Expand this Topic clickable element to expand a topic
Skip to content
Optica Publishing Group

Generalized method for retrieving effective parameters of anisotropic metamaterials

Open Access Open Access

Abstract

Electromagnetic or acoustic metamaterials can be described in terms of equivalent effective, in general anisotropic, media and several techniques exist to determine the effective permeability and permittivity (or effective mass density and bulk modulus in the context of acoustics). Among these techniques, retrieval methods use the measured reflection and transmission coefficients (or scattering coefficients) for waves incident on a metamaterial slab containing few unit cells. Until now, anisotropic effective slabs have been considered in the literature but they are limited to the case where one of the axes of anisotropy is aligned with the slab interface. We propose an extension to arbitrary orientations of the principal axes of anisotropy and oblique incidence. The retrieval method is illustrated in the electromagnetic case for layered media, and in the acoustic case for array of tilted elliptical particles.

© 2014 Optical Society of America

1. Introduction

Metamaterials have generated an intense research interest in recent years for their ability to produce unusual properties of wave propagations, in electromagnetism, acoustics, elasticity and for water waves, see, e.g., [15]. These structures are most commonly made of a periodic arrangement of unit cells, that may be very complex, and there is a need for proper characterization of the resulting behavior at large scale in order to better understand and further exploit them.

The notion of effective parameters for metamaterials has been introduced in analogy with the notion of homogeneous parameters defined for natural materials, being inhomogeneous at small scales. This is a simplifying approximation expected to be valid when the scales of the material inhomogenities are much smaller than the scale associated to the spatial variation in the acoustic, or electromagnetic, field. In the low frequency regime, homogenization theory provides the effective parameters unambiguously (see, e.g, [6]) and recent studies have shown the possibility to extend such approach to higher frequencies, typically when the unit cell presents resonances [7]. Alternatively, and we may say more pragmatically, a popular method consists in determining the effective parameters using an inversion technique [811] assuming that the reflection and transmission coefficients by a metamaterial slab illuminated by a plane wave are known. The idea beyond such methods is to get material properties with broadband validity and to avoid mathematical approaches hardly suitable in practice. For resonant structures, this means that the dependence of the effective parameters on the frequency are looked for, and this may lead to negative effective parameters [1115]. For non resonant structures, the parameters are expected to remain relatively unchanged but the idea is to get more accurate parameter values in the frequency range of interest [1618] (more accurate than the values obtained from homogenization). As pointed out in [19], the main virtue of this technique is that it does not require any assumption except that the concept of effective medium is valid, which is not incidental [2024].

The retrieval method leads to the effective refractive index n and the impedance ratio ξ (between, say, air and the effective medium) using the reflection and transmission coefficients through a thin slab containing few unit cells of the metamaterial. In its initial form, the method was adapted to isotropic effective medium and the method used the scattering properties of a slab illuminated by plane waves incident normally to the slab interfaces. Firstly proposed by Smith in 2002 [12] in the context of electromagnetic waves, the method has been improved to the case of an asymmetric slab [9], to the case of an isotropic effective medium with obliquely incidences [25] and to the case of an anisotropic effective medium with normal incidence [26] and with oblique incidence [27]; however, in [26, 27], it was assumed that the interfaces were parallel to one principal axis of anisotropy. In addition to these improvements, the ambiguities in the inversion technique have been clearly addressed [19, 28].

Note that in 2007, Fokin and coauthors propose a slightly different retrieval method [11] in the context of acoustic waves. Although the situation in both contexts of waves is strictly the same in two dimensions (except there is no polarization option in acoustics), the two literatures, in acoustic and electromagnetism, remain largely separated.

In this paper, the main remaining restriction of the retrieval methods, namely the assumption that the principal directions of anisotropy coincide with the directions of the unit cell, is addressed (angle α in Fig. 1). This is done together considering oblique incidence of the plane wave on the air-metamaterial interface (angle θ). This may be of practical importance. Indeed, retrieval methods have been applied to characterize anisotropic metamaterial with α = 0, and have revealed interesting phenomena: the design of anisotropic metamaterials with negative index operating at variety of incidence angles [29], the highly directive emission of a source embedded in an anisotropic metamaterial [30, 31], the design of a slab lens with low loss using an anisotropic metamaterial with negative properties involving pure dielectric rod-type structure, the rainbow-like emission of a source placed in a uniaxial metamaterial slab [32], the inversion of the Brewster angle in anisotropic metamaterials with negative properties. We think that the ability to inspect effective properties when the resulting anisotropy is not aligned with the unit cell (thus, α ≠ 0) may open doors to interesting applications.

 figure: Fig. 1

Fig. 1 (a) Typical configuration of the real medium, the reflection R and transmission T through a slab containing few cells (otherwise x-periodic in the x-direction) are calculated for an obliquely incident plane wave with angle θ and wavenumber k. (b) Effective anisotropic medium leading to the same scattering properties (R and T).

Download Full Size | PDF

The paper is organized as follows. In Sec. 2, the 2D wave equations for electromagnetism and acoustic are briefly recall, to clarify the equivalences. In Sec. 3, the retrieval method for anisotropic effective media and oblique incidence is presented. This is done following [11] and by considering first the case α = 0 (Sec. 3.1). Although a similar calculation has been already done in [27], this “warm up” allows to better understand the case α ≠ 0 (presented in Sec. 3.2). Indeed, contrary to what is said in [28], the case of an isotropic slab (even at oblique incidence) has strong differences with the case of an anisotropic slab with arbitrary principal axes. Sec. 4 presents results in the particular case of layered media, for which available results coming from the homogenization theory have been tested [33]. In this particular configuration, our retrieval technique is validated, and the variations of the retrieved effective parameters with the frequency and the anisotropy angle are inspected. We also report in Sec. 5 results of the retrieval method for array of tilted elliptical particles (the acoustic case is considered here).

2. 2D Wave equation in anisotropic media for electromagnetic and acoustic waves

In this section, the 2D wave equations for wave propagating in an anisotropic medium is briefly derived, for electromagnetic polarized waves and for acoustic waves. A time harmonic dependence eiωt is assumed in the following.

