25 and (c) ϵ1=11.55. Note that frequencies can be expressed in units of the plasma frequency, Ω=ω/ωp.

In Figure 4(a), we represent the permittivities ϵ and ϵ of our plasmonic crystals for a wide range of frequencies. Note that the metal-filling factor takes control on the dissipative effects in the metamaterial; accordingly low values of f are of great convenience. We set f=1/4 in our numerical simulations. For low frequencies, Ω1, the following approximations can be used: ϵfϵ2<0 and 0<ϵϵ1/1f. Therefore, propagating TEx modes (Ex=0) cannot exist in the bulk crystal since it behaves like a metal in these circumstances. On the other hand, TMx waves propagate following the spatial dispersion curve of Eq. (37). This is a characteristic of Type II hyperbolic media. As mentioned earlier, Eq. (37) denotes a hyperboloid of one sheet (see Figure 4(b) for Ω=0.20).

Figure 4.

(a) Variation of relative permittivities ϵ (blue solid line) and ϵ (magenta solid line) as a function of normalized frequency Ω, for the plasmonic crystal including a lossless Drude metal and fused silica (ϵ1=2.25) as dielectric material. Ω1 and Ω2 yield 0.359 and 0.759, respectively. (b) and (c) plot Eq. (37) in the kxky plane for extraordinary waves (TMx modes) for a plasmonic effective crystal, including a dielectric of permittivity ϵ1=2.25, in the range Ω<1. Solid line corresponds to kz=0 and shaded regions are associated with harmonic waves with kz>0 (nonevanescent fields).

Furthermore, the hyperbolic dispersion exists up to a frequency

Ω1=11+ϵ11f/f,E38

for which ϵ=0. For slightly higher frequencies, both ϵ and ϵ are positive and Eq. (37) becomes an ellipsoid of revolution. Since its minor semi-axis is Ωϵ, the periodic multilayer simulates a uniaxial medium with positive birefringence. Raising the frequency even more, ϵ diverges at

Ω2=11+ϵ1f/1f,E39

leading to the so-called canalization regime. In general, Ω1<Ω2 provided that f<1/2. Beyond Ω2, Eq. (37) turns to a hyperboloidal shape. In the range Ω2<Ω<1; however, the dispersion curve has two sheets (Type I hyperbolic medium). Figure 4(c) illustrates this case. Note that the upper limit of this hyperbolic band is determined by the condition ϵ=0 or in an equivalent way, ϵ2=0, occurring at the plasma frequency.

Advertisement

5. Dyakonov surface waves

Dyakonov surface waves (DSWs) are another kind of surface waves, supported at the interface between an optically isotropic medium and a uniaxial-birefringent material. In the original work by Dyakonov (English version was reported in 1988 [3]), the optical axis of the uniaxial medium was assumed in-plane with respect to the interface. This is the case we deal with here.

The importance of DSWs for integrated optical applications, such as sensing and nano-waveguiding, was appreciated in a series of papers [17, 18]. Indeed Dyakonov-like surface waves also emerge in the case that a biaxial crystal [19] or a structurally chiral material [20] takes the place of the uniaxial medium. The case of metal-dielectric (MD) multilayers as structurally anisotropic media is especially convenient since small-filling fractions of the metallic inclusions enable metamaterials with an enormous birefringence, thus enhancing density of DSWs and relaxing their prominent directivity [21, 22, 23].

5.1. Dispersion equation of DSWs

The system under study is the plotted in Figure 5, where we have two semi-infinite media, one of them is isotropic and the second one is an MD lattice. In our case, the indices “1” and “2” make reference to the plasmonic lattice and the isotropic medium, respectively. We have previously reported a comprehensive analysis of this case in [4]. As we have seen earlier, the plasmonic lattice can be taken as an effective uniaxial crystal. In this case, the permittivity along its optical axis, ϵz1=ϵ, is given by Eq. (36); also, the permittivity in the transverse direction ϵx1=ϵy1=ϵ may be appropriately averaged. From hereon, the permittivity ϵ2 of the isotropic medium in x>0 will be denoted by ϵ. Note that our analysis serves for natural birefringent materials characterized by permittivities ϵ and ϵ.

Figure 5.

Schematic setup under study, consisting in a semi-infinite dielectric-metal superlattice x<0 and an isotropic substrate x>0.