First, we define in electromagnetism the permittivity tensor εdiag and permeability tensor μdiag, and in acoustics, the mass density tensor ρdiag (the bulk modulus B is assumed to be scalar) expressed in the principal directions (X, Y, Z)

εdiag=(εX000εY000εZ),μdiag=(μX000μY000μZ),inelectromagnetism,ρdiag=(ρX000ρY000ρZ),inacoustics.
The equations can be solved in a (x, y, z = Z) space deduced from (X, Y, Z) by a rotation R of angle α in the (x, y) plane,
R=(cosαsinα0sinαcosα0001)
and we get
inelectromagnetism{×E=ik(RtμdiagR)H,×H=ik(RtμdiagR)E,inacoustics{p=ik(RtρdiagR)u,Bu=ikp,
which correspond to the Maxwell equations for the electric field E and the magnetic field H, and to the Euler equations for the pressure p and velocity u, respectively. In the 2D case, the medium does not couple the electric E and magnetic H fields and for polarized waves, either Transverse Magnetic (TM) for which H = (0, 0, H) or Transverse Electric (TE) for which E = (0, 0, E). A wave equation for respectively E or H similar to Eq. (5) is obtained, namely,
[(1/εx1/εxy1/εxy1/εy)H]+μZk2H=0,TMwaves,[(1/μx1/μxy1/μxy1/μy)E]+εZk2E=0,TEwaves.

In the context of acoustics, 2D case corresponds to a configuration where the pressure field p(x, y) and velocity u(x, y) do not depend on z and we get

[(1/ρx1/ρxy1/ρx1/ρy)p]+k2Bp=0,
with the definitions and correspondances indicated in the Table 1.

Tables Icon

Table 1. Coefficients in the 2D wave equation for anisotropic media in electromagnetism and acoustics.

3. Retrieval method

The basic idea of the retrieval method lays in the assumption that the scattering properties of the medium, Fig. 1(a), can be described by the scattering properties of an equivalent effective homogeneous anisotropic medium, Fig. 1(b).

We start with the warm-up case α = 0, already considered in [27]. First, (R, T) are determined in the configuration of Fig. 1(b), afterwards the retrieval method is presented. Then, the case α ≠ 0 is presented.

3.1. Case of anisotropy direction α = 0

To begin with, the retrieval method is presented assuming α = 0. Outside the slab, say, in air, the wave field H satisfies (Δ + k2)H = 0. In the slab, that contains several unit cells in [0, y], the wavefield satisfies

[(1/εY001/εX)H]+μZk2H=0,
and (εX, εY, μZ) are the permittivities and permeability relative to the permittivity εext and permeability μext outside the slab (and k2 = ω2εext μext). The wave field inside and outside the slab can be written
{H(y0)=eiksinθx[eikcosθy+Reikcosθy],H(0<yy)=eiksinθx[aeinky+beinky],H(yy)=eiksinθxTeikcosθ(yy)
where nk = kY denotes the horizontal effective wavenumber (n is the index of the effective medium). Applying the continuity, at each interfaces y = 0 and y = y, of the field H and yH/εX (inside the slab) and yH (outside the slab), we get the reflection R and transmission T coefficients
R=(1ξ2)(einkyeinky)(1+ξ)2einky(1ξ)2einky,T=4ξ(1+ξ)2einky(1ξ)2einky
where R = R(θ), T = T (θ) and, for TM waves,
nkYk=μZεXεXεYsin2θ,ξ=εXcosθn.
In [34], the case of TE waves is considered. The authors obtain (with our notations)
n=εZμXμXμYsin2θ,andξ=μXcosθn,
and ξ is used to calculate the reflection coefficient for a single interface Rhs = (ξ − 1)/(ξ + 1), afterwards R is deduced (see Eq. (3) of [34]).

At this stage, let us just make a parenthesis and stress the analogy with the acoustics case:

n=ρYBρYρXsin2θ,andξρYcosθn.
The above expressions are in agreement with the equivalents given in the Table 1.

Next, we come back to the retrieval method. Following [11], it is easy to show that n and ξ can be calculated from (R, T) owing to the relations

{n=iky[log(1R2+T2+ξ˜)2T+2imπ],withξ˜±(1R2+T2)24T2,ξ=ξ˜(1R)2T2,
with m an integer. The determination of the sign in ξ̃ is known as “sign ambiguity”, and the problem of the 2π-multivaluation of n (choice of m) as “branch ambiguities”. These ambiguities have been addressed in [19,28] and they are disregarded here. For weak scattering, these ambiguities are easy to solve. Indeed, we expect in that case R ∼ 0 and Teikcosθℓy, associated to an almost perfect impedance matching ξ ∼ 1 and to almost equal horizontal wavenumbers: nk in the slab and k cosθ outside, namely n ∼ cosθ. Then, ξ̃ = [(1 − R)2T2]ξ ∼ (1 − e2ikcosθℓy) (which answers the sign ambiguity) leading to n ∼ cosθ in Eq. (12) if the principal determination is considered, (branch m = 0).

The steps of the inversion towards the effective parameters are as follows:

  1. n(θ) and ξ (θ) are computed from (R, T) according to Eq. (12).
  2. εX = ξ (θ)n(θ)/cosθ, from Eq. (9), is deduced.
  3. n and εX being known, the function n2/εX = C1C2 sin2 θ can be fitted as a function of θ to get μZ = C1 and εY = 1/C2, Eq. (9).

In the following, we adapt this inversion to the case where α ≠ 0.

3.2. Case of tilted principal axes, α ≠ 0

If α ≠ 0, the wave field H satisfies the wave equation (Eq. (3))

[(1/εy1/εxy1/εxy1/εx)H]+μzk2H=0,
with
{1εx=cos2αεY+sin2αεX,1εy=sin2αεY+cos2αεX,1εxy=sinαcosα(1εX1εY).
The main difference with the previous calculation of R and T is that the horizontal wavenumber of the right- and left- going waves in the slab are not just of opposite sign. Denoting them k±, the field inside the slab can be written
H(0<yy)=eiksinθx[aeik+y+beiky].
Inserting this form in the wave equation (13), we get the dispersion relation
k±2/εy+2k±ksinθ/εxy+k2(sin2θ/εxμZ)=0,
leading to
k±=εyεxyksinθ±nk,withnμZεyεy2εXεYsin2θ,
where we have used 1/(εxεy)1/εxy2=1/(εXεY). At the interfaces y = 0 and y = y, we apply the continuity of the field H and the continuity of yH outside the slab and yH/εy + xH/εxy inside the slab. We get
R=(1ξ2)(einkyeinky)(1+ξ)2einky(1ξ)2einky,T=4ξeikysinθεy/εxy(1+ξ)2einky(1ξ)2einky
and in this case, n and ξ are defined by
nμZεyεy2εXεYsin2θ,ξεycosθn.

Obviously, α = 0 leads to 1/εxy = 0, εx = εY and εy = εX, which is consistent with the previous calculation.