Since we treat the plasmonic lattice as a uniaxial crystal, we may establish analytically the diffraction equation that gives the 2D wave vector kD=0kykz in x=0. For that purpose, we follow Dyakonov [3] by considering hybrid-polarized surface modes. In the isotropic medium we consider TEx (Ex=0) and TMx (Hx=0) waves whose wave vectors have the same real components ky and kz in the plane x=0. Therefore the electric field in both media may be set as

Etot=Exexpikyy+ikzziωt.E40

Moreover, these fields are evanescent in the isotropic medium and in the superlattice. In the anisotropic medium ( x<0) the evanescent electric amplitude can be written as

Ex=Bo1 b^o1expiko1x+Be1 b^e1expike1x,E41

where the ordinary and extraordinary waves in the effective uniaxial medium decay exponentially with rates given by κo=iko1 and κe=ike1, respectively. Taking the formulation given in Section 2.3.1, the amplitudes Ao1 and Ae1 are identically zero. In the isotropic medium (x>0) the amplitude of the electric field is

Ex=ATE2 a^TE2expikTE2x+ATM2 a^TM2expikTM2x,E42

where the evanescent decay for TE and TM modes is κ=ikTE2=ikTM2. Now the amplitudes BTE2 and BTM2 are zero.

Once we have the amplitudes in both sides of the interface, we apply the boundary conditions at x=0, D1·v1=D2·v2, where D1 is provided by Eq. (16), D2 from Eq. (26), v1 from Eq. (17), and v2 from Eq. (27). This equation reduces to:

0Bo10Be1=MhATE20ATM20,E43

where the transmission matrix Mh=D11·D2 establishes a relationship between the amplitudes of hybrid polarization modes. Using the elements Mij of the matrix Mh, and defining Mi and Ma as

Mi=M11M13M31M33=ikzκo+κ2kz2k02ϵk02kyϵκo+κϵ2κokz2k02ϵκe+κky2κekz2k02ϵikzκκeϵ+ϵκo22κeϵkz2k02ϵ,E44a
Ma=M21M23M41M43=ikzκoκ2kz2k02ϵk02kyϵκoκϵ2κokz2k02ϵκκeky2κekz2k02ϵikzκκeϵϵκo22κeϵkz2k02ϵ.E44b

Eq. (43) can be rewritten as a set of two independent matrix equations, namely

00=Mi·ATE2ATM2,E45a
Bo1Be1=Ma·ATE2ATM2.E45b

Note that Mi governs the amplitudes ATE2 and ATM2 of the isotropic medium, and Ma (also Mh) may be used to determine the amplitudes Bo1 and Be1 of the anisotropic medium.

Dyakonov equation is obtained by means of letting the determinant of Mi equal to zero, giving

k02ky2ϵκ+κeϵκo+ϵκ=κokz2κ+κoϵκo2+ϵκκe,E46

which provides a spectral map of allowed values kykz. After fairly tedious algebraic transformations we can reduce Eq. (46) to a more convenient form [3].

κ+κeκ+κoϵκo+ϵκe=ϵϵϵϵk02κo.E47

Assuming that ϵ, ϵ and all decay rates are positive, the additional restriction ϵ<ϵ<ϵ can be deduced for the existence of surface waves. As a consequence, positive birefringence is mandatory to ensure a stationary solution of Maxwell’s equations. Therefore, layered superlattices supporting Dyakonov-like surface waves cannot be formed by all dielectric materials [21].

5.2. DSWs in nano-engineered materials

To illustrate the difference between using conventional birefringent materials and plasmonic crystals, we solve Eq. (47) for liquid crystal E7 with ϵ=2.98 i ϵ=2.31 at a wavelength of λ0=1.55μ m and N-BAK1 substrate of dielectric constant ϵ=2.42. In this case, DSWs propagate in a narrow angular region Δθ=θmaxθmin, where θ stands for the angle between the in-plane vector kykz and the optical axis. More specifically, the angular range yields Δθ=0.92 around the mean angle θ¯=26.6. It appears that the resulting angular range Δθ is short. But this range would become much smaller when using other optical crystals like quartz, exhibiting a common birefringence. In order to gain in angular extent Δθ, we consider a GaAs-Ag crystal (ϵ1=12.5 and ϵ2=103.3, where we neglect losses) that leads to values of ϵ=14.08, derived from Eq. (36), and ϵ=0.92 with a metal-filling factor f=0.10. Form birefringence now yields Δn=2.79. Moreover, solutions to Eq. (47) can be found in the region of angles comprised between θmin=39.0 and θmax=71.3. It is important to point out that the total angular range of existence of DSWs, Δθ=32.3, grows by more than an order of magnitude. Figure 6 shows the dispersion curve for DSWs; one can observe that θmin is attained under the condition κ=0 (red solid line), where TEx and TMx waves are uniform in the substrate x>0. Looking at the other side of the dispersion curve, θmax is established by ke=0, shown as a black solid line, for which the extraordinary wave will not decay spatially at x.