To extend the retrieval method to this case, the main problem is that εxy appears only in the expression of T through a phase term that is unknown a priori. Thus, it will not be possible to build combinations of R and T able to isolate n or ξ only, as it is done for α = 0, Eq. (9). This phase term has to be eliminated first, before any combination of R and T. This is a preliminary step in the inversion, which consists in building a phase compensated transmission coefficient Tc, and this requires T (θ) and T (−θ) to be known. If so, it is sufficient to remark that θ → −θ affects the extra phase term but does not affect ξ and n, and this leads to

e2ikysinθεy/εxy=T(θ)T(θ),
from which we deduce
Tc(θ)T(θ)T(θ)=4ξ(1+ξ)2einky(1ξ)2einky,

Now, the inversion procedure can be apply straightforwardly to (R, Tc) following the same steps as described in Sec. 3.1:

  1. n(θ) and ξ (θ) are computed from (R, Tc) according to Eq. (12)
  2. εy = ξ (θ)n(θ)/cosθ, from Eq. (19), is deduced.
  3. n and εy being known, the function n2/εy = C1C2 sin2 θ can be fitted as a function of θ to get μZ = C1 and εX εY = εy/C2.

At that stage, εy and the product εX εY have been determined and two more steps are needed to get (εxyx):

  1. From Eq. (20), εxy can be calculated, and this may be done by fitting Eq. (20) or by using a direct inversion (εxy = −2ikℓy sinθεy/log[T (θ)/T(−θ)]). Note that a fitting procedure is possible here since εxy is expected to be independent of θ (at least in a certain range of θ), which is not the case for n in Eq. (12).
  2. Finally, εx is determined using
    1εx=εyεXεY+εyεxy2.
    The full tensor can be described equivalently by (εX, εY, α) owing to
    {1εX=12[1εx+1εy+(1εx1εy)2+4εxy2],1εY=12[1εx+1εy(1εx1εy)2+4εxy2],
    where we assume εXεY and
    {α=sin1[1/εY1/εx1/εY1/εX],ifεxy0α=180°sin1[1/εY1/εx1/εY1/εX],ifεxy>0,
    where the determination of α is done in [0, 180°]. The choice to assign the largest permittivity to the axis X in Eq. (23) is not restrictive since it only reflects the equivalence between the situations, for say ε1 < ε2, (εX = ε1, εY = ε2, α) and (εY = ε1, εX = ε2, 90° + α).

In the following sections, we illustrate our retrieval method in two cases: (1) the case of layered medium and we inspect the influence of the anisotropy direction α and of the wavenumber k on the retrieved parameters (electromagnetic wave are considered) and (2) the case of arrays of tilted elliptical particles and we inspect the influence of the array size, beyond the dominant effects of the filling fraction and of the orientation of the particles (acoustic waves are considered).

4. Results for layered dielectric media

The results presented in this section corresponds to a structure made of layers (slab, see Fig. 2). The layers are composed of alternating air and a dielectric material chosen with parameters relative to those of air ε1 = 10 and μ1 = 0.2, with periodicity d and filling fraction φ = 0.5. In the low frequency regime, layered structures are in general well described by homogenization theory [6, 33]; for comparison with the retrieved effective parameters, we give the values of the homogenized parameters

{εXhomog=(1φ+φ/ε1)1=1.818,εYhomog=1φ+φε1=5.500,μhomog=1φ+φμ1=0.6,
with the notation of Eq. (6) (for α = 0, εx = εY and εy = εX). To apply the retrieval method, one consider a slab composed of two unit cells (x = d/cosα, y = 2d/sinα when α ≠ 0 -see Fig. 2 - and x = d, y = 2d when α = 0). Along the x-direction, periodic, or Floquet, boundary conditions are imposed. Numerical calculations are performed using a multimodal method, similar to the Rigorous Coupled Wave Approach [35, 36]. From the numerics, the (R, T) coefficients are calculated, and allow then to deduce the retrieved parameters.

 figure: Fig. 2

Fig. 2 Configuration of the study. The slab contains slanted layers, made of alternating air and a material with parameters relative to those of the air ε1 = 10 and μ1 = 0.2; the wave is incident with angle θ and we measure R and T.

Download Full Size | PDF

4.1. Validation of the retrieval method

To begin with, and to anticipate on the validation of our retrieval method, Figs. 3 and 4 show the comparisons between two wavefields, in the cases α = 0 and α = 45°, respectively: the wavefield computed considering the real structure (direct numerics, Figs. 3(a) and 4(a)), and the wavefield computed in an effective anisotropic slab of length L with parameters deduced from the retrieval method (Figs. 3(b) and 4(b)). For α = 0 and kd = 0.5, we get εX = 1.880, εY = 4.573 and μZ = 0.612, and for α = 45° and kd = 0.25, we get εX = 1.848, εY = 5.014 and μZ = 0.605. Also, in both cases, the value of the angle α is correctly retrieved (values are found to be respectively 0 and 44.62°).

 figure: Fig. 3

Fig. 3 (a) Propagation of an incident plane wave (with θ = 45°) through a slab with kL = 5, containing a layered structure with kd = 0.5 and α = 0 (the layers are represented only on half of the slab to make visible the wavefield inside the slab). (b) Propagation through the slab (kL = 5) filled with the effective anisotropic medium defined by the retrieved parameters (α ≃ 0, εX = 1.880, εY = 4.573 and μZ = 0.612).

Download Full Size | PDF

 figure: Fig. 4

Fig. 4 Same representation as in Fig. 3 for (a) slanted layers in a slab with length kL = 2.5, with kd = 0.25 and α = 45°. (b) The effective parameters derived from the inversion method are α = 44.62°, εX = 1.848, εY = 5.014 and μZ = 0.605.

Download Full Size | PDF

Comparisons are presented for α = 0 and kL = 5 (Fig. 3) and for α = 45° and kL = 2.5 (Fig. 4). As it can be seen, not only the scattering coefficients are correctly restituted (these scattering coefficients give the field outside the slab, and we find few percents of discrepancy on R and T in the presented cases), but also the fields inside the slab. The reader may notice however the scars visible in the wavefield obtained from direct numerics inside the layered structures (the layers are represented on only half of the slab to make visible the wavefield inside the slab).

To be more explicit, we now illustrate the different steps of our retrieval method in the case a = 45°. The preliminary step is illustrated in Fig. 5, for a ≠ 0, where the compensated transmission coefficient Tc is obtained from T (θ) and T (−θ), Eq. (21).

 figure: Fig. 5

Fig. 5 Illustration of the preliminary step of the retrieval method: Calculation of the compensated transmission coefficient Tc, Eq. (21). Here T (θ) and T (−θ) have been calculated for slanted layers with α = 45° and kd = 0.5.