Figure 6.

Dispersion Eq. (47) for DSWs (dotted-dashed line) propagating on the interface of a semi-infinite dielectric-metal lattice with metal filling factor f = 0.1. The solid elliptical line and the black dashed line are associated with homogeneous extraordinary waves κe=0 and homogeneous ordinary waves κo=0, respectively. The isofrequency curve κ=0 of isotropic N-BAK1 are represented in by the red solid line, which applies for TEx and TMx waves. Adapted with permission from [4] of copyright ©2013 IEEE photonics society.

Consequently, the solution for Eq. (47) can be traced near the curves κ=0 and κe=0; thus, DSWs are always found close to the crosspoint P0ky0kz0 of both curves.

5.2.1. Nonlocal effects

As we discussed in Section 4.1, the EMA is limited to metallic slabs’ width wmλ0. However, this condition must be taken into account with care, since the skin depth of noble metals is extremely short, δc/ωp. For instance, we estimate δ=24 nm in the case of silver. If the metal thickness is comparable to its skin depth, the EMA will substantially deviate from exact calculations. Note that experimental studies from multilayer optics rarely incorporate metallic slabs with a thickness below 10 nm.

We emphasize that moderate changes in the birefringence of the plasmonic crystal will substantially affect the existence of DSWs. More specifically, an enlargement of ϵ driven by increasing wm, provided f is fixed (see Section 4.1), will lead to a significant modification of the DSW dispersion curves. Ultimately, this phenomenon is clearly attributed to nonlocal effects in the effective-medium response of nanolayered metamaterials [24], which is associated with a strong variation of the fields on the scale of a single layer.

We conclude that, in order to excite DSWs, one may counterbalance the decrease of birefringence in the plasmonic lattice by means of a dielectric substrate of higher index of refraction. To illustrate this matter, the dispersion Eq. (47) for DSWs is represented in Figure 7, in addition to using the values of ϵ and ϵ from nonlocal estimators [4]. When wm grows but f is kept fixed, the dispersion curve of the Dyakonov surface waves tends to approach the optic axis.

Figure 7.

Solutions to Dyakonov equation, drawn in dotted-dashed lines, for a MD lattice with the same filling factor f=0.10 but different wm, using estimates from the nonlocal birefringence approach described in [4]. Note that the red solid line designates the isofrequency curve of isotropic substrate (a) N-BAK1 and (b) P-SF68. Adapted with permission from [4] of copyright ©2013 IEEE photonics society.

For an N-BAK1 substrate, as shown in Figure 7(a), θmax=68.2 and 58.7 for wm=3 nm and 6 nm, respectively. Also, θmin=37.6 and 32.1 for these two cases. As a consequence the angular range Δθ shrinks when wm increases. In the limit ϵϵ, which occurs for wm=10.3 nm using a N-BAK1 substrate, DSWs are not supported at the interface of the MD lattice and the isotropic dielectric. A substrate with greater relative permittivity ϵ would be necessary. For example, if we use a substrate with greater relative permittivity ϵ as P-SF68 [see Figure 7(b)], DSWs exist for wm=12 nm with an angular range Δθ=12.5.

5.2.2. Dissipative effects

Up to now, we have avoided another important aspect of plasmonic devices namely dissipation in metallic elements. In this regard, effective permittivities are fundamentally complex, and consequently the Dyakonov Eq. (47) is expected to give complex values of kykz. This procedure has been discussed by Sorni et al. [25] recently. In order to tackle this problem, we evaluate numerically the value of the Bloch wavenumber kz for a given real value ky. The spatial frequency kz becomes complex since Imϵm=8.1. As a consequence, the surface wave cannot propagate indefinitely, undergoing an energy attenuation given by l=2Imkz1. Furthermore, we naturally assume that the real part of the parameters κ,κo, and κe are all positive. These positive values correlate with a decay at x and thus with a confinement of the wave near x=0.

Figure 8(a) depicts the dispersion curve corresponding to dissipative DSWs, for the case of a plasmonic MD lattice with f=0.10 and wm=12 nm. We used a commercial software (COMSOL Multiphysics) based on the finite-element method (FEM) in order to perform our numerical simulations. We cannot observe surface waves by setting an N-BAK1 substrate with n=1.56, suggesting that this is a retardation effect. More specifically, Figure 8(a) shows the isofrequency curves when n=1.95, corresponding to P-SF68.

Figure 8.

(a) Isofrequency curve that corresponds to hybrid surface waves existing at the boundary between a semi-infinite P-SF68 substrate and a lossy MD superlattice of f=0.10 and wm=12 nm. (b) Ratio of Im(kz) over re(kz) representing dissipation effects in the propagation of DSWs. (c) Three contour plots of the magnetic field Hx computed using the finite-element method. The superlattice is set on the left, for which only one period is represented. Capital letters a, B, and C, designate the transverse spatial frequencies ky=0.8k0, 1.2k0, and 1.6k0, respectively. Adapted with permission from [4] of copyright ©2013 IEEE photonics society.

We observe that the dispersion curve for dissipative DSWs is flatter and larger than the curve obtained by neglecting losses. Specifically θmax=49.9 and θmin=23.7, giving an angular range Δθ=26.2. Figure 8(b) shows Im(kz)/Re(kz) in the range of existence of the surface waves. In these two figures, capital letters A, B, and C designate the transverse spatial frequencies ky=0.8k0, 1.2k0, and 1.6k0, respectively. Figure 8(c) shows the magnetic field Hx for the three different cases denoted by capital letters A, B, and C. Note that in the case of paraxial surface waves, for which ky reaches a minimum value (case A), one achieves ImkzRekz as depicted in Figure 8(b). This is induced by an enormous shift undergone by the field maximum in the direction to the isotropic medium, as shown in Figure 8(c), where dissipation effects are barely disadvantageous on surface-wave propagation. Moreover, this would be consistent with a condition ReκReκe. On the other hand, for nonparaxial waves, having the largest values of ky, the fields show slow energy decay inside the plasmonic superlattice. In case C, the magnetic field Hx is localized around the metallic layer and takes significant values far from the boundary of the substrate. As a consequence, losses in the metal translate into a significant rise in the values of Im(kz).

5.3. New families of DSWs in lossy media

In this section we carry out a thorough analysis of DSWs that takes place in lossy uniaxial metamaterials. Special emphasis is put when the effective-medium approach induces satisfactory results. The introduction of losses leads to a transformation of the isofrequency curves, which deviates from spheres and ellipsoids, as commonly considered by ordinary and extraordinary waves, respectively. As a consequence, one can find two different families of surface waves as reported by Sorni et al. [25]. One family of surface waves is directly related with the well-known solutions derived by Dyakonov [3]. Importantly, the existence of a new family of surface waves is revealed, closely connected to the presence of losses in the uniaxial effective crystal. We point out that the solutions to Dyakonov equation presented earlier are partial ones insofar as the z-component of the wavevector is kept real valued. Nevertheless, the whole set of solutions includes all possible wavevectors that feature a complex-valued kx.

Advertisement

6. Dyakonov surface waves in hyperbolic media

In this section we perform a thorough analysis of DSWs taking place in semi-infinite MD lattices exhibiting hyperbolic dispersion. Part of this section was previously reported by Zapata-Rodríguez et al. [26]; we point out that recently further studies on DSW in hyperbolic metamaterials have been reported by other authors [27]. Our approach puts emphasis on the EMA. Under these conditions, different regimes can be found including DSWs with nonhyperbolic dispersion. The system under analysis is again as depicted in Figure 5. For simplicity, we assume that dielectric materials are nondispersive; indeed, we set ϵ=1 and ϵd=2.25 in our numerical simulations. Furthermore, Drude metals are included, and frequencies will be expressed in units of its plasma frequency, Ω=ω/ωp. Again, Dyakonov Eq. (47) provides the spectral map of wave vectors kD=0kykz. Note that in this section, spatial frequencies will be expressed in units of kp.