Download Full Size | PDF

Then, Figs. 6 and 7 show the retrieved index n, the impedance ratio ξ (step 1), and the parameter εy that is deduced from the previous two (step 2). In the presented cases, the sign ambiguity has been solved by selecting the sign of ξ̃ (from Eqs. (9) and (12), with cosθ > 0 and assuming εY > 0) in order to impose the real part or the imaginary part of n to be positive; this corresponds to the selection of a right-going wave of the form einky with n either real positive for propagating waves (in the convention eiωt), or imaginary positive for an evanescent wave. Also, for our relatively low contrasts in mass density and bulk modulus, it has been sufficient to consider the principal determination of n in Eq. (12) (branch m = 0) and no branch ambiguity has appeared when varying the parameters θ, α and kd in the considered ranges.

 figure: Fig. 6

Fig. 6 Step 1 of the retrieval method (main step): the impedance ratio ξ (θ) and the effective index (along the x-axis) n(θ) are retrieved, Eq. (12); for kd = 0.01 (○, blue), 0.1 (•, red) and 0.5 (⋄, green).

Download Full Size | PDF

 figure: Fig. 7

Fig. 7 Step 2: εy is deduced from ξ and n; the final corresponding εX values are indicated. Same color and symbol conventions as in Fig. 6 are used.

Download Full Size | PDF

In Fig. 7, it can be observed that the retrieved results for εy calculated at different incidence angles agree well with each other for low kd with a small variation which increases with kd; this is expected since the assumption that the layered medium behaves as an effective medium fails when the wavelength becomes able to inspect the microstructure of size d. To anticipate, the values of εX (obtained after the whole retrieval procedure has been done) is indicated in each case in Fig. 7. As expected, it is angle and frequency independent, and it is found to be close to the homogenized one.

Finally, Fig. 8 illustrates the step 3 of the retrieval method; a fit of the quantity n2/εy is performed to determine μZ and the product εX εY and Fig. 9 illustrates the step 4, where εxy is determined from T (θ)/T (−θ); here, a fit of T (θ)/T (−θ) has been performed in the form of a complex exponential function, according to Eq. (20).

 figure: Fig. 8

Fig. 8 Step 3: Determination of μZ and of the product εX εY by fitting the quantity n2/εy = μZ − sin2 θ εyX εY. Symbols correspond to the retrieved values (with the same color and symbol conventions as in Fig. 6), and plain lines to the fit.

Download Full Size | PDF

 figure: Fig. 9

Fig. 9 Step 4: T (θ)/T (−θ) is fitted to find εxy, Eq. (20).

Download Full Size | PDF

The step 5 is straightforward: εx is deduced from Eq. (22), since εy and the product εX εY have been retrieved. We find εx = 2.660, 2.664 and 2.750 for kd = 0.01, 0.1 and 0.5 respectively.

As previously said, alternatively to the components (εx, εy, εxy) of the permittivity tensor, we can calculate the full set of retrieved parameters (εX, εY, μZ, α), Eqs. (23)(24). They are given below

  • For kd = 0.01, εX = 1.845, εY = 4.944, μZ = 0.607, α = 0.774,
  • For kd = 0.1, εX = 1.846, εY = 4.953, μZ = 0.606, α = 0.775,
  • For kd = 0.5, εX = 1.855, εY = 5.215, μZ = 0.603, α = 0.780.

Although we have not considered the extreme low frequency regime, the retrieved parameters are in quite good agreement with the homogenized parameters, Eq. (25), which are known to properly describe layered structures. This tends to validate our retrieval method. In the following section, we further inspect the retrieval method for varying α and kd.

4.2. Inspection of variations in α and k

Our retrieval method has been applied for varying α values, and kd = 0.01, 0.1 and 0.5. The results are shown in Fig. 10 for the set of parameters (εx, εy, εxy) and μZ. For α > 20°, a slab containing two unit cells y = d/sinα has been considered. However, as α goes to zero, the length of the unit cell diverges; thus, we have imposed a fixed value of y corresponding to α = 20°; this makes the results for α < 20° less reliable since the wave does not capture the periodicity of the structure along the y-axis, particularly for larger wavelength (this is why the results for kd = 0.5 and α < 20° have been omitted in the plots). For comparison, we have reported the behaviors of μZ and of (εx, εy, εxy) coming from homogenization theory, using the values given by Eq. (25) in Eq. (14). It can be seen that the retrieved parameters are close to the homogenized one, which is expected for these layered structures.

 figure: Fig. 10

Fig. 10 Variation of the retrieved parameters with α, for kd = 0.01 (○, blue), kd = 0.1 (•, red) and kd = 0.5 (⋄, green). Gray dotted lines correspond to the homogenized values, Eqs. (25) and (14).

Download Full Size | PDF

Results for varying kd are reported in Fig. 11. To change the representations, we show (εX, εY), the retrieved angle αr, and μZ, and corresponding homogenized values are reported. Again, the retrieval values are close to the homogenized ones, with significant variations (sometime outside the plot) for kd approaching unity.

 figure: Fig. 11

Fig. 11 Variation of the retrieved parameters (as defined in Eq. (6)) as a function of kd, for α = 0 (○, blue) and α = 45° (⋄, green). Gray dotted lines correspond to the homogenized values, Eqs. (25).

Download Full Size | PDF

5. Results for tilted elliptical particles and acoustic waves

As presented in Section 2, the case of acoustic waves is similar to the case of electromagnetic waves. In this section, we consider periodic arrangements of elliptical particles with relative (with respect to the surrounding medium) mass density ρ1 = 100 and relative bulk modulus B1 = 200. The particles are on a square lattice x-periodic, and they are defined by the major and minor diameters a = 0.8x and b = 0.2x; being the case, they can be tilted by an angle α inside the unit cell (Figures 12).

 figure: Fig. 12

Fig. 12 Square lattices of tilted elliptical particles; the unit cell is square x × x and (a, b) are the major and minor diameters. Note that tilting the ellipse inside the unit cell produces an array (b) with arrangement different from tilting the whole array (a). The surrounding medium has normalized mass density and bulk modulus; the elliptical particles have relative mass density ρ1 and relative bulk modulus B1.

Download Full Size | PDF

As first attempts, we consider the array of the Figure 12(b) with α = 45°. The retrieval method is performed starting from (R, T) calculated for y = 4x and kℓx = 10−2; we find αr = 44.4° (close to value of the tilt angle α = 45°) and

ρX=1.171,ρY=1.785,B=1.138.