In the special case of the surface wave propagation perpendicular to the optical axis (kz=0), Eq. (47) reveals the following solution: ϵκo+ϵκ=0. In the case: ϵ<0 and ϵ<ϵ, this equation has the well-known solution

ky=Ωϵϵϵ+ϵ,E48

which resembles the dispersion equation of conventional SPPs [see Eq. (31)]. Here we have purely TMx polarized waves, as expected. Note that no solutions to Eq. (47) can be found, in the form of surface waves, considering wave propagation parallel to the optical axis (ky=0) for hyperbolic metamaterials: ϵϵ<0. That means that a threshold value of ky can be found for the existence of surface waves.

Next we describe a specific configuration governing DSWs, subject to a low value of the refractive index n=ϵ, namely ϵ<ϵ (ϵ<ϵ) occurring at low and moderate frequencies; other cases are treated elsewhere [26]. In the effective-uniaxial medium, it is easy to realize that κ<κo and also κe<κo. Under these circumstances, all brackets in Dyakonov Eq. (47) are positive provided ϵκo+ϵκe>0. This happens within the spectral band Ω0<Ω<Ω1, where

Ω0=11+ϵ/f+ϵd1f/f.E49

Note that Ω0=0.292 in our numerical simulation.

In Figure 9(a) and (b), we illustrate the dispersion equation of DSWs for two different frequencies within the spectral range 0<Ω<Ω0. In these cases, the dispersion curve approaches a hyperbola. We find a bandgap around kz=0 in Figure 9(a), unlike what occurs in (b). Note that hybrid solutions near kz=0 are additionally constrained to the condition kyΩϵ [see also Eq. (48)], which is a necessary condition for κe to exhibit real and positive values. We consider the quasi-static regime (Ω0) where kD=kDΩ to determine the asymptotes of the hyperbolic-like DSW dispersion curve. Under this approximation, κ=kD, κo=kD and κe=ΘkD, where

Θ=cos2θ+ϵ/ϵsin2θ,E50

being ky=kDcosθ and kz=kDsinθ. These asymptotes establish a canalization regime leading to a collective directional propagation of DSW beams [4, 28]. At this point it is necessary to remind that the asymptotes of the e-waves dispersion curve, in the kykz plane, have slopes satisfying the condition θD<θe, as illustrated in Figure 9(b).

Figure 9.

Solutions to Eq. (47), drawn in solid line, providing the spatial dispersion of DSWs which can exist in the arrangement of Figure 5, at different frequencies: (a) Ω=0.20, (b) Ω=0.28, and (c) Ω=0.85. Here, the isotropic medium is air and the multi-layered metamaterial has a filling factor f=0.25 and. Also, we include equations κ=0 (dotted line) and κe=0 (dashed line). Adapted with permission from [26] of copyright ©2013 Optical Society of America.

In the high-frequency band Ω2<Ω<1 we find that ϵ<0<ϵ, as occurs in Figure 9(c). Note the relevant proximity of DSW dispersion curve to κe=0, the same way we also find in Figure 9(a) and (b). Conversely it crosses the e-wave hyperbolic curve at two different points, where solutions to the Dyakonov equation begin and end, respectively. It is clear that the angular range of DSWs now turns to be significantly low.

6.1. DSWs in band-gap hyperbolic media

In previous sections we demonstrated that the presence of metallic nano-elements leads to nonlocal effects and dissipation effects which reshape the propagation dynamics of the surface signal. Here, we briefly discuss the extraordinary favorable conditions which may appear in band-gap metal–insulator-layered media for the existence of DSWs. As reported thoroughly by Miret et al. [29], engineering secondary bands by tuning the plasmonic-crystal geometry may lead to a controlled optical anisotropy, which is markedly dissimilar to the prescribed hyperbolic regime that is derived by the EMA, however, assisting the presence of DSWs on the interface between such hyperbolic metamaterial and an insulator.

In particular, a surface wave propagating on an Ag-Ge grating was considered, where the environment medium that is set above the metallic grating is formed by SiO2. If the metal-filling factor was f=0.25, the effective permittivities of the anisotropic metamaterial would be estimated as ϵ=11.48+i2.05 and ϵ=25.96+i0.14 at a wavelength of λ0=1550 nm. Disregarding losses, the DSW dispersion curve describes an incomplete hyperbolic curve, finding an endpoint under the condition κe=0, where the extraordinary wave breaks its confinement in the vicinities of the isotropic-uniaxial interface [26].