These values are consistent with the results available in the literature for this configuration. Indeed, B is close to Bth, the volume average of the bulk moduli in the unit cell; with φ=πab/4x2 the filling fraction of the particles, we have

Bth=[φ+B1+1φ]1=1.143,
often refereed as Wood’s law, see, e.g., [37]. For the effective density tensor, a simple expressions of (ρX, ρY) is given in the dilute limit (small φ-values, and φ ≃ 0.126 in our case) [38]. This derivation is performed for electromagnetic waves and for an array corresponding to α = 0 (Eq. (38) in this reference); we give below the acoustic equivalent (accounting for the non convenient equivalences εXρY and εYρX):
{ρXth=2rXφ+ρ1(rX+φ)2rX+φ+ρ1(rXφ)1.168,ρYth=2rYφ+ρ1(rY+φ)2rY+φ+ρ1(rYφ)1.853,
where we have defined rX ≡ 2a/(a + b) and rY ≡ 2b/(a + b). The agreement with our retrieval values is reasonable (with differences less than 4%).

To get insight of the robustness of the retrieval parameters, we use the values of (ρX, ρY, B) (in Eq. (26)) to characterize the effective behavior of arrays at significant higher frequency kℓx = 0.25, for a larger slab L = 50x and for three α values (α = 0, 45°, 90°). Figures 13 show the fields calculated for an incident wave propagating through the real lattices, compared with the fields calculated for an incident wave propagating through the slabs filled with the effective anisotropic media; we used the effective parameters previously determined, Eq. (26) and the direction of anisotropy is fixed by the tilt angle α. The agreement is good in the three cases.

 figure: Fig. 13

Fig. 13 Top panel: Propagation of an incident wave (with θ = 45°) through a slab with length L = 50x containing tilted elliptical particles in a square lattice with unit cell of size kℓx = 0.25 (the ellipse is given by a = 0.8x, b = 0.2x). The particles are materialized with black line only on the lower part of the slab; the insets show a single particle in the unit cell. Bottom : Propagation through the slab of same length filled with the effective anisotropic medium defined using the retrieved parameters (ρX = 1.171, ρY = 1.785, B = 1.138), and (a) α = 0, (b) α = 45°, α = 90°.

Download Full Size | PDF

The observed agreement, although intuitive, is not so trivial : indeed, the particles have been tilted by α inside the unit cell, but the lattice remains a square lattice; this is not the case if the whole array (at α = 0) is tilted of α; for instance, rotating the whole array of angle α = 45° would result in a hexagonal lattice while it remains a square lattice in our case. This suggests that the composition (ρ1, B1), the filling fraction φ and the orientations α of the particles are the pertinent parameters which characterize the effective medium, at least at leading order (as in the theoretical expressions Eqs. (27) and (28)).

To be more quantitative, we report in Figure 14 the variations of the retrieved parameters (ρX, ρY, B, αr) as a function of α; the retrieval method has been applied for kℓx = 10−2 and with y = Nℓx, N = 5, 15 and 30.

 figure: Fig. 14

Fig. 14 Variation of the retrieved parameters with α for N = 5 (○, blue), N = 15 (•, red) and N = 30 (⋄, green). Dotted gray lines are the theoretical predictions, Eqs. (27) and (28).

Download Full Size | PDF

As a first comment, the retrieved parameters (ρX, ρY, B) are found to be rather constant and the retrieved angle of anisotropic αr is found to coincide with the tilt angle α within 2% for any N. This is consistent with the conclusions drawn from the fields in Figure 13, that the effective medium is not very sensitive to the type of lattice. Note also that (ρX, ρY) satisfy the symmetry α → (180° − α), as expected.

Then, two facts are remarkable. (1) The amplitude in the variations of (ρX, ρY) does not decrease with N, and the amplitude is quite significant for ρX (10 %). We also checked that it does not decrease for lower frequencies. We think that this variation may be attributable to the change in the lattice configurations when α varies, as previously commented. (2) An influence of N, the size of the array chosen for the inversion, has visible consequence. On the one hand, the retrieved bulk modulus B slightly increases with N. On the other hand, for ρX and ρY, the symmetry α → (90° − α) is only approached for increasing N values. These two facts illustrate the “size dependance” of the effective anisotropic medium, which is a second order effect, compared with the dominant order (influence of (ρ1, B1), φ). Indeed the symmetry α → (90° − α) is expected for infinite lattices; for small N, it turns out that the effective parameters of an array with elongated particles orientated along x (square lattice with α = 0) or orientated along y (square lattice with α = 90°) are not identical. This is illustrated further in the last Figure 15 where we report the variations of the effective parameters with the array size N for α = 0 and α = 90°. While αr is found independent of N (within 2% variation), the retrieved parameters ρX (and ρY) for α = 0 and 90° tend to a commun value only asymptotically for large N.

 figure: Fig. 15

Fig. 15 Variation of the retrieved parameters with the array size N (y = Nℓx) for α = 0 (○, blue) and α = 90° (⋄, green). The dotted gray lines indicate the theoretical predictions, Eqs. (27) and (28).).

Download Full Size | PDF

6. Concluding remarks

We have proposed an extension of the well-known parameter retrieval method toward oblique incidence and arbitrary anisotropy directions of the effective medium. The retrieval method has been validated in a simple case of slanted layer structures, where the homogenization theory of layered medium is known to produce accurate results in the low frequency regime. Consistent results have been found for the retrieved parameters. We also reported results for tilted elliptical particles in the acoustic case; again, the retrieved parameters were consistent with available results of the literature, with dominant effects being due to the composition, the filling fraction and the orientation of the particles; beyond these dominant effects, the weak influence of the size of the medium has been evidenced.

However, in both cases, the effective parameters have been shown to be roughly independent on the characteristics of the incident wave (providing that the assumption of low frequency regime is reasonably satisfied). In many cases of interest, this is not the case; complex resonant structures are at least dependent on the frequency and may even depend on the incidence angle, and, in the worth cases, on the sample shape and size [23]. In these cases, the concept of effective medium becomes questionable or meaningless; enlightening critical discussions on this problem can be found in [21, 24, 25]. It is stressed in [25] that the retrieved parameters have to be understood rather as wave parameters able to restitute the scattering properties of the structure than as material parameters, independently of the scattering problem. Obvisouly, the interest in determining these effective parameters is to avoid heavy computations of a real macrostructure including its microstructure. Note that this implies at least the wave parameters to be independent on the size of the structures. With respect to this goal, it is of interest to introduce a conceptual homogeneous medium, associated to a known physics (its response with respect to wave propagation) and because the structures are involved, it is of interest that this medium possesses many degrees of freedom. This is the reason for previous improvements of the retrieval method and it is the meaning that we attribute to the extra degree of freedom considered in the present study (the angle α of the principal direction of anisotropy). Adding more and more degrees of freedom (within a prescribed known wave physics) should help to isolate more material parameters and less wave parameters. At least, this is what we can expect from such improved models.

Acknowledgments

The authors acknowledge the financial support of the Agence Nationale de la Recherche through the grant ANR ProCoMedia, project ANR-10-INTB-0914. A. C. and J-F. M. acknowledge the funding from the European Community’s Seventh Framework Program (FP7/2007–2013) under grant agreement no 285549: SIMPOSIUM project.

References and links

1. N. Engheta and R. W. Ziolkowski, eds. Metamaterials: Physics and Engineering Explorations (John Wiley & Sons, 2006). [CrossRef]  

2. L. Solymar and E. Shamonina, Waves in Metamaterials (Oxford University, Oxford, 2009).

3. R. V. Craster and S. Guenneau, eds. Acoustic metamaterials: Negative Refraction, Imaging, Lensing and Cloaking, Vol. 166 (Springer, 2012).

4. J. Sánchez-Dehesa, D. Torrent, and J. Carbonell, “Anisotropic metamaterials as sensing devices in acoustics and electromagnetism,” Proc. of SPIE 8346, 834606 (2012). [CrossRef]  

5. C.P. Berraquero, A. Maurel, P. Petitjeans, and V. Pagneux, “Experimental realization of a water-wave metamaterial shifter,” Phys. Rev. E 88, 051002(R) (2013). [CrossRef]  

6. O. A. Oleinik, A. S. Shamaev, and G. A. Yosifian, Mathematical Problems in Elasticity and Homogenization (North Holland, Amsterdam, 1992).

7. T. Antonakakis, R. V. Craster, and S. Guenneau, “High-frequency homogenization of zero-frequency stop band photonic and phononic crystals,” New J. Phys. 15, 103014 (2013). [CrossRef]  

8. P. Markos and C. M. Soukoulis, “Transmission properties and effective electromagnetic parameters of double negative metamaterials,” Opt. Express 11, 649–661 (2003). [CrossRef]   [PubMed]  

9. D. R. Smith, D. C. Vier, T. Koschny, and C. M. Soukoulis, “Electromagnetic parameter retrieval from inhomogeneous metamaterials,” Phys. Rev. E 71, 036617 (2005). [CrossRef]  

10. X. Chen, T. M. Grzegorczyk, B. Wu, J. Pacheco, and J. A. Kong, “Robust method to retrieve the constitutive effective parameters of metamaterials,” Phys. Rev. E., 70, 016608 (2004). [CrossRef]  

11. V. Fokin, M. Ambati, C. Sun, and X. Zhang, “Method for retrieving effective properties of locally resonant acoustic metamaterials,” Phys. Rev. B 76, 144302 (2007). [CrossRef]  

12. D. R. Smith, S. Schultz, P. Markos, and C. M. Soukoulis, “Determination of effective permittivity and permeability of metamaterials from reflection and transmission coefficients,” Phys. Rev. B 65, 195104 (2002). [CrossRef]  

13. H. Chen, J. Zhang, Y. Bai, Y. Luo, L. Ran, Q. Jiang, and J. A. Kong, “Experimental retrieval of the effective parameters of metamaterials based on a waveguide method,” Opt. Express 14, 12944–12949 (2006). [CrossRef]   [PubMed]  

14. Z. Liang and J. Li, “Extreme acoustic metamaterial by coiling up space,” Phys. Rev. Lett. 108, 114301 (2012). [CrossRef]   [PubMed]  

15. Y. Xie, B.-I. Popa, L. Zigoneanu, and S. A. Cummer, “Measurement of a broadband negative index with space-coiling acoustic metamaterials,” Phys. Rev. Lett. 110, 175501 (2013). [CrossRef]   [PubMed]  

16. B.-I. Popa and S. A. Cummer, “Design and characterization of broadband acoustic composite metamaterials,” Phys. Rev. B 80, 174303 (2009). [CrossRef]  

17. L. Zigoneanu, B.-I. Popa, and S. A. Cummer, “Design and measurements of a broadband two-dimensional acoustic lens,” Phys. Rev. B 84, 024305 (2011). [CrossRef]  

18. L. Zigoneanu, B.-I. Popa, A.F. Starr, and S. A. Cummer, “Design and measurements of a broadband two-dimensional acoustic metamaterial with anisotropic effective mass density,” J. Appl. Phys. 109, 054906 (2011). [CrossRef]  

19. Y. Hao and R. Mittra, FDTD Modeling of Metamaterials: Theory and Applications, Chap. 7, (Artech House, Boston, 2009).

20. C. Menzel, T. Paul, C. Rockstuhl, T. Pertsch, S. Tretyakov, and F. Lederer, “Validity of effective material parameters for optical fishnet metamaterials,” Phys. Rev B, 81, 035320 (2010). [CrossRef]  

21. S. B. Raghunathan and N. V. Budko, “Effective permittivity of finite inhomogeneous objects,” Phys. Rev. B 81, 054206 (2010). [CrossRef]  

22. C. R. Simovski and S. A. Tretyakov, “On effective electromagnetic parameters of artificial nanostructured magnetic materials,” Photonics and Nanostructures - Fundamentals and Applications 8, 254–263 (2010). [CrossRef]  

23. A. I. Căbuz, A. Nicolet, F. Zolla, D. Felbacq, and G. Bouchitté, “Homogenization of nonlocal wire metamaterial via a renormalization approach,” J. Opt. Soc. Am. B, 28, 1275–1282 (2011). [CrossRef]  

24. C. R. Simovski, “On electromagnetic characterization and homogenization of nanostructured metamaterials,” J. Opt. 13, 013001 (2011). [CrossRef]  

25. C. Menzel, C. Rockstuhl, T. Paul, F. Lederer, and T. Pertsch, “Retrieving effective parameters for metamaterials at oblique incidence,” Phys. Rev. B 77, 195328 (2008). [CrossRef]  

26. F.-J. Hsieh and W.-C. Wang, “Full extraction methods to retrieve effective refractive index and parameters of a bianisotropic metamaterial based on material dispersion models,” J. Appl. Phys. 112, 064907 (2012) [CrossRef]  