Considering now a realistic nanostructure consisting of Ag layers of w2=40 nm interspersed between Ge layers of w1=120 nm, thus maintaining a metal-filling factor of f=0.25 as analyzed earlier, a first TM band with hyperbolic-like characteristics dominates at high in-plane frequencies kt. Additionally, a second band emerges for TM Bloch modes, which exhibits a moderate anisotropy, demonstrating near-elliptical dispersion curves (with positive effective permittivities) and positive birefringence. Furthermore, TE modal dispersion is roughly isotropic. Therefore, satisfactory conditions are found near the second TM band for the existence of Dyakonov-like surface waves. Finally, in order to numerically obtain the dispersion curves and wave fields associated with DSWs, one may follow the same computational procedure followed by Zapata-Rodríguez et al. and [4, 26].

Advertisement

7. Summary

In this chapter we provide several methods to analytically calculate and numerically simulate modal propagation of DSWs governed by material anisotropy. We focused on the spatial properties of DSWs at optical and telecom wavelengths, particularly using uniaxial metamaterials formed of dielectric and metallic nanolayers. We developed an electromagnetic matrix procedure enabling different aspects reviewed in this chapter, specially adapted to complex multilayered configurations. The EMA results are particularly appropriate for the characterization of the form birefringence of a multilayered nanostructure, though limitations driven by the layers width have been discussed. Through a rigorous full-wave analysis, we showed that hybrid-polarized surface waves may propagate obliquely at the boundary between a plasmonic bilayer superlattice and an isotropic loss-free material. We revealed that realistic widths of the slabs might lead to solutions which deviate significantly from the results derived directly from the EMA and Dyakonov analysis. Finally, we showed that excitation of DSWs at the boundary of an isotropic dielectric and a hyperbolic metamaterial enables a distinct regime of propagation. It is important to note that the properties of the resulting bound states change drastically with the index of refraction of the surrounding medium, suggesting potential applications in chemical and biological sensing and nanoimaging.

This chapter was supported by the Qatar National Research Fund (Grant No. NPRP 8-028-1-001) and the Spanish Ministry of Economy and Competitiveness (Grants No. TEC2014-53727-C2-1-R and TEC2017-86102-C2-1R).