27. Z. H. Jiang, J. A. Bossard, X. Wang, and D. H. Werner, “Synthesizing metamaterials with angularly independent effective medium properties based on an anisotropic parameter retrieval technique coupled with a genetic algorithm,” J. Appl. Phys. 109, 013515 (2011) [CrossRef]  

28. S. Arslanagić, T. V. Hansen, N. A. Mortensen, A. H. Gregersen, O. Sigmund, R. W. Ziolkowski, and O. Breinbjerg, “A review of the scattering-parameter extraction method with clarification of ambiguity issues in relation to metamaterial homogenization,” IEEE Antennas and Propagation Magazine 55, 91–106 (2013). [CrossRef]  

29. T.-C. Yang, Y.-H. Yang, and T.-J. Yen, “An anisotropic negative refractive index medium operated at multiple-angle incidences,” Opt. Express 17, 24189–24197 (2009). [CrossRef]  

30. B. L. Wu, W. J. Wang, J. Pacheco, X. Chen, J. Lu, T. M. Grzegorczyk, J. A. Kong, P. Kao, P. A. Theophelakes, and M. J. Hogan, “Anisotropic metamaterials as antenna substrate to enhance directivity,” Microw. and Opt. Techn. Lett. 48, 680–683 (2006). [CrossRef]  

31. Y. Yuan, L. Shen, L. Ran, T. Jiang, J. Huangfu, and J. A. Kong, “Directive emission based on anisotropic meta-materials,” Phys. Rev. A 77, 053821 (2008). [CrossRef]  

32. T. Jiang, Y. Luo, Z. Wang, L. Peng, J. Huangfu, W. Cui, W. Ma, H. Chen, and L. Ran, “Rainbow-like radiation from an omni-directional source placed in a uniaxial metamaterial slab,” Opt. Express 17, 7068–7073 (2009). [CrossRef]   [PubMed]  

33. A. Maurel, S. Félix, and J.-F. Mercier, “Enhanced transmission through gratings: Structural and geometrical effects,” Phys. Rev. B 88, 115416 (2013). [CrossRef]  

34. T. M. Grzegorczyk, Z. M. Thomas, and J. A. Kong, “Inversion of critical angle and Brewster angle in anisotropic left-handed metamaterials,” Appl. Phys. Lett. 86, 251909 (2005). [CrossRef]  

35. A Maurel, J.-F. Mercier, and S. Félix, “Wave propagation through penetrable scatterers in a waveguide and though a penetrable grating,” J. Acoust. Soc. Am. 135, 165–174 (2014). [CrossRef]   [PubMed]  

36. E. Popov and M. Nevière, “Grating theory: New equations in Fourier space leading to fast converging results for TM polarization,” J. Opt. Soc. Am. A 17, 1773–1784 (2000). [CrossRef]  

37. D. Torrent and J. Sánchez-Dehesa, “Effective parameters of clusters of cylinders embedded in a nonviscous fluid or gas,” Phys. Rev. B 74, 224305 (2006). [CrossRef]  

38. N. A. Nicorovici and R. C. McPhedran, “Transport properties of arrays of elliptical cylinders,” Phys. Rev. E 54(2), 1945–1957 (1996). [CrossRef]  

Cited By

Optica participates in Crossref's Cited-By Linking service. Citing articles from Optica Publishing Group journals and other participating publishers are listed here.

Alert me when this article is cited.


Figures (15)

Fig. 1
Fig. 1 (a) Typical configuration of the real medium, the reflection R and transmission T through a slab containing few cells (otherwise x-periodic in the x-direction) are calculated for an obliquely incident plane wave with angle θ and wavenumber k. (b) Effective anisotropic medium leading to the same scattering properties (R and T).
Fig. 2
Fig. 2 Configuration of the study. The slab contains slanted layers, made of alternating air and a material with parameters relative to those of the air ε1 = 10 and μ1 = 0.2; the wave is incident with angle θ and we measure R and T.
Fig. 3
Fig. 3 (a) Propagation of an incident plane wave (with θ = 45°) through a slab with kL = 5, containing a layered structure with kd = 0.5 and α = 0 (the layers are represented only on half of the slab to make visible the wavefield inside the slab). (b) Propagation through the slab (kL = 5) filled with the effective anisotropic medium defined by the retrieved parameters (α ≃ 0, εX = 1.880, εY = 4.573 and μZ = 0.612).
Fig. 4
Fig. 4 Same representation as in Fig. 3 for (a) slanted layers in a slab with length kL = 2.5, with kd = 0.25 and α = 45°. (b) The effective parameters derived from the inversion method are α = 44.62°, εX = 1.848, εY = 5.014 and μZ = 0.605.
Fig. 5
Fig. 5 Illustration of the preliminary step of the retrieval method: Calculation of the compensated transmission coefficient Tc, Eq. (21). Here T (θ) and T (−θ) have been calculated for slanted layers with α = 45° and kd = 0.5.
Fig. 6
Fig. 6 Step 1 of the retrieval method (main step): the impedance ratio ξ (θ) and the effective index (along the x-axis) n(θ) are retrieved, Eq. (12); for kd = 0.01 (○, blue), 0.1 (•, red) and 0.5 (⋄, green).
Fig. 7
Fig. 7 Step 2: εy is deduced from ξ and n; the final corresponding εX values are indicated. Same color and symbol conventions as in Fig. 6 are used.
Fig. 8
Fig. 8 Step 3: Determination of μZ and of the product εX εY by fitting the quantity n2/εy = μZ − sin2 θ εyX εY. Symbols correspond to the retrieved values (with the same color and symbol conventions as in Fig. 6), and plain lines to the fit.
Fig. 9
Fig. 9 Step 4: T (θ)/T (−θ) is fitted to find εxy, Eq. (20).
Fig. 10
Fig. 10 Variation of the retrieved parameters with α, for kd = 0.01 (○, blue), kd = 0.1 (•, red) and kd = 0.5 (⋄, green). Gray dotted lines correspond to the homogenized values, Eqs. (25) and (14).
Fig. 11
Fig. 11 Variation of the retrieved parameters (as defined in Eq. (6)) as a function of kd, for α = 0 (○, blue) and α = 45° (⋄, green). Gray dotted lines correspond to the homogenized values, Eqs. (25).
Fig. 12
Fig. 12 Square lattices of tilted elliptical particles; the unit cell is square x × x and (a, b) are the major and minor diameters. Note that tilting the ellipse inside the unit cell produces an array (b) with arrangement different from tilting the whole array (a). The surrounding medium has normalized mass density and bulk modulus; the elliptical particles have relative mass density ρ1 and relative bulk modulus B1.
Fig. 13
Fig. 13 Top panel: Propagation of an incident wave (with θ = 45°) through a slab with length L = 50x containing tilted elliptical particles in a square lattice with unit cell of size kℓx = 0.25 (the ellipse is given by a = 0.8x, b = 0.2x). The particles are materialized with black line only on the lower part of the slab; the insets show a single particle in the unit cell. Bottom : Propagation through the slab of same length filled with the effective anisotropic medium defined using the retrieved parameters (ρX = 1.171, ρY = 1.785, B = 1.138), and (a) α = 0, (b) α = 45°, α = 90°.
Fig. 14
Fig. 14 Variation of the retrieved parameters with α for N = 5 (○, blue), N = 15 (•, red) and N = 30 (⋄, green). Dotted gray lines are the theoretical predictions, Eqs. (27) and (28).
Fig. 15
Fig. 15 Variation of the retrieved parameters with the array size N (y = Nℓx) for α = 0 (○, blue) and α = 90° (⋄, green). The dotted gray lines indicate the theoretical predictions, Eqs. (27) and (28).).

Tables (1)

Tables Icon

Table 1 Coefficients in the 2D wave equation for anisotropic media in electromagnetism and acoustics.

Equations (28)

Equations on this page are rendered with MathJax. Learn more.

ε diag = ( ε X 0 0 0 ε Y 0 0 0 ε Z ) , μ diag = ( μ X 0 0 0 μ Y 0 0 0 μ Z ) , in electromagnetism , ρ diag = ( ρ X 0 0 0 ρ Y 0 0 0 ρ Z ) , in acoustics .
R = ( cos α sin α 0 sin α cos α 0 0 0 1 )
in electromagnetism { × E = i k ( R t μ diag R ) H , × H = i k ( R t μ diag R ) E , in acoustics { p = i k ( R t ρ diag R ) u , B u = i k p ,
[ ( 1 / ε x 1 / ε x y 1 / ε x y 1 / ε y ) H ] + μ Z k 2 H = 0 , TM waves , [ ( 1 / μ x 1 / μ x y 1 / μ x y 1 / μ y ) E ] + ε Z k 2 E = 0 , TE waves .
[ ( 1 / ρ x 1 / ρ x y 1 / ρ x 1 / ρ y ) p ] + k 2 B p = 0 ,
[ ( 1 / ε Y 0 0 1 / ε X ) H ] + μ Z k 2 H = 0 ,
{ H ( y 0 ) = e i k sin θ x [ e i k cos θ y + R e i k cos θ y ] , H ( 0 < y y ) = e i k sin θ x [ a e i n k y + b e i n k y ] , H ( y y ) = e i k sin θ x T e i k cos θ ( y y )
R = ( 1 ξ 2 ) ( e i n k y e i n k y ) ( 1 + ξ ) 2 e i n k y ( 1 ξ ) 2 e i n k y , T = 4 ξ ( 1 + ξ ) 2 e i n k y ( 1 ξ ) 2 e i n k y
n k Y k = μ Z ε X ε X ε Y sin 2 θ , ξ = ε X cos θ n .
n = ε Z μ X μ X μ Y sin 2 θ , and ξ = μ X cos θ n ,
n = ρ Y B ρ Y ρ X sin 2 θ , and ξ ρ Y cos θ n .
{ n = i k y [ log ( 1 R 2 + T 2 + ξ ˜ ) 2 T + 2 i m π ] , with ξ ˜ ± ( 1 R 2 + T 2 ) 2 4 T 2 , ξ = ξ ˜ ( 1 R ) 2 T 2 ,
[ ( 1 / ε y 1 / ε x y 1 / ε x y 1 / ε x ) H ] + μ z k 2 H = 0 ,
{ 1 ε x = cos 2 α ε Y + sin 2 α ε X , 1 ε y = sin 2 α ε Y + cos 2 α ε X , 1 ε x y = sin α cos α ( 1 ε X 1 ε Y ) .
H ( 0 < y y ) = e i k sin θ x [ a e i k + y + b e i k y ] .
k ± 2 / ε y + 2 k ± k sin θ / ε x y + k 2 ( sin 2 θ / ε x μ Z ) = 0 ,
k ± = ε y ε x y k sin θ ± n k , with n μ Z ε y ε y 2 ε X ε Y sin 2 θ ,
R = ( 1 ξ 2 ) ( e i n k y e i n k y ) ( 1 + ξ ) 2 e i n k y ( 1 ξ ) 2 e i n k y , T = 4 ξ e i k y sin θ ε y / ε x y ( 1 + ξ ) 2 e i n k y ( 1 ξ ) 2 e i n k y
n μ Z ε y ε y 2 ε X ε Y sin 2 θ , ξ ε y cos θ n .
e 2 i k y sin θ ε y / ε x y = T ( θ ) T ( θ ) ,
T c ( θ ) T ( θ ) T ( θ ) = 4 ξ ( 1 + ξ ) 2 e i n k y ( 1 ξ ) 2 e i n k y ,
1 ε x = ε y ε X ε Y + ε y ε x y 2 .
{ 1 ε X = 1 2 [ 1 ε x + 1 ε y + ( 1 ε x 1 ε y ) 2 + 4 ε x y 2 ] , 1 ε Y = 1 2 [ 1 ε x + 1 ε y ( 1 ε x 1 ε y ) 2 + 4 ε x y 2 ] ,
{ α = sin 1 [ 1 / ε Y 1 / ε x 1 / ε Y 1 / ε X ] , if ε x y 0 α = 180 ° sin 1 [ 1 / ε Y 1 / ε x 1 / ε Y 1 / ε X ] , if ε x y > 0 ,
{ ε X homog = ( 1 φ + φ / ε 1 ) 1 = 1.818 , ε Y homog = 1 φ + φ ε 1 = 5.500 , μ homog = 1 φ + φ μ 1 = 0.6 ,
ρ X = 1.171 , ρ Y = 1.785 , B = 1.138 .
B th = [ φ + B 1 + 1 φ ] 1 = 1.143 ,
{ ρ X th = 2 r X φ + ρ 1 ( r X + φ ) 2 r X + φ + ρ 1 ( r X φ ) 1.168 , ρ Y th = 2 r Y φ + ρ 1 ( r Y + φ ) 2 r Y + φ + ρ 1 ( r Y φ ) 1.853 ,
Select as filters


Select Topics Cancel
© Copyright 2024 | Optica Publishing Group. All rights reserved, including rights for text and data mining and training of artificial technologies or similar technologies.