References

  1. 1. Zenneck J. Über die Fortpflanzung ebener elektromagnetischer wellen längs einer ebenen Lieterfläche und ihre Beziehung zur drahtlosen Telegraphie. Annalen der Physik (Leipzig). 1907;23:846-866
  2. 2. Maier SA. Plasmonics: Fundamentals and Applications. New York: Springer; 2007
  3. 3. D'yakonov MI. New type of electromagnetic wave propagating at an interface. Sov. Phys. JETP. 1988;67:714-716
  4. 4. Zapata-Rodríguez CJ, Miret JJ, Sorni JA, Vuković SM. Propagation of dyakonon wave-packets at the boundary of metallodielectric lattices. IEEE Journal of Selected Topics in Quantum Electronics. 2013;19:4601408
  5. 5. Takayama O, Crasovan LC, Johansen SK, Mihalache D, Artigas D, Torner L. Dyakonov surface waves:A review. Electromagnetics. 2008;28:126-145
  6. 6. Berreman DW. Optics in stratified and anisotropic media: 4×4 matrix formulation. Journal of the Optical Society of America. 1972;62(4):502-510
  7. 7. Hodgkinson IJ, Kassam S, Wu QH. Eigenequations and compact algorithms for bulk and layered anisotropic optical media: Reflection and refraction at a crystal-crystal interface. Journal of Computational Physics. 1997;133(1):75-83
  8. 8. Barnes WL, Dereux A, Ebbesen TW. Surface plasmon subwavelength optics. Nature. 2003;424(6950):824
  9. 9. Zayats AV, Smolyaninov II, Maradudin AA. Nano-optics of surface plasmon polaritons. Physics Reports. 2005;408:131-314
  10. 10. Yariv A, Yeh P. Electromagnetic propagation in periodic stratified media. II. Birefringence, phase matching, and x-ray lasers. Journal of the Optical Society of America. 1977;67:438-448
  11. 11. Vuković SM, Shadrivov IV, Kivshar YS. Surface Bloch waves in metamaterial and metal-dielectric superlattices. Applied Physics Letters. 2009;95:041902
  12. 12. Rytov SM. Electromagnetic properties of a finely stratified medium. Soviet Physics - JETP. 1956;2:466-475
  13. 13. Smith DR, Schurig D, Rosenbluth M, Schultz S. Limitations on subdiffraction imaging with a negative refractive index slab. Applied Physics Letters. 2003;82:1506-1508
  14. 14. Smolyaninov II, Hwang E, Narimanov E. Hyperbolic metamaterial interfaces: Hawking radiation from Rindler horizons and spacetime signature transitions. Physical Review B. 2012;85:235122
  15. 15. Guo Y, Newman W, Cortes CL, Jacob Z. Applications of hyperbolic metamaterial substrates. Advances in OptoElectronics. 2012;ID 452502
  16. 16. Cortes CL, Newman W, Molesky S, Jacob Z. Quantum nanophotonics using hyperbolic metamaterials. Journal of Optics. 2012;14:063001
  17. 17. Takayama O, Crasovan L, Artigas D, Torner L. Observation of Dyakonov surface waves. Physical Review Letters. 2009;102:043903
  18. 18. Takayama O, Artigas D, Torner L. Lossless directional guiding of light in dielectric nanosheets using Dyakonov surface waves. Nature Nanotechnology. 2014;9:419-424
  19. 19. Walker DB, Glytsis EN, Gaylord TK. Surface mode at isotropic uniaxial and isotropic-biaxial interfaces. Journal of the Optical Society of America. A. 1998;15:248-260
  20. 20. Gao J, Lakhtakia A, Lei M. Dyakonov-Tamm waves guided by the interface between two structurally chiral materials that differonly in handedness. Physical Review A. 2010;81:013801
  21. 21. Vuković SM, Miret JJ, Zapata-Rodríguez CJ, Jaksić Z. Oblique surface waves at an interface of metal-dielectric superlattice and isotropic dielectric. Physica Scripta. 2012;T149:014041
  22. 22. Ghasempour Ardakani A, Naserpour M, Zapata Rodríguez CJ. Dyakonov-like surface waves in the THz regime. Photonics and Nanostructures - Fundamentals and Applications. 2016;20:1-6
  23. 23. Miret JJ, Zapata-Rodrïguez CJ, Jaksić Z, Vuković SM, Belić MR. Substantial enlargement of angular existence range for Dyakonov-like surface waves at semi-infinite metal-dielectric superlattice. Journal of Nanophotonics. 2012;6. DOI: 063525
  24. 24. Elser J, Podolskiy VA, Salakhutdinov I, Avrutsky I. Nonlocal effects in effective-medium response of nanolayered metamaterials. Applied Physics Letters. 2007;90:191109
  25. 25. Sorni JA, Naserpour M, Zapata-Rodríguez CJ, Miret JJ. Dyakonov surface waves in lossy metamaterials. Optics Communication. 2015;355:251-255
  26. 26. Zapata-Rodríguez CJ, Miret JJ, Vuković S, Belić MR. Engineered surface waves in hyperbolic metamaterials. Optics Express. 2013;21:19113-19127
  27. 27. Xiang Y, Guo J, Dai X, Wen S, Tang D. Engineered surface Bloch waves in graphene-based hyperbolic metamaterials. Optics Express. 2014;22:3054-3062
  28. 28. Jacob Z, Narimanov EE. Optical hyperspace for plasmons: Dyakonov states in metamaterials. Applied Physics Letters. 2008;93:221109
  29. 29. Miret JJ, Sorní JA, Naserpour M, Ghasempour Ardakani A, Zapata-Rodríguez CJ. Nonlocal dispersion anomalies of Dyakonov-like surface waves at hyperbolic media interfaces. Photonics and Nanostructures - Fundamentals and Applications. 2016;18:16-22

Written By

Carlos J. Zapata-Rodríguez, Slobodan Vuković, Juan J. Miret, Mahin Naserpour and Milivoj R. Belić

Submitted: 10 May 2017 Reviewed: 17 January 2018 Published: 02 May 2